Next Article in Journal
Tauroursodeoxycholic Acid Enhances Osteogenic Differentiation through EGFR/p-Akt/CREB1 Pathway in Mesenchymal Stem Cells
Previous Article in Journal
Impact of Short-Term (+)-JQ1 Exposure on Mouse Aorta: Unanticipated Inhibition of Smooth Muscle Contractility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epidrugs in the Therapy of Central Nervous System Disorders: A Way to Drive on?

by
Marina G. Gladkova
1,*,
Este Leidmaa
2,3 and
Elmira A. Anderzhanova
4,†
1
Faculty of Bioengineering and Bioinformatics, Lomonosov Moscow State University, 119234 Moscow, Russia
2
Institute of Molecular Psychiatry, Medical Faculty, University of Bonn, 53127 Bonn, Germany
3
Institute of Biomedicine and Translational Medicine, Department of Physiology, University of Tartu, 50411 Tartu, Estonia
4
School of Medicine, BAU International University Batumi, Batumi 6010, Georgia
*
Author to whom correspondence should be addressed.
Present address: Federal State Budgetary Institution “Centre for Strategic Planning and Management of Biomedical Health Risks” of the Federal Medical Biological Agency, Pogodinskaya 10, 119121 Moscow, Russia.
Cells 2023, 12(11), 1464; https://doi.org/10.3390/cells12111464
Submission received: 22 January 2023 / Revised: 1 May 2023 / Accepted: 16 May 2023 / Published: 24 May 2023

Abstract

:
The polygenic nature of neurological and psychiatric syndromes and the significant impact of environmental factors on the underlying developmental, homeostatic, and neuroplastic mechanisms suggest that an efficient therapy for these disorders should be a complex one. Pharmacological interventions with drugs selectively influencing the epigenetic landscape (epidrugs) allow one to hit multiple targets, therefore, assumably addressing a wide spectrum of genetic and environmental mechanisms of central nervous system (CNS) disorders. The aim of this review is to understand what fundamental pathological mechanisms would be optimal to target with epidrugs in the treatment of neurological or psychiatric complications. To date, the use of histone deacetylases and DNA methyltransferase inhibitors (HDACis and DNMTis) in the clinic is focused on the treatment of neoplasms (mainly of a glial origin) and is based on the cytostatic and cytotoxic actions of these compounds. Preclinical data show that besides this activity, inhibitors of histone deacetylases, DNA methyltransferases, bromodomains, and ten-eleven translocation (TET) proteins impact the expression of neuroimmune inflammation mediators (cytokines and pro-apoptotic factors), neurotrophins (brain-derived neurotropic factor (BDNF) and nerve growth factor (NGF)), ion channels, ionotropic receptors, as well as pathoproteins (β-amyloid, tau protein, and α-synuclein). Based on this profile of activities, epidrugs may be favorable as a treatment for neurodegenerative diseases. For the treatment of neurodevelopmental disorders, drug addiction, as well as anxiety disorders, depression, schizophrenia, and epilepsy, contemporary epidrugs still require further development concerning a tuning of pharmacological effects, reduction in toxicity, and development of efficient treatment protocols. A promising strategy to further clarify the potential targets of epidrugs as therapeutic means to cure neurological and psychiatric syndromes is the profiling of the epigenetic mechanisms, which have evolved upon actions of complex physiological lifestyle factors, such as diet and physical exercise, and which are effective in the management of neurodegenerative diseases and dementia.

1. Epigenetics and Approaches to Its Study

The deoxyribonucleic acid (DNA) molecule encoding for proteins is essential for life and is predominantly responsible for immutable traits. The DNA in a cell exists in a strictly structured form that is achieved with the help of histones and architectural proteins. This structural organization defines the differential functional activity of the DNA molecule, which can be transcriptionally inaccessible in dense regions (heterochromatin) or transcriptionally accessible (euchromatin) in loose regions of chromatin (Figure 1).
Epigenetics, on the contrary, gives a living creature the opportunity to adapt to dynamically changing environmental factors, hence, playing an important role in the balance of molecular processes, which, if violated, can lead to various diseases. An epigenetic phenomenon is a change in gene expression that is not mediated by changes in the nucleotide sequence of a DNA molecule. The alterations in DNA accessibility and transcriptional activity are mostly achieved by a few mechanisms: (1) chemical modifications (reversible methylation) of DNA, (2) chemical modifications (e.g., reversible acetylation or methylation) of histone proteins, (3) chromatin remodeling and the spatial three-dimensional (3D) organization of the genome, as well as via action of (4) non-coding ribonucleic acids (ncRNAs) on gene expression. At the cellular and systemic levels, epigenetic mechanisms controlling gene activity influence molecular, biochemical, and biological processes. Epigenetic control over the genome is essential in determining the unique properties of a cell, which develops from a single precursor via specialized gene expression programs.
The upper-mentioned epigenetic functional and structural alterations normally appear in a given organism but may be transmitted to the next generation(s) due to the stabilization of the specific expression programs (with no changes in the genome sequence per se), for instance, via the preservation of the residual methylation after fertilization and via gamete-delivered RNAs [4,5].
Epigenetic mechanisms of all known types can be reliably monitored nowadays. For instance, histone modifications are well recognized as an integrative marker of genome regulation. Chromatin immunoprecipitation (ChIP)-sequencing is commonly used to assess the profile of histone lysine (K) acetylation and methylation. The most validated specific markers of genome suppression are histone H3 methylation at lysine in two positions: H3K9me3 and H3K27me3. In turn, H3K27ac, H3K4me3, and H3K4me1 are validated as genome activation factors. However, to date, a variety of other post-translational histone2-4 modifications, such as arginine acetylation, phosphorylation, SUMOylation, and ubiquitination, are shown to result in changes in genome activity [6].
Table 1 summarizes the contemporary methods of epigenomics, including epigenome next-generation sequencing (NGS), which are used to analyze the epigenetic component of cell activity.
The sensitivity and efficiency of modern analytical methods allow one to assess the epigenetic landscape and reliably identify the promoter, enhancer, as well as associated insulator sequences at the single-cell level [7,8,9].
Progress in entire genome chromatin profiling and epigenome-wide association studies (EWASs) (by analogy with a genome-wide association study, GWAS) fostered a surge of data about epigenetic states in various tissues and cells at their normal and pathological functionings. The contemporary challenges of epigenetic data management are the (1) development of open-access databases; (2) standardisation of experimental methods; (3) setting of standards for the analysis and integration of EWAS and GWAS data; (4) annotation of epigenetic markers at different levels of biological complexity; and (5) selection of a significant set of epigenetic markers in the context of a disease or treatment [8,10,11,12,13,14].
Aside from the general constraints related to big data management, there are a few additional obstacles to understanding the role of epigenetic mechanisms in the pathogenesis of neurological and psychiatric diseases. Although a big pool of information on changes in the epigenome is amassed to date, we are still limited in detailed aspects of the dynamics of epigenetic landscapes in different neuronal cells in various brain regions. Similarly, a systematic search for reliable peripheral epigenetic markers of central nervous system (CNS) disorders has still not been accomplished. The rationale for using peripheral epigenetic markers, in addition to their easy accessibility in humans, is the involvement of peripheral systems (namely, the immune system, circulating stress hormones, myokines, and the microbiome) in the normal maintenance of brain functions and the generation of pathological endophenotypes. One of the possible platforms for finding significant correlations can be the concept of a so-called epigenetic “clock”, which assumes the age-related accumulation of methyl groups on a DNA molecule, allowing one to determine the biological age of cells, tissues, organs, or organisms. The best-validated model is Steve Horvath’s epigenetic “clock”, comprising 353 epigenetic markers of the human genome [15,16].

2. Epigenetic Therapy of Diseases of the CNS

2.1. Epigenetically Active Drugs (Epidrugs)

At present, the field of epidrug pharmacology is developing to serve mainly the needs of cancer therapy. However, epigenetically active substances, being, by default, multitarget and multifunctional, would represent the cluster of drugs of which specific and unspecific effects are potentially even of higher interest with respect to other nosologies or syndromes [17]. The known compounds that are able to interfere with epigenetic mechanisms are represented by drugs with a primary epigenetic activity or by those that possess epigenetic effects as a side phenomenon. Table 2 lists existing epigenetically active substances (epidrugs) in accordance with the biochemical mechanism [18,19,20]. To date, the compounds of three types have been studied the most: (1) inhibitors of DNA methyltransferases (DNMTis); (2) inhibitors of histone deacetylases (HATis); and (3) inhibitors of DNA demethylases (TETis). Several compounds with primary epigenetic activity have been approved as anticancer therapy by the EMA (European Medicines Agency) and/or the FDA (Food and Drug Administration) of the USA, as well as the FDA of the People’s Republic of China (e.g., chidamide).
Table 2. Clusters of epidrugs and compounds exerting epigenetic activities. Approved anticancer compounds with primary epigenetic activity are in bold; subclasses of inhibitors are italicized.
Table 2. Clusters of epidrugs and compounds exerting epigenetic activities. Approved anticancer compounds with primary epigenetic activity are in bold; subclasses of inhibitors are italicized.
Epigenetic Modification/
Pathway (Figure 2)
Target Domain and
Mechanism of Action
AbbreviationExamples of Approved
Compounds
Primary Target
Pathology
DNA methylationDNA
methyltransferase
inhibitors
DNMTis5-azacytidine, decitabine [21],
hydralazine [22], and
procainamide [23]
myeloid neoplasms,
malaria, and cardiovascular diseases
DNA demethylationTen-eleven
translocation (TET)
proteins inhibitors
TETis-
Histone acetylationHistone
acetyltransferase
inhibitors
HATisanacardic acid, curcumin,
garcinol, catechin, and thiazole + bistubstrate
inhibitors [24]
antimicrobial therapy, anti-inflammatory
therapy, and cancer
Histone deacetylationHistone deacetylases
inhibitors
HDACisvorinostat,
panobinostat
(withdrawn by FDA in 2022),
romidepsin
(withdrawn by FDA in 2021), belinostat,
chidamide (tucidinostat) [25],
entinostat
(granted breakthrough therapy status by the FDA in 2013),
valproic acid, magnesium salt of valproic acid,
phenylbutyric acid [2],
nicotinamide
(affirmed as GRAS (Generally Recognized as Safe) by the
FDA in 2005 as a direct
human food ingredient) [26], and carbamazepine [27]
lymphomas, myeloid
neoplasms, other types of cancer, epilepsy, and dietary supplement
Histone methylationHistone
methyltransferase
inhibitors
HMTis-
Lysine-specific histone
methyltransferases
inhibitors
HKMTisphenelzine [28],
tranylcypromine [29], and
tazemetostat [30]
major depression, anxiety, and epithelioid sarcoma
Histone demethylationHistone demethylase
inhibitors
HDMisdeferiprone [31] and
deferasirox [32]
treatment of iron
overload in thalassemia and major or long-term
blood transfusions
Lysine-specific histone
demethylases inhibitors
LSDis/
KDMis
-
Protein arginine
methylation
Protein arginine
methyltransferase
inhibitors
PRMTis-
Protein arginine
citrullination
(deamination)
Protein arginine
deiminase inhibitors
PADisstreptomycin and
methotrexate [33]
antimicrobial therapy, chemotherapy agent,
and immune system
suppressant
PhosphorylationHistone kinase inhibitors-ruxolitinib, pacritinib,
pazopanib, vandetanib,
lapatinib, and erlotinib [25]
myelofibrosis, atopic
dermatitis, vitiligo,
renal cell carcinoma,
soft tissue sarcoma,
medullary thyroid
cancer, breast cancer, non-small cell lung
cancer, and pancreatic cancer
Ubiquitination/
deubiquitination
Inhibitors of ubiquitin signaling modulators (proteasome, target E1, E3, or DUB modulators)-bortezomib, carfizomib,
ixazomib, thalidomide,
lenalidomide, and
pomalidomide [34]
multiple myeloma, mantel cell lymphoma, and myelodysplastic syndromes
Poly(ADP-ribosyl)ation
(PARylation)
Poly(ADP-ribose) polymerase (PARP1) inhibitorsPARPisolaparib, niraparib,
rucaparib, and talazoparib [35]
different types of cancer
OthersInhibitors of proteins binding to
acetylated histones
PAHis-
Bromodomain and Extra terminal
motif proteins (BETs) inhibitors
BETisdinaciclib (granted orphan drug
status by the FDA in 2011) [36]
different types of cancer
Inhibitors of proteins binding
to methylated histones
PMHis-
Regulation by non-coding RNAncRNAspatisiran, givosiran,
and pegatanib [37]
familial amyloid
polyneuropathy, hepatic porphyria, and macular degeneration
Limitations to use the epidrugs are related to (1) their chemical instability; (2) ubiquitous activity of DNMTis accompanied by the development of side effects; (3) unspecific acetylation, deacetylation, and methylation of non-histone proteins resulting in undesirable side effects (such as fatigue and dysfunction of the gastrointestinal tract); and (4) cytotoxic effects and significant inhibition of hematopoiesis [38].
Figure 2. Epigenetic mechanisms. A DNA strand can be dynamically modified, mostly by adding methyl groups. Cytosine residues in DNA can be methylated by DNA methyltransferases (DNMTs) to 5-methylcytosine (5mC) (found primarily at CpG sites), which is oxidized to 5-hydroxymethylcytosine (5hmC) by ten-eleven translocation (TET) proteins. The 5hmC can be further oxidized by TET proteins to become 5-formylcytosine (5fC) and then 5-carboxylcytosine (5caC), or deaminated by activation-induced cytidine deaminase (AID) or APOBECs. In turn, accessibility can be controlled via histones. The parts of histone molecules protruding from nucleosomes can be modified by special “writer” proteins, adding different chemical groups (methyl, acetyl, phosphate, crotonyl, citrate, and serotonyl) or small proteins (ubiquitin and SUMO). Special “eraser” proteins remove marks from histones. Other proteins, called “readers”, recognize and interact with the certain marks of histone “tails” (N-terminus) to mediate specific transcription. Sometimes only one modification can be enough to regulate processes on chromatin. After the discovery of the first histone modifications, David Allis and Brian Strahl put forward the “histone code” hypothesis: a combination of histone marks reflects a code that is read by other proteins and determines the molecular processes at the site [39]. However, it seems that the real picture is more complicated, and not only histone marks control molecular processes on chromatin. Abbreviations: Cit = citrulline; DNMT = DNA methyltransferase; DUB = deubiquitylating enzyme; E1 = ubiquitin-activating enzyme; E2 = ubiquitin-conjugating enzyme; E3 = ubiquitin ligase; HAT = histone acetyltransferase; HDAC = histone deacetylase; HDM = histone demethylase; HMT = histone methyltransferase; PAD = peptidyl arginine deiminase; P = phosphate group; PRMT = protein arginine methyltransferase; R = arginine histone residue; S = serine histone residue; SENP = sentrin-specific protease; T = threonine histone residue; TET = TET protein; K = lysine histone residue; and Met = methyl group. Created with BioRender.com (accessed on 2 August 2022).
Figure 2. Epigenetic mechanisms. A DNA strand can be dynamically modified, mostly by adding methyl groups. Cytosine residues in DNA can be methylated by DNA methyltransferases (DNMTs) to 5-methylcytosine (5mC) (found primarily at CpG sites), which is oxidized to 5-hydroxymethylcytosine (5hmC) by ten-eleven translocation (TET) proteins. The 5hmC can be further oxidized by TET proteins to become 5-formylcytosine (5fC) and then 5-carboxylcytosine (5caC), or deaminated by activation-induced cytidine deaminase (AID) or APOBECs. In turn, accessibility can be controlled via histones. The parts of histone molecules protruding from nucleosomes can be modified by special “writer” proteins, adding different chemical groups (methyl, acetyl, phosphate, crotonyl, citrate, and serotonyl) or small proteins (ubiquitin and SUMO). Special “eraser” proteins remove marks from histones. Other proteins, called “readers”, recognize and interact with the certain marks of histone “tails” (N-terminus) to mediate specific transcription. Sometimes only one modification can be enough to regulate processes on chromatin. After the discovery of the first histone modifications, David Allis and Brian Strahl put forward the “histone code” hypothesis: a combination of histone marks reflects a code that is read by other proteins and determines the molecular processes at the site [39]. However, it seems that the real picture is more complicated, and not only histone marks control molecular processes on chromatin. Abbreviations: Cit = citrulline; DNMT = DNA methyltransferase; DUB = deubiquitylating enzyme; E1 = ubiquitin-activating enzyme; E2 = ubiquitin-conjugating enzyme; E3 = ubiquitin ligase; HAT = histone acetyltransferase; HDAC = histone deacetylase; HDM = histone demethylase; HMT = histone methyltransferase; PAD = peptidyl arginine deiminase; P = phosphate group; PRMT = protein arginine methyltransferase; R = arginine histone residue; S = serine histone residue; SENP = sentrin-specific protease; T = threonine histone residue; TET = TET protein; K = lysine histone residue; and Met = methyl group. Created with BioRender.com (accessed on 2 August 2022).
Cells 12 01464 g002

2.2. Significance and Requirements for Epigenetic Therapy for CNS Diseases

The increase in life expectancy, changes in lifestyle, and informational load resulted in the emergence of new pathogenic factors (such as aging and stress) and the decrease in significance of old pathogenic factors (such as infections and nutritional deficiencies). The profiling of CNS diseases in the world population shows a growing dominance of neurodegenerative disorders and psychophysiological stress-associated disorders [40,41,42]. Such a bias in the profile of pathologies defines new therapeutic goals and further stresses the fact that the contemporary therapeutic pharmacological approaches are not sufficiently effective. This situation challenges scientific society and the pharmaceutical industry and, among others, increases an interest in epidrugs as a means to treat CNS disorders.
Table 3 provides an overview of epidrugs showing activity to treat pathologies of the CNS in preclinical (mostly) and clinical studies.
While we already know a number of compounds with primary epigenetic activity, their rapid transfer from preclinical studies into the therapy of the CNS is hindered by some flaws and difficulties.
A list of the most crucial demands on epidrugs includes:
  • Permeability of the blood–brain barrier (BBB);
  • Regionality of action (for example, mainly in cortical regions);
  • Differential specificity of action at the level of neuronal or glial populations (might be crucial for a selective targeting of neuroinflammation [63]);
  • High selectivity of compounds to their molecular targets (at the level of protein isoforms);
  • If the drug is multitargeted, then its molecular action has to be highly balanced with no or few side effects;
  • Absence of cytotoxic effects on cells with a non-cancer phenotype;
  • Lack of systemic immunosuppressive and hematopoiesis-inhibiting effects;
  • Possibility of a differential therapy according to the phase of the disease (for example, conducting early postnatal treatment of pathologies associated with genetic disorders).
Based on the requirements and recent promising findings on the mechanisms of CNS disorders, the following cues for epidrug research and development can be delineated: (1) the search for compounds with good BBB permeability among selective inhibitors of TET proteins and DNA methyltransferases; (2) the search for compounds with high potency to selectively regulate neuroimmune inflammation and neurotrophic mechanisms; (3) the epigenetic activity profiling and assessing of the possible “off-label” use of compounds with canonical effects, such as anticonvulsant and antidepressant agents or metabolites; and (4) the employment of CRISPRon and CRISPRoff gene-editing technologies, as well as more advanced CRISPR/Cas system versions (such as Prime editing). The last is of particular interest because it enables cell- and locus-specific reactivation or repression of the transcription of a single gene or a combination of genes (multiplexed gene targeting) under a variety of promoters [64]. However, the clinical use of the transgenic approaches is still limited due to the problem of targeted delivery of compounds and genetic constructs to the human brain. The approach also needs to be more careful in addressing possible off-target effects.
When determining the strategies of therapeutic intervention in the treatment of CNS disorders, one should take into account the ability of complex physiological factors (lifestyle factors), such as physical activity, dietary factors, and fasting, as well as cognitive activity, to modulate the epigenetic landscape. Being ministers of beneficiary action (e.g., antidepressant and pro-cognitive), these factors do not lead to the development of unwanted side effects. Therefore, their epigenetic sequels may be considered as a prototype to explore the potential targets for specific epigenetic therapy.

3. Biological Targets of Epidrugs

The very nature of epidrugs points to their complexity and diversity of effects. On the cellular level, epidrugs influence fundamental processes, such as central metabolism, autophagy, cell survival, and the reprogramming of pluripotent or cancer cells. On the tissue level, the effects of epidrugs appear as an activation of neuroprotective and neuroplastic mechanisms as well as of the modulation of neuroinflammation, which involves the intercellular interaction and reorganization of the extracellular matrix. At the next level of complexity, epidrugs target particular endophenotypes and syndromes. A final achievement would be the correction of pathologies on the organismal level.
Figure 3 summarizes the most-studied effects of epidrugs, which are grouped according to their biological complexity levels.
To make an inventory of the available substances with epigenetic activity and to overview the contemporary tendencies in preclinical and clinical studies of their applicability for CNS treatments, we scrutinized respective publications based on the following logic.
Substances belonging to the continuously growing cluster of epidrugs were, first, stratified based on the main known biochemical/epigenetic mechanisms.
Second, epidrugs from each cluster with defined biochemical mechanisms were looked for in reports on the results of preclinical or clinical studies of therapeutic effects. Then, they were clustered in groups, which were formed with respect to the pathologies and pathology-related mechanisms that should be targeted (cytostatic action, neuroprotection, or neuroregeneration, treatment of neurological neurodegenerative disorders, and treatment of psychopathological syndromes).
We lined up the results of our search with respect to the number of published studies, which reflect the current interest in the respective epidrugs as therapeutic means for treating neurological and psychiatric disorders.
We also stressed the role of complex epigenetic regulators in the therapy of certain diseases, assuming their actual benefits.

3.1. Epigenetic Therapy for Brain Cancer

The strategy for treating cancer with epigenetically active compounds implies their use as adjuvant therapy [65]. To this end, the cytotoxic and/or cytostatic effects, as well as the ability of epidrugs to activate immune mechanisms, are exploited [66].
Nonselective HDACis are currently undergoing phase II clinical trials as a component of complex therapy for brain cancer [67,68,69,70]. The study on selective targets of HDACis singled out HDAC6. This enzyme shows a relatively high level of expression in the brain, which may be a prerequisite for the high efficacy of specific inhibitors. Moreover, selective HDAC6is, MPT0B291 and tubacin, show cytotoxic and cytostatic activity against human and rat glioma cells, but not normal astrocytes, both in vitro and in vivo [71,72]. The efficiency of the HDAC6 inhibitor JOC1 in the treatment of multiform glioblastoma was also demonstrated in clinical studies [73].
The spectrum of pharmacological activity of nonselective HDACis may be the determinant of a complex pharmacotherapy for brain cancer. For example, panobinostat, vorinostat, and romidepsin inhibit glycolysis in cancer cells by enhancing the oxidative metabolism (potentiating the expression of peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC1α) and peroxisome proliferator-activated receptor beta (PPARδ)). Since the cancer oxidative metabolism is mainly driven by fatty acid oxidation, a combination of fatty acid oxidation inhibitors and HDACis might be effective in cancer treatment, as it was proved in a patient-derived xenograft model of glioblastoma in mice [74].
BET inhibitors show cytotoxic activity against glioblastoma cells in vitro [75]. The synergistic cytotoxic effect of the BET inhibitor JQ-1 and the transcriptional activator CREB-binding protein (CBP), which recruiats bromodomain (BRD) proteins, showed to be effective as a therapy for diffuse intrinsic pontine glioma [76]. JQ-1 inhibits the growth of glioma implanted in mice when a transfer across the BBB is ensured using PEGylated nanoparticles [77].
DNMTis are also studied as adjuvant components in the treatment of brain cancer [78,79].
The experimental evidence and results of clinical trials accumulated to date strongly suggest that epidrugs may be effective in the treatment of brain neoplasms.

3.2. Neuroprotective and Pro-Regenerative Effects of Compounds with Epigenetic Activity

Cell death, which involves a wide spectrum of scenarios, can, besides being a part of the pathogenesis of neurodegenerative disorders, be assumed to be a leading component of degeneration in brain and spinal cord injuries, brain hemorrhage, ischemic stroke, chronic insufficiency of cerebral circulation, and toxicity-related damage (including that related to toxic metabolic syndromes).
HDACis showed neuroprotective effects in models of traumatic brain injury. Among the few mechanisms involved, the decrease in the neuroinflammatory response is particularly noteworthy [80]. Suppression of neuroinflammation is determined by an epidrug-evoked decline in the secretion of proinflammatory cytokines, suppression of glial cell proliferation, and changes in glial morphology [81,82,83,84,85]. This action of epidrugs is also associated with the increased expression of neurotrophic factors BDNF and NGF [86,87]. Additionally, several studies showed that the components of intracellular signaling, which increase cell survival (for example, the AKT-GSK3ß kinase cascade), are affected by HDACis [88,89]. Notably, the involvement of the PI3K-AKT-GSK3ß signaling pathway in the mechanisms of the neuroprotective effects of epidrugs emphasizes a synergism of HDACis and the activity of “conventional” drugs targeting the insulin- and serotonin-dependent intracellular signaling pathways [90]. Insufficiency of this cooperation may be an important determinant of resistance to tricyclic antidepressants and SSRIs [91,92,93].
The neuroprotective effect of nonspecific HDACis in models of ischemia, oxidative stress, and glutamate neurotoxicity is associated with the transcriptional suppression of pro-apoptotic factors, such as p75(NTR)-dependent caspase-3 and ubiquitin-conjugating enzyme E2N Ube2n [94,95,96]. Additionally, nonspecific HDACi valproic acid affects mitochondrial biogenesis [97] and reduces the oxidative stress caused by psychophysiological stress or global cerebral ischemia [98]. HDACis also show suppression of glial differentiation and activity and potentiation of neuron differentiation from progenitor cells during adult neurogenesis [99,100]. However, the prospects for the use of HDACis for nerve tissue bioengineering, with the help of pluripotent cell reprogramming technologies, is not entirely clear yet [101,102]. A limitation to the use of nonspecific epidrugs is related to a possible acetylation/deacetylation imbalance. Garcinol, which combines the properties of an HAT p300 inhibitor and an HDAC11 inhibitor, potentiates toxic effects in the 1-methyl-4-phenylpyridinium (MPP+) toxicity model [103].
Several observations further illustrate the importance of selective modulators of histone acetylation and deacetylation in realization of a neuroprotective effect [104,105]. The therapeutic effect of exifone, a selective activator of HDAC1, was previously attributed to its scavenger properties [106]. However, exifone-driven neuroprotection in the model of oxidative stress associated with the accumulation of 8-oxo-2-deoxyguanosine rather depends on the activation of HDAC1 and the reduction in H4K12 acetylation levels in neurons. This protective effect of exifone was not observed in astrocytes, which implies a cell-type-specific action [55,107]. Advantages of the selective inhibition of histone acetylation were also demonstrated in the model of peripheral nerve regeneration [108]. The effect of a nonspecific HDACi MS-275 in the axonotomy regeneration model was weaker than the effect of the histone acetyltransferase p300/CBP-associated factor (PCAF) overexpression [108,109].
A nonspecific DNMTi, 5-aza-deoxycytidine, potentiated MPP+ toxicity in dopaminergic neurons [43]. Another unspecific DNMT inhibitor zebularine, a nucleoside analogue of cytidine, increased the accumulation of β-amyloid in N2a mouse neuroblastoma cells in vitro [44]. In contrast, the DNMT1i RG108 suppressed apoptosis in motor neurons, providing neuroprotective effects [46]. Although the use of nonselective DNMTis does not appear to be a promising way for regulating neuronal function, the isoform-selective DNMT1 and DNMT3a/b inhibitors may nevertheless be more suitable [110,111].
BET inhibitors, especially those selective for BRD2/4 proteins, exhibit a characteristic suppression of the expression levels of pro-apoptotic factors and cytokines regulating inflammation (tumor necrosis factor alpha (TNF-α), C-X-C motif ligands 1 and 10 (CXCL1; CXCL10), and C-C motif ligand 2 (CCL2)), as well as matrix metalloproteinase 9 (MMP9), the extracellular cleaving enzyme involved in the maintenance of the extracellular matrix and membrane receptors [112,113,114,115,116].
Overall, known epidrugs facilitate the neurotrophic and neuroplastic processes, inhibit the mechanisms of programmed cell death, and suppress neuroinflammation. These features prompt researchers to look at the applicability of certain epidrugs as neuroprotectors and stimulators of neuronal regeneration but also suggest the domain of neurodegenerative disorders as a potential area of applicability.

3.3. Specific Therapy for Neurodegenerative Diseases

Epidrugs potentially can be used in the treatment of neurodegenerative disorders because they activate neuroprotective mechanisms, suppress neuroinflammation, and evoke the mechanisms of synaptogenesis and neuroplasticity [117]. Considering this, epidrugs affect specific aspects of the development and pathogenesis of neurodegenerative diseases by modulating expression and post-translational modifications of functional proteins as well as pathoproteins.
The most prevalent neurodegenerative diseases, such as Huntington’s disease [118], Parkinson’s disease [119,120], Alzheimer’s disease [121,122,123], and amyotrophic lateral sclerosis [124], are characterized by a dysregulation in histone acetylation in the brain. However, it is worth mentioning that in the case of Parkinson’s disease, the genome-wide H3K27ac hyperacetylation was found in the cortex [120], not only in the midbrain [119].
Despite the differential regulation of histones in neurodegenerative disorders, it is assumed that the use of HDACis can compensate for these pathological changes [125,126] via a complex mechanism including downregulation of pathoprotein production and inflammatory cytokines [127,128]. Besides targeting specific pathogenetic mechanisms, HDACis may influence the efficiency of learning and memory [129], therefore compensating for pathological phenotypes.
HDAC1/2 inhibitor BG45 enhances the level of synaptic proteins PSD95, spinophilin, and synaptophysin in the model of toxic β-amyloid expression in vitro [130]. In turn, selective HDACis BG45 and MGCD0103 (mocetinostat) not only reduce the neurotoxic effect of β-amyloid but also decrease its production [130,131]. At the same time, HDACis MS275 and CI994 increase the expression of apolipoprotein E4 [132], a risk factor gene associated with Alzheimer’s disease [133]. HDAC4/5 inhibitors and nonspecific HDACis suppress α-synuclein production Parkinson’s disease [134,135].
The use of HDACi LBH589 (panobinostat) in the early postnatal period in the genetic model of Huntington’s disease leads to a significant improvement in behavioral and neurochemical endophenotypes [136], which points to interference with the neurodevelopmental mechanisms of this disorder. Nonspecific HDACis sodium butyrate and trichostatin A normalize mitochondrial activity and, therefore, prevent impairments in motor learning and coordination in YAC128 transgenic mice (full-length HTT expression) [137]. The HDAC3 inhibitor RGFP966 activates the expression of genes characteristic of neuroplastic processes in the hippocampus of Huntington’s disease gene homolog (Hdh) knockout mice (HdhQ111-KO), which is accompanied by an improvement in cognitive functions [138].
Besides interference with histone acetylation, DNA methylation is also considered as a target for epigenetic therapy of neurodegenerative disorders [139]. A differential profile of DNA methylation is shown in the brain of patients with Parkinson’s [140] and Alzheimer’s [141] diseases. The use of lysine demethylase inhibitors (LSDis or KDMis) appears to be a promising strategy to prevent neurodegeneration. ORY-2001 (vafidemstat), combining properties of a KDM1A inhibitor and a monoamine oxidase B (MAO B) inhibitor [142], was reported to reduce agitation and aggression, as well as cause a rise in the total Neuropsychiatric Index and measures of caregiver burden in an open-label study in 12 people with Alzheimer’s disease (a six-month course of 1.2 mg of the drug daily) [143]. A clinical trial on the safety of this compound in the treatment of Alzheimer’s disease was recently completed [143].
The BETi JQ-1 affects pathological behavioral and molecular biological endophenotypes in a model of Alzheimer’s disease in mice. JQ-1 reduces Tau protein phosphorylation, an expression of pro-inflammatory factor genes (Il-1b, Il-6, Tnfa, Ccl2, Nos2, and Ptgs2) in 3xTg mice [144], prevents a decrease in cognitive functions, and restores the physiologically relevant expression of genes (e.g., ion channels and DNA repair) in the hippocampus in APP/PS1-21 mice [145]. Apabetalone, a compound with BET inhibitory activity, improves cognitive performance in people over 70 years of age. However, this may be due to its beneficial effects on the cardiovascular system [146].
However, existing HDACis and BETis have pronounced immunosuppressive and cytotoxic effects, which may be an obstacle to their use for CNS disorders, which require long-term treatment. The presence of side effects motivates a search for selectively acting compounds and/or routes of targeted drug delivery or drugs with local action. In this regard, a novel hybrid technology that exploits the capabilities of the CRISPR/dCas (dead Cas) complex, and an epigenetic enzyme recruited by it, is of increasing interest. The expression of such constructs allows a locus-specific reactivation (CRISPRon) or repression (CRISPRoff) of DNA transcription [64,147,148]. However, this promising approach requires significant optimization to achieve a high cell-specificity of the construct’s expression in humans. It should be mentioned that in addition to the use of viral vectors with BBB permeability, the possibility of using PEGylated immunoliposomes demonstrating good BBB permeability and high biocompatibility is being considered [149].

3.4. Epigenetic Therapy of Psychopathological Syndromes

The design or selection of epidrugs for the treatment of psychiatric disorders is determined by the characteristic biological features of psychiatric syndromes: (1) high polygenicity with small effect sizes of individual risk variants; (2) strong influence of external and internal environmental factors on the development and dynamics of psychopathological endophenotypes; and (3) overlap of pathogenetic mechanisms between psychopathological syndromes and neurodegenerative diseases (neuroinflammation, maladaptive neuroplasticity, and neurodegeneration).
It is also important to note that known psychoactive pharmacological substances are being proven to have epigenetic mechanisms of action. In turn, it is not surprising that many epigenetic drugs mediate psychical activity because they affect neuroplasticity, which is essential to maintaining functional homeostasis and adaptive changes in the brain.

3.4.1. Possibility for Off-Label Use of Substances with Known Psychopharmacological Activity

Antidepressants, antipsychotics, and drugs with antiepileptic activity are constantly confirmed to have epigenetic effects. These epigenetic activities may be the main determinants of the clinical effectiveness of these drugs, instead of their initially proposed modes of action. For instance, imipramine was introduced for depression treatment as early as 1957. As shown more recently, its antidepressant effect in a chronic stress mouse model is mediated by an increase in H3 acetylation in the Bdnf gene promoter and BDNF expression [150]. Valproic acid has been in use since 1962. Besides acting as an antiepileptic drug, it was proposed as an add-on therapeutic means to cure schizophrenia and bipolar disorder [151,152]. The administration of valproic acid is accompanied by an increase in the levels of H3 and H4 acetylation in the cortex and hippocampus. This regionality of epigenetic changes seems to be necessary for the effects of valproic acid in the treatment of panic attacks and anxiety disorders in humans, conditions with insufficient top–down control of emotions and general excitability [53,54]. Together with the revitalization of the known drugs and their off-label use, identifying their epigenetic effects may provide a new avenue for developing derivatives with primary epigenetic activity [150,153,154,155,156].

3.4.2. Drugs with Primary Epigenetic Activity

Effects of HDACis and BETis in the CNS are well known to be associated with an increase in the transcription of genes regulating neurogenesis, synaptogenesis, and neuroplasticity, a decrease in cytokine production, and the suppression of microglial activity. An increasing number of evidence indicates the effects of HDACis and BETis on the expression of neurotransmitter and neurohormone receptors and transporters (e.g., dopamine, serotonin, gamma-aminobutyric acid, glutamate, and corticosterone) [157,158,159,160,161]. The regulation of the expression of conventional and physiologically significant targets of psychoactive drugs further supports the idea of the prospective use of epigenetically active compounds as therapeutic agents for psychopathological conditions.
Furthermore, DNMTis hold promise for their use in treatments of stress-associated psychiatric diseases. Global and locus-specific DNA methylation reflects the active response of neuronal tissue to stress [162] and/or its ability to cope with stress (stress resilience) [150]. The direct [163] and indirect [154] reduction in global methylation leads to an antidepressant effect [164]. Psychophysiological stress is one of the leading pathogenetic factors of depression and post-traumatic stress disorder (PTSD) [165,166]. Increased methylation of the exon 1F of the NR3C1 glucocorticoid receptor is associated with higher hypothalamic–pituitary–adrenal (HPA) axis activity, increased production of cortisol, as well as pathological changes in the behavior of patients with depression and those who have suffered mental trauma in childhood. This may lead to a decrease in the expression of glucocorticoid receptors and impaired activity of the stress hormone axis [154,167,168,169]. The HPA axis can be one of the targets of epigenetic therapy for alleviating stress-related psychopathological syndromes.
Notably, DNMTis are expressed at different levels in various brain regions and are usually present in neurons and not glial cells [110]. Moreover, DNMT-mediated epigenetic regulation in neurons is isoform-specific. Thus, the expression of DNMT3A2 increases in response to stable synaptic activity [170], indicating the dependence of DNA methylation on the excitability of neurons. The effect of DNMT inhibitors on memory and learning appears to be centered on the phases of memory formation (learning, consolidation, and reconsolidation) [171,172,173,174,175]; however, these effects are not fully understood.
The BETi JQ-1, on the other hand, shows contradictory effects in mouse models of PTSD. It reduces the extinction of traumatic memory [176], decreases traumatic memory [177], or reduces the recognition of a familiar object during the object recognition tests assessing memory [60]. These observations indicate the dependence of neuroplastic processes on the activity of BRD1-4 proteins. However, the mechanism behind BETi effects during different phases of memory formation remains unclear.

3.5. Epigenetic Therapy for Developmental Disorders, Drug Addiction, Epilepsy, and Pain Disorders

3.5.1. Developmental Disorders

The role of epigenetic mechanisms in developmental genetic disorders has been widely acknowledged [178]. The efficiency of epidrugs in preclinical studies for these disorders is, therefore, not surprising. Thus, both the selective HDAC6 inhibitor SW-100 in the model of Martin–Bell syndrome (fragile X syndrome) [179] and sodium salt of valproic acid in the model of Angelman syndrome [180] prevented the development of the pathological phenotype (impairment of memory and learning, social behavior, and motor functions).
The BETi JQ-1 showed to be beneficial in the model of Rett syndrome [181]. The BET inhibitor I-BET858 altered the expression of genes belonging to the annotation clusters of “neuroplasticity” and “synaptogenesis” in a mouse model of autism. Interestingly, among all the clusters analyzed, only the expression of genes associated with Wnt signaling changed both in the cases of acute and chronic administration of the substance [182].

3.5.2. Drug Addiction

Changes in the epigenome have been demonstrated during the development of addiction to alcohol, nicotine, cocaine, amphetamine, cannabinoids, and opiates [183,184,185,186,187,188,189,190]. Notably, these changes are cell-type-specific; the profiles of epigenetic markers in neurons and astrocytes in response to psychostimulants and opiates differ [191].
Preclinical test results indicate the potential for using HDACis to treat drug addiction [192,193,194,195,196,197]. However, an increase in the dependence rate upon administration of HDACis was also noted, which can result from the interaction of the epidrug activity and the phase of addiction development [198].
The potential of BETis in the treatment of drug addiction is generally poorly studied, although the possibility for their use in the treatment of cocaine addiction has been discussed [194].
Neuroplastic changes involved in forming addiction to different drugs seem to be an attainable target for epigenetic therapy. However, since the answer to whether treatment should preferentially affect neuroplasticity in a selected neuronal population is “yes” [199], any success seems to depend on progress in the targeted delivery of epigenetic drugs.
It can also be stressed here that epigenome profiling is crucial to determine the mechanisms leading to sensitivity, resistance, and tolerance to pharmacological drugs [184,200].

3.5.3. Pain Syndromes and Epilepsy

The effects of epidrugs are mainly rapid. At the same time, they are fundamentally reversible. That makes the therapy of epilepsy and pain syndromes another potential field of epidrug application [201]. Indeed, the BET inhibitor JQ-1 reduces seizure activity in the pentylenetetrazole seizure model [60]. The analysis of clusters of enrichment of the full-genomic effect of JQ-1 indicates the significant changes in ionotropic receptors [202]. Changes in the expression of ion channels or other proteins in peripheral nerves, e.g., the mitochondrial transmembrane protein FUNDC1 [203], may underlie the antinociceptive effects of drugs with epigenetic activity [204,205,206].

3.6. Lifestyle Factors in Epigenome Modulation in the Treatment of Neurodegenerative Diseases and Psychopathological Conditions

Epigenetics is likely to play a major role in the interaction between the environment (both physical and social) and gene expression. The term “lifestyle” is defined as a complex of modifiable habits and a typical way of living for an individual. It includes factors such as diet, behavior, stress, physical and cognitive activity, working habits, smoking, and alcohol consumption, all of which are shown to alter epigenetic landscapes [207]. Moreover, the measurement of DNA methylation patterns allows to discriminate between individuals with a healthy versus unhealthy lifestyle, quantified by assessing diet, physical activity, and smoking and alcohol intake by individual [208]. A strategy that induces a complex therapeutic effect on the epigenome could thus consider the modulatory influence of lifestyle factors: adherence to sports and a healthy diet, cognitive activity, and the frequency of psychological stress. On the other hand, the specific constellation of epigenetic mechanisms that mediate the action of lifestyle factors may be a reference in the search for more specific epidrugs with similar features.
It is worth mentioning that due to the systemic effect that specific lifestyle interventions and epidrugs themselves have on metabolism and the immune system, their overall positive impact on brain function and homeostasis of nervous tissue can be, at least in part, a secondary phenomenon.

3.6.1. Inflammation as a Proxy for Epigenetic Changes Evoked by Systemic Lifestyle Factors

Both lifestyle factors and systemic neuroinflammation have been linked to several pathological conditions, including psychiatric and neurodegenerative disorders [209,210,211]. Inflammation is also elevated in obesity, a condition particularly amenable to lifestyle changes [212,213]. Moreover, psychiatric disorders and obesity are shown to co-occur [214]. In obesity, expanding adipose tissue secretes proinflammatory adipokines [215], such as interleukin-6 (IL-6) and TNF-a, which can lead to inflammation and metabolic dysfunction associated with obesity [216]. Interestingly, both IL-6 and TNF-a are also elevated in neurodegenerative and psychiatric disorders [217,218,219,220,221,222,223]. It is thus possible that obesity and psychiatric conditions further potentiate each other, or perhaps, one could lead to another through the mediation of inflammatory changes.
Chronic low-grade inflammation induces a multitude of epigenetic changes. Meta-analysis of EWASs of serum C-reactive protein (CRP), a sensitive marker for low-grade inflammation that is also a proxy for IL-6 levels, found differential methylation at 218 CpG sites to be associated with CRPs [224]. Authors propose that these genetic loci may underlie inflammation and serve as targets for the development of novel therapeutic interventions for inflammation. The more recent meta-analysis found 1511 independent differentially methylated loci associated with CRPs, mostly situated in TF binding sites and genomic enhancer regions [225]. It is also suggested that most of the identified differentially methylated genes are hypomethylated in inflammatory processes [226].
Importantly, increased systemic inflammation caused by an energy-dense Western diet is reversible by a change in diet, as switching back to normal food rescues the inflammatory profile in mice susceptible to atherosclerosis [227]. A reduction in body weight might thus be helpful in the management of psychiatric conditions via reduced inflammation and corresponding epigenetic changes. Nevertheless, Western diet-induced transcriptomic and epigenomic reprogramming of myeloid progenitor cells led to their increased proliferation and enhanced innate immune responses, which were not fully reversible after the change to a healthy diet [227].
Furthermore, obesity also increases oxidative stress through different mechanisms, including hyperleptinemia, chronic inflammation, low antioxidant defense, and postprandial reactive oxygen species (ROS) generation [228]. ROS and oxidative stress induce a reduction in the activity of DNMTis, which can lead to global hypomethylation. However, by decreasing TET activity, increasing DNMT expression via hypoxia-inducible factor 1 (HIF1), or recruiting DNMT/SIRT1 (HDAC)-containing complexes to H2O2-induced DNA double-strand breaks, ROS can also result in local hypermethylation. Increased levels of ROS have also been widely associated with increased histone acetylation [229].

3.6.2. Dietary Factors Trigger Epigenetic Mechanisms

Dietary change is presumably the most obvious lifestyle intervention to tackle obesity or metabolic dysfunction, and thus modulates systemic inflammation, oxidative stress, and potentially neuroinflammation. At the moment, the dietary factors that have epigenetic effects can be grouped into three classes: single nutrients, particularly vitamins and polyphenols [230]; microRNA from food, which can be safely delivered to organisms and is incorporated into extracellular vesicles [231]; and caloric excess or restriction [232].
The involvement of central metabolites in the regulation of the activity of epigenetic enzymes provides a mechanistic link between dietary and calorie intake and changes in the epigenetic landscape [233]. For example, ATP is required for activation of chromatin-remodeling complexes [234], acetyl-CoA is essential for histone acetylation [235], and NAD+ is a co-factor for sirtuin-type HDACis [236].
In turn, calorie restriction is an appreciated longevity factor, and mechanisms behind its action are mostly found to be linked to reduced nutrient signaling [237]. Fasting or a low-calorie diet drives changes in metabolism and the activation of autophagy and also initiates a complex shift in the epigenome [238,239,240]. Calorie restriction is also accompanied by the augmentation of neuroprotective mechanisms [241,242,243]. However, it is not only high body weight but also body fat distribution that affects the epigenome. Lower DNA methylation age acceleration distances correlate with waist-to-hip ratios in participants with healthy lifestyles [208].
Neurodevelopmental factors are particularly relevant in the etiology of psychiatric disorders [244]. Somewhat contrary to the beneficial effects of calorie restriction listed above, severe food restriction during pregnancy can lead to serious repercussions for the health outcomes of the offspring. Thus, individuals prenatally exposed to famine during the Dutch Hunger Winter had reduced glucose tolerance and increased risk for metabolic syndrome and showed less DNA methylation of the imprinted insulin-like growth factor 2 (IGF2) gene even six decades later [245,246]. Moreover, both the Dutch Hunger Winter and the Great Leap Forward of China, which also involved mass starvation, indicated that exposure to famine during the first trimester of gestation is associated with a higher incidence of schizophrenia and depression in adulthood [247,248].
On the other hand, a maternal high-calorie diet is also associated with the development of both metabolic and psychiatric diseases in offspring through epigenetic mechanisms [198,201,202,203,246,249,250,251]. Additionally, the choice and amounts of nutrients are important for the epigenetic milieu. Polyunsaturated fatty acids (PUFAs) that are very high in Western diets are shown to have epigenetic effects per se [252,253]. Particularly strong effects are seen during pregnancy. Mouse pups from dams treated with an omega-6 PUFA-rich diet prior to and throughout pregnancy showed disrupted cortical lamination and anxiety-like behavior, even when the omega-6 PUFA-rich diet contained a recommended omega-3-to-omega-6 ratio. Omega-6 PUFA enrichment reduced chromatin accessibility in general, but particularly of TFs downstream of cannabinoid receptor 1 (CB1), for example, that of signal transducer and activator of transcription 3 (STAT3), a factor necessary for inflammatory signaling [254]. Since omega-6 PUFAs are precursors to arachidonic acid, which in turn is a precursor to endocannabinoids, omega-6 PUFAs and, therefore, an endocannabinoid-rich diet were proposed to downregulate CB1 receptors in the offspring of the diet-treated dams via de-sensitization. It was hypothesized that de-sensitized CB1 receptors would downregulate several TFs through methylation and eventually downregulate immunoglobulin cell adhesion molecules, which alter brain development and lead to permanent neuroanatomical and psychopathological changes in the adult offspring of the omega-6 PUFA-treated dams. Dietary excesses before conception and during pregnancy thus have the potential to program the wiring of brain circuits through epigenetic mechanisms and affect emotional processing and cognition for life [254].
Several dietary factors exert epigenetic effects and modulate psychopathological conditions independent of the fatty acid or caloric content of food. The dietary factor acetyl-L-carnitine [235], which is also naturally produced by the body, increases global histone acetylation by donating its acetyl groups. This is associated with neuroprotective effects and decreased neuroinflammatory responses [255,256,257]. Another class of important dietary components, polyamines, is also suggested to exert their action by inhibiting HAT activity [258]. Polyphenols, a large family of natural compounds from plant foods present in cruciferous vegetables, soy products, and green tea, have been shown to modify the activity of DNMTis, HATis, and HDACis [259,260,261,262]. Polyphenols include resveratrol and curcumin that are well known for their antiaging properties and are considered potential adjuvant treatments for mental and brain health. However, there are mixed clinical results, potentially owing to interindividual differences and antioxidant effects of those compounds [263]. Of note, maternal resveratrol supplementation of high-calorie diets during pregnancy was found to have some benefits for maternal and fetal metabolic health in several animal studies. Resveratrol was therefore proposed to mimic the positive effects of calorie restriction and physical activity in the case of maternal overeating, which might be otherwise difficult to address while avoiding malnutrition during pregnancy [246]. It should be noted that biologically active substances of plant and animal origin are often used as prototypes or as precursors for the synthesis of new compounds with epigenetic activity [264,265].

3.6.3. The Role of Physical Activity in Shaping the Epigenetic Landscape

Exercise increases the activity of histone acetyltransferases and histone deacetylases, reduces the level of global DNA methylation, and thus sets back the “time” of the epigenetic clock [266,267,268,269,270,271]. The changes in the epigenetic landscape go along with the therapeutic efficacy of physical exercises, which exert beneficial effects on cognitive impairment during aging, and can be used as an add-on therapy for neurodegenerative diseases [272,273] and, with less evidence, for psychopathological conditions [238,274].

3.7. Other Molecular Targets of Epigenetic Therapy for CNS Disorders

3.7.1. TET Proteins

TET proteins perform hydroxylation of DNA at 5-methylcytosine (5mC) [275], therefore changing the probability of DNA methylation and demethylation [276]. TET proteins are involved in all aspects of neuronal development and neuroplasticity; therefore, their functional activity seems to be a significant factor in the pathogenesis of neurodegenerative and psychopathological disorders [277]. Unlike cells of other lineages, which lose the expression of TET proteins during development and differentiation, a high level of expression of TET proteins is retained in fully differentiated neurons. The TET protein expression is believed to be caused by a high level of 5-hydroxymethylcytosine (5hmC) [278] and high transcriptional competence of DNA. The neuron-specific TET expression is a promising feature that allows one to consider TET proteins to develop an accurate epigenetic intervention specific to neurons [279,280].

3.7.2. Non-Coding RNAs

The most important regulators of the epigenome are non-coding RNAs. They are mostly represented by microRNAs (miRNAs), or enhancer and long non-coding RNAs (eRNAs and lncRNAs) which show a complex tertiary structure. MicroRNAs control transcription and intranuclear and cytoplasmic processing of genetic information; enhancer RNAs mainly mediate changes in the structural organization of DNA [281]. There is a potential for the use of miRNAs as antitumor agents and regenerative agents; but, at present, many drugs are withdrawn from clinical trials in the early phases due to the high frequency of side effects [282]. Currently, the possibility of using miRNAs in the treatment of neurological diseases and psychopathological conditions based on their impact on neuroimmune mechanisms is being questioned [283].
There is ample evidence of significant associations between specific miRNA expression profiles and psychopathological conditions [284,285,286,287,288,289]. However, the use of miRNAs as therapeutic agents is extremely difficult. Firstly, the spectrum of miRNA effects is wide, and it is not limited to the effect in the CNS, which increases the likelihood of side effects. This could be overcome with cell-specific delivery of miRNAs (e.g., with antibody- and nanobody-based delivery). Secondly, miRNAs are readily degraded, and there is a need to optimize the methods for their targeted organ delivery [290], for example, using “minicells” [291].

3.8. Clinical Trials

To assess the current state of clinical trials for substances with epigenetic activity in the treatment of pathological conditions of the CNS, an analysis of the US National Institutes of Health database (clinicaltrials.gov) [292] was carried out. Table 4 presents data on the number of active + completed (A + C) and terminated + withdrawn (T + W) studies of the compounds with primary epigenetic activity that are approved for use. The largest number of trials was conducted to assess the safety or therapeutic activity of compounds in the treatment of brain cancers, mainly of glial origin. This bias is primarily due to the exploitation of the cytostatic and cytotoxic properties of the compounds, which are especially pronounced in glial cells.
Upon analysis, the breadth of the goals of clinical trials with valproic acid or its salts used as monotherapy or as a component of complex therapy was noteworthy. In addition to the most common pathologies targeted with valproic acid (epilepsy, affective disorders, and schizophrenia), the therapeutic efficacy of this drug is being studied in the treatment of pain syndromes, depression and PTSD, Alzheimer’s disease (one study), spinal muscular atrophy (six studies), and oncological diseases of the brain (glioblastomas, gliomas of different localizations I-IV degrees of malignancy, gliomatosis, neuroblastoma, and medulloblastoma). Hydralazine is being studied as a primary therapy for Alzheimer’s disease in one database-registered clinical trial [293]. Its effect on autophagy is seen as the leading mechanism of action. The rationale for the study of phenelzine and tranylcypromine in the treatment of depression is based on their MAO inhibitory action [294]; epigenetic effects are not considered.

4. Conclusions

It is well accepted that epigenetic mechanisms are the core links between pathological environmental factors and disease. In analogy, it is true that interfering with the epigenetic mechanism might be the way to prevent the development or slow down the progress of neurological and psychiatric diseases. Therefore, epigenetic therapy of diseases of the CNS appears to be a promising field in pharmacology.
The development of epigenetic therapy for CNS diseases nonetheless faces several challenges:
  • Need for inhibitors and activators specific to different types and isoforms of epigenetic enzymes that also have good BBB permeability;
  • Discovery of compounds with bimodal or complex activity;
  • Understanding the changes in genome–epigenome interactions;
  • Expansion of the bioinformatical annotation epigenetic markers at different biological levels;
  • Evaluation of the applicability of the CRISPRon and CRISPRoff tools in humans;
  • Further development and optimization of the cell-specific delivery of compounds or genetic constructs using viral vectors, anti- or nanobodies, or PEGylated immunoliposomes.
The expanding amount of data from studies interrogating the mechanisms of action and effects of epigenetically active drugs and compounds in the CNS strongly suggests a potential for the therapeutic use of epidrugs in brain cancer. The present review particularly outlines the prospects of using HDACis, DNMTis, and BETis for the therapy of brain neoplasms (mainly of glial origin). Analysis of the CNS-specific targets and effects of epidrugs with respect to functional regulation of brain activity reveals that they also often target glial elements, affecting microglia differentiation and secretory activity. Glia provide support, protection, and, at the same time, foster neuroinflammation processes, thus regulating them negatively and positively. Although the epigenetic biomarkers of neuroinflammation in neurological (and psychiatric) disorders are not comprehensively described [295,296,297] to give us exact targets for drugs, the strategy of epigenetic therapeutic intervention can evolve based on the glia-interfering effects of epidrugs. Fundamentally, epidrugs may control neuroinflammation, the mechanism that is at the core of many neurological and psychiatric disorders, and regulate the expression and post-translational modifications of pathological proteins. Therefore, epidrugs can be generally considered symptomatic therapies with a wide spectrum of action. Additionally, it seems reasonable to make use of complex and balanced physiological modulators of the epigenome, such as single nutrients, physical activity, and a low-calorie diet, as an add-on therapy for psychopathological and neurodegenerative conditions and age-associated dementia.

Author Contributions

Conceptualization, E.A.A.; writing—original draft preparation, E.A.A., E.L. and M.G.G.; writing—review and editing, E.A.A., E.L. and M.G.G.; visualization, M.G.G.; project administration, M.G.G.; supervision, E.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Authors thank Anton A. Keskinov for the initial idea and Valentin V. Makarov for support of this writing project.

Conflicts of Interest

The authors declare no conflict of interest.

Glossary

architectural proteinsproteins playing a structural role in chromatin organization, include linker and core histone proteins and insulator proteins
acetylated histone proteinan epigenetic marker; an indicator of transcriptional competence
BBBblood–brain barrier
BDNFbrain-derived neurotrophic factor
BETi(s)BET inhibitor(s) of transcription factors of the BRD family
BRDbromodomain [bromo- and extra-terminal motif (BET) domains] family of proteins (BRD1, BRD2, BRD3, BRD4, and BRDT) that bind to acetylated histone and promote the recruitment of macromolecular complexes with an increase in DNA transcriptional competence; belong to the functional class of readers
CB1cannabinoid receptor 1
CBPcyclic adenosine monophosphate response element-binding protein [CREB-binding protein]; transcriptional coactivator
CCL2C-C motif ligand 2; chemokine
ChIPchromatin immunoprecipitation; protocols for profiling chemical modification of chromatin using immunoprecipitating methods
chromatin accessibilitythe degree to which nuclear macromolecules are able to physically contact chromatinized DNA and is determined by the occupancy and topological organization of nucleosomes and other chromatin-binding factors that regulate access to the DNA
chromatin-binding factorsnon-histone macromolecules binding to DNA (either directly or indirectly)
CNScentral nervous system
CpGcytosine–guanine site; a dinucleotide sequence subject to preferred methylation; CpG sequences are predominantly concentrated in non-coding regions of the genome (CpG islands [CpG island]) and are associated with promoter DNA regions
CRPC-reactive protein
CXCL1/10C-X-C motif ligand 1/10; chemokine
DNAdeoxyribonucleic acid
DNMTDNA methyltransferase; methyltransferase of the DNMT3a type is represented by several isoforms and is responsible for de novo methylation, and DNMT1 (and DNMT3b)—interact with methylated loci; DNMT belongs to the functional class of writers
DNMTi(s)DNMT inhibitor(s)
enhancera non-coding regulatory region of DNA, which, after binding of transcription factors to it, stimulates gene transcription
epigenome (epigenetic landscape)a set of epigenetic phenomena in a cell at the estimated time
euchromatinareas of decondensed, transcriptionally competent state of chromatin
EWASsepigenome-wide association studies
GWASsgenome-wide association studies
Hhistone proteins (H2A, H2B, H3, H4), constituting, in the form of dimers, a nucleosome complex
HAThistone acetyltransferase; belongs to the functional group of writers; in general, HAT increases gene expression
HATi(s)HAT inhibitor(s)
HDAChistone deacetylase; HDACs are epigenetic erasers and make the DNA less accessible to transcription factors; the HDAC family includes zinc-dependent enzymes of the HDACI, HDACII, HDACIV, and NAD+-dependent HDACIII classes (sirtuins, SIRT)
HDACi(s)HDAC inhibitor(s)
heterochromatinareas of condensed, transcriptionally incompetent state of chromatin (DNA and histone proteins)
HIF1hypoxia-inducible factor 1
IGF2insulin-like growth factor 2
IL-6interleukin-6; has both pro- and anti-inflammatory activities depending on the tissue type
insulatora non-coding protein-binding regulatory DNA sequence that regulates inter-chromosomal (enhancer-blocking insulators) and intra-chromosomal (barrier to the spread of neighboring heterochromatin) interactions within a particular DNA locus, flanked by insulators
KDMlysine-specific demethylase [lysine demethylase]; KDM belongs to the functional class of erasers
KDMi(s)KDM inhibitor(s)
KMTlysine histone methyltransferase; KMT belongs to the functional class of writers
KMTi(s)KMT inhibitor(s)
MAO Bmonoamine oxidase B
MMP9matrix metalloproteinase 9
methylated cytosine DNA (5mC, 5-methylcytosine)epigenetic marker; an indicator of transcriptional incompetence; 5hmC—5-hydroxymethylcytosine, derivative of 5mC acting in further oxidation
ncRNAnon-coding functionally active RNA sequences [non-coding RNA]; the most important regulators of the epigenome are microRNAs [miRNAs], enhancer and long-length non-coding RNAs [eRNAs and lncRNAs]
NGFnerve growth factor
NGSnext-generation sequencing; massively parallel DNA and RNA sequencing methods providing ultra-high throughput, scalability, and speed of analysis
nucleosome (complex)the main structure-forming element of chromatin, an octamer of histone proteins encircled by approximately 147 DNA bp; together the composition and post-translational modification of nucleosomes regulate chromatin accessibility
p75(NTR)p75 neurotrophin receptor, mediating a set of signaling pathways, including caspase-dependent signaling
PCAFp300/CBP-associated factor, regulates crosstalk-dependent acetylation of histone H3 by distal site recognition
PGC1αperoxisome proliferator-activated receptor gamma coactivator 1-alpha; transcriptional coactivator
PK/PDpharmacokinetics and pharmacodynamics
PPARδperoxisome proliferator-activated receptor delta; nuclear hormone receptor
promoternon-coding DNA region that binds to RNA polymerase and helper proteins initiating the transcription
PUFA(s)polyunsaturated fatty acid(s)
RNAribonucleic acid
ROSreactive oxygen species
STAT3signal transducer and activator of transcription 3
TADtopologically associated domain; a genomic region that self-interacts to form a three-dimensional structure, which determines the possibility of active interaction of regulatory elements (promoter–enhancer loops) and the transcriptional competence of a particular DNA locus
TF(s)transcription factor(s); non-histone protein(s) binding directly to DNA and controlling the transcription rates
TET proteinsten-eleven translocation proteins; enzymes that hydroxylate methylated cytosine (5mC); TET proteins belong to the functional class of erasers
TNF-αtumor necrosis factor alpha; cytokine

References

  1. Passarge, E. Emil Heitz and the Concept of Heterochromatin: Longitudinal Chromosome Differentiation Was Recognized Fifty Years Ago. Am. J. Hum. Genet. 1979, 31, 106–115. [Google Scholar] [PubMed]
  2. Ganesan, A.; Arimondo, P.B.; Rots, M.G.; Jeronimo, C.; Berdasco, M. The Timeline of Epigenetic Drug Discovery: From Reality to Dreams. Clin. Epigenetics 2019, 11, 174. [Google Scholar] [CrossRef] [PubMed]
  3. Luger, K.; Mäder, A.W.; Richmond, R.K.; Sargent, D.F.; Richmond, T.J. Crystal Structure of the Nucleosome Core Particle at 2.8 Å Resolution. Nature 1997, 389, 251–260. [Google Scholar] [CrossRef]
  4. Hales, B.F.; Grenier, L.; Lalancette, C.; Robaire, B. Epigenetic Programming: From Gametes to Blastocyst. Birth Defects Res. Part A Clin. Mol. Teratol. 2011, 91, 652–665. [Google Scholar] [CrossRef]
  5. Ostermeier, G.C.; Miller, D.; Huntriss, J.D.; Diamond, M.P.; Krawetz, S.A. Delivering Spermatozoan RNA to the Oocyte. Nature 2004, 429, 154. [Google Scholar] [CrossRef] [PubMed]
  6. Ramazi, S.; Allahverdi, A.; Zahiri, J. Evaluation of Post-Translational Modifications in Histone Proteins: A Review on Histone Modification Defects in Developmental and Neurological Disorders. J. Biosci. 2020, 45, 135. [Google Scholar] [CrossRef]
  7. Allis, C.D.; Jenuwein, T. The Molecular Hallmarks of Epigenetic Control. Nat. Rev. Genet. 2016, 17, 487–500. [Google Scholar] [CrossRef]
  8. Pidsley, R.; Zotenko, E.; Peters, T.J.; Lawrence, M.G.; Risbridger, G.P.; Molloy, P.; Van Djik, S.; Muhlhausler, B.; Stirzaker, C.; Clark, S.J. Critical Evaluation of the Illumina MethylationEPIC BeadChip Microarray for Whole-Genome DNA Methylation Profiling. Genome Biol. 2016, 17, 208. [Google Scholar] [CrossRef]
  9. Weaver, I.C.G.; Cervoni, N.; Champagne, F.A.; D’Alessio, A.C.; Sharma, S.; Seckl, J.R.; Dymov, S.; Szyf, M.; Meaney, M.J. Epigenetic Programming by Maternal Behavior. Nat. Neurosci. 2004, 7, 847–854. [Google Scholar] [CrossRef]
  10. Cazaly, E.; Saad, J.; Wang, W.; Heckman, C.; Ollikainen, M.; Tang, J. Making Sense of the Epigenome Using Data Integration Approaches. Front. Pharmacol. 2019, 10, 126. [Google Scholar] [CrossRef]
  11. EWAS Hub. Available online: https://ngdc.cncb.ac.cn/ewas/datahub (accessed on 16 July 2022).
  12. Human Epigenome Project. Available online: https://www.epigenome.org/ (accessed on 23 July 2022).
  13. Sharing Epigenomes Globally. Nat. Methods 2018, 15, 151. [CrossRef]
  14. Saulnier, K.M.; Bujold, D.; Dyke, S.O.M.; Dupras, C.; Beck, S.; Bourque, G.; Joly, Y. Benefits and Barriers in the Design of Harmonized Access Agreements for International Data Sharing. Sci. Data 2019, 6, 297. [Google Scholar] [CrossRef] [PubMed]
  15. Horvath, S. DNA Methylation Age of Human Tissues and Cell Types. Genome Biol. 2013, 14, R115. [Google Scholar] [CrossRef] [PubMed]
  16. Horvath, S.; Ritz, B.R. Increased Epigenetic Age and Granulocyte Counts in the Blood of Parkinson’s Disease Patients. Aging 2015, 7, 1130–1142. [Google Scholar] [CrossRef]
  17. Umehara, T. Epidrugs: Toward Understanding and Treating Diverse Diseases. Epigenomes 2022, 6, 18. [Google Scholar] [CrossRef] [PubMed]
  18. Qi, Y.; Wang, D.; Wang, D.; Jin, T.; Yang, L.; Wu, H.; Li, Y.; Zhao, J.; Du, F.; Song, M.; et al. HEDD: The Human Epigenetic Drug Database. Database 2016, 2016, baw159. [Google Scholar] [CrossRef]
  19. Histome2. Epidrugs. Available online: http://www.actrec.gov.in/histome2/Epidrugs/epidrugs.php (accessed on 23 July 2022).
  20. Shah, S.G.; Mandloi, T.; Kunte, P.; Natu, A.; Rashid, M.; Reddy, D.; Gadewal, N.; Gupta, S. HISTome2: A Database of Histone Proteins, Modifiers for Multiple Organisms and Epidrugs. Epigenetics Chromatin 2020, 13, 31. [Google Scholar] [CrossRef]
  21. Kronfol, M.M.; Dozmorov, M.G.; Huang, R.; Slattum, P.W.; McClay, J.L. The Role of Epigenomics in Personalized Medicine. Expert Rev. Precis. Med. Drug Dev. 2017, 2, 33–45. [Google Scholar] [CrossRef]
  22. Ruiz-Magaña, M.J.; Martínez-Aguilar, R.; Lucendo, E.; Campillo-Davo, D.; Schulze-Osthoff, K.; Ruiz-Ruiz, C. The Antihypertensive Drug Hydralazine Activates the Intrinsic Pathway of Apoptosis and Causes DNA Damage in Leukemic T Cells. Oncotarget 2016, 7, 21875–21886. [Google Scholar] [CrossRef]
  23. Qureshi, I.A.; Mehler, M.F. Developing Epigenetic Diagnostics and Therapeutics for Brain Disorders. Trends Mol. Med. 2013, 19, 732–741. [Google Scholar] [CrossRef]
  24. Huang, M.; Huang, J.; Zheng, Y.; Sun, Q. Histone Acetyltransferase Inhibitors: An Overview in Synthesis, Structure-Activity Relationship and Molecular Mechanism. Eur. J. Med. Chem. 2019, 178, 259–286. [Google Scholar] [CrossRef] [PubMed]
  25. Luan, Y.; Li, J.; Bernatchez, J.A.; Li, R. Kinase and Histone Deacetylase Hybrid Inhibitors for Cancer Therapy. J. Med. Chem. 2019, 62, 3171–3183. [Google Scholar] [CrossRef] [PubMed]
  26. 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. [Google Scholar] [CrossRef] [PubMed]
  27. Volmar, C.-H.; Wahlestedt, C. Histone Deacetylases (HDACs) and Brain Function. Neuroepigenetics 2015, 1, 20–27. [Google Scholar] [CrossRef]
  28. Prusevich, P.; Kalin, J.H.; Ming, S.A.; Basso, M.; Givens, J.; Li, X.; Hu, J.; Taylor, M.S.; Cieniewicz, A.M.; Hsiao, P.-Y.; et al. A Selective Phenelzine Analogue Inhibitor of Histone Demethylase LSD1. ACS Chem. Biol. 2014, 9, 1284–1293. [Google Scholar] [CrossRef]
  29. Duan, Y.-C.; Ma, Y.-C.; Qin, W.-P.; Ding, L.-N.; Zheng, Y.-C.; Zhu, Y.-L.; Zhai, X.-Y.; Yang, J.; Ma, C.-Y.; Guan, Y.-Y. Design and Synthesis of Tranylcypromine Derivatives as Novel LSD1/HDACs Dual Inhibitors for Cancer Treatment. Eur. J. Med. Chem. 2017, 140, 392–402. [Google Scholar] [CrossRef]
  30. Liu, A.; Han, J.; Nakano, A.; Konno, H.; Moriwaki, H.; Abe, H.; Izawa, K.; Soloshonok, V.A. New Pharmaceuticals Approved by FDA in 2020: Small-molecule Drugs Derived from Amino Acids and Related Compounds. Chirality 2022, 34, 86–103. [Google Scholar] [CrossRef]
  31. Khodaverdian, V.; Tapadar, S.; MacDonald, I.A.; Xu, Y.; Ho, P.-Y.; Bridges, A.; Rajpurohit, P.; Sanghani, B.A.; Fan, Y.; Thangaraju, M.; et al. Deferiprone: Pan-Selective Histone Lysine Demethylase Inhibition Activity and Structure Activity Relationship Study. Sci. Rep. 2019, 9, 4802. [Google Scholar] [CrossRef]
  32. Roatsch, M.; Hoffmann, I.; Abboud, M.I.; Hancock, R.L.; Tarhonskaya, H.; Hsu, K.-F.; Wilkins, S.E.; Yeh, T.-L.; Lippl, K.; Serrer, K.; et al. The Clinically Used Iron Chelator Deferasirox Is an Inhibitor of Epigenetic JumonjiC Domain-Containing Histone Demethylases. ACS Chem. Biol. 2019, 14, 1737–1750. [Google Scholar] [CrossRef]
  33. Ahmad Nadzirin, I.; Chor, A.L.T.; Salleh, A.B.; Rahman, M.B.A.; Tejo, B.A. Discovery of New Inhibitor for the Protein Arginine Deiminase Type 4 (PAD4) by Rational Design of α-Enolase-Derived Peptides. Comput. Biol. Chem. 2021, 92, 107487. [Google Scholar] [CrossRef]
  34. Wu, H.Q.; Baker, D.; Ovaa, H. Small Molecules That Target the Ubiquitin System. Biochem. Soc. Trans. 2020, 48, 479–497. [Google Scholar] [CrossRef] [PubMed]
  35. Ummarino, S.; Hausman, C.; Di Ruscio, A. The PARP Way to Epigenetic Changes. Genes 2021, 12, 446. [Google Scholar] [CrossRef] [PubMed]
  36. Alqahtani, A.; Choucair, K.; Ashraf, M.; Hammouda, D.M.; Alloghbi, A.; Khan, T.; Senzer, N.; Nemunaitis, J. Bromodomain and Extra-Terminal Motif Inhibitors: A Review of Preclinical and Clinical Advances in Cancer Therapy. Future Sci. OA 2019, 5, FSO372. [Google Scholar] [CrossRef] [PubMed]
  37. Damase, T.R.; Sukhovershin, R.; Boada, C.; Taraballi, F.; Pettigrew, R.I.; Cooke, J.P. The Limitless Future of RNA Therapeutics. Front. Bioeng. Biotechnol. 2021, 9, 628137. [Google Scholar] [CrossRef] [PubMed]
  38. Cherblanc, F.; Chapman-Rothe, N.; Brown, R.; Fuchter, M. Current Limitations and Future Opportunities for Epigenetic Therapies. Future Med. Chem. 2012, 4, 425–446. [Google Scholar] [CrossRef]
  39. Strahl, B.D.; Allis, C.D. The Language of Covalent Histone Modifications. Nature 2000, 403, 41–45. [Google Scholar] [CrossRef]
  40. Global Burden of Disease Study 2017; Institute for Health Metrics and Evaluation: Seattle, WA: IHME, 2018. Available online: https://www.healthdata.org/policy-report/findings-global-burden-disease-study-2017 (accessed on 31 December 2022).
  41. James, S.L.; Abate, D.; Abate, K.H.; Abay, S.M.; Abbafati, C.; Abbasi, N.; Abbastabar, H.; Abd-Allah, F.; Abdela, J.; Abdelalim, A.; et al. Global, Regional, and National Incidence, Prevalence, and Years Lived with Disability for 354 Diseases and Injuries for 195 Countries and Territories, 1990–2017: A Systematic Analysis for the Global Burden of Disease Study 2017. Lancet 2018, 392, 1789–1858. [Google Scholar] [CrossRef]
  42. Spencer, S. Global Burden of Disease 2010 Study: A Personal Reflection. Glob. Cardiol. Sci. Pract. 2013, 2013, 15. [Google Scholar] [CrossRef]
  43. Wang, Y.; Wang, X.; Li, R.; Yang, Z.-F.; Wang, Y.-Z.; Gong, X.-L.; Wang, X.-M. A DNA Methyltransferase Inhibitor, 5-Aza-2′-Deoxycytidine, Exacerbates Neurotoxicity and Upregulates Parkinson’s Disease-Related Genes in Dopaminergic Neurons. CNS Neurosci. Ther. 2013, 19, 183–190. [Google Scholar] [CrossRef]
  44. Li, W.; Jiang, M.; Zhao, S.; Liu, H.; Zhang, X.; Wilson, J.X.; Huang, G. Folic Acid Inhibits Amyloid β-Peptide Production through Modulating DNA Methyltransferase Activity in N2a-APP Cells. Int. J. Mol. Sci. 2015, 16, 25002–25013. [Google Scholar] [CrossRef]
  45. Lin, N.; Qin, S.; Luo, S.; Cui, S.; Huang, G.; Zhang, X. Homocysteine Induces Cytotoxicity and Proliferation Inhibition in Neural Stem Cells via DNA Methylation in Vitro. FEBS J. 2014, 281, 2088–2096. [Google Scholar] [CrossRef] [PubMed]
  46. Martin, L.J.; Wong, M. Aberrant Regulation of DNA Methylation in Amyotrophic Lateral Sclerosis: A New Target of Disease Mechanisms. Neurotherapeutics 2013, 10, 722–733. [Google Scholar] [CrossRef] [PubMed]
  47. Huang, S.; Dong, W.; Jiao, Z.; Liu, J.; Li, K.; Wang, H.; Xu, D. Prenatal Dexamethasone Exposure Induced Alterations in Neurobehavior and Hippocampal Glutamatergic System Balance in Female Rat Offspring. Toxicol. Sci. 2019, 171, kfz163. [Google Scholar] [CrossRef] [PubMed]
  48. Moloney, R.D.; Stilling, R.M.; Dinan, T.G.; Cryan, J.F. Early-Life Stress-Induced Visceral Hypersensitivity and Anxiety Behavior Is Reversed by Histone Deacetylase Inhibition. Neurogastroenterol. Motil. 2015, 27, 1831–1836. [Google Scholar] [CrossRef]
  49. Gurbani, S.S.; Yoon, Y.; Weinberg, B.D.; Salgado, E.; Press, R.H.; Cordova, J.S.; Ramesh, K.K.; Liang, Z.; Velazquez Vega, J.; Voloschin, A.; et al. Assessing Treatment Response of Glioblastoma to an HDAC Inhibitor Using Whole-Brain Spectroscopic MRI. Tomography 2019, 5, 53–60. [Google Scholar] [CrossRef]
  50. Covington, H.E.; Maze, I.; LaPlant, Q.C.; Vialou, V.F.; Ohnishi, Y.N.; Berton, O.; Fass, D.M.; Renthal, W.; Rush, A.J.; Wu, E.Y.; et al. Antidepressant Actions of Histone Deacetylase Inhibitors. J. Neurosci. 2009, 29, 11451–11460. [Google Scholar] [CrossRef]
  51. Sah, A.; Sotnikov, S.; Kharitonova, M.; Schmuckermair, C.; Diepold, R.P.; Landgraf, R.; Whittle, N.; Singewald, N. Epigenetic Mechanisms within the Cingulate Cortex Regulate Innate Anxiety-Like Behavior. Int. J. Neuropsychopharmacol. 2019, 22, 317–328. [Google Scholar] [CrossRef] [PubMed]
  52. Bach, D.R.; Korn, C.W.; Vunder, J.; Bantel, A. Effect of Valproate and Pregabalin on Human Anxiety-like Behaviour in a Randomised Controlled Trial. Transl. Psychiatry 2018, 8, 157. [Google Scholar] [CrossRef]
  53. Kinrys, G.; Pollack, M.H.; Simon, N.M.; Worthington, J.J.; Nardi, A.E.; Versiani, M. Valproic Acid for the Treatment of Social Anxiety Disorder. Int. Clin. Psychopharmacol. 2003, 18, 169–172. [Google Scholar] [CrossRef]
  54. Primeau, F.; Fontaine, R.; Beauclair, L. Valproic Acid and Panic Disorder. Can. J. Psychiatry 1990, 35, 248–250. [Google Scholar] [CrossRef]
  55. Patnaik, D.; Pao, P.-C.; Zhao, W.-N.; Silva, M.C.; Hylton, N.K.; Chindavong, P.S.; Pan, L.; Tsai, L.-H.; Haggarty, S.J. Exifone Is a Potent HDAC1 Activator with Neuroprotective Activity in Human Neuronal Models of Neurodegeneration. ACS Chem. Neurosci. 2021, 12, 271–284. [Google Scholar] [CrossRef] [PubMed]
  56. Wiese, M.; Schill, F.; Sturm, D.; Pfister, S.; Hulleman, E.; Johnsen, S.A.; Kramm, C.M. No Significant Cytotoxic Effect of the EZH2 Inhibitor Tazemetostat (EPZ-6438) on Pediatric Glioma Cells with Wildtype Histone 3 or Mutated Histone 3.3. Klin. Padiatr. 2016, 228, 113–117. [Google Scholar] [CrossRef] [PubMed]
  57. Xu, H.; Wang, J.; Zhang, K.; Zhao, M.; Ellenbroek, B.; Shao, F.; Wang, W. Effects of Adolescent Social Stress and Antidepressant Treatment on Cognitive Inflexibility and Bdnf Epigenetic Modifications in the MPFC of Adult Mice. Psychoneuroendocrinology 2018, 88, 92–101. [Google Scholar] [CrossRef] [PubMed]
  58. Morse, S.J.; Butler, A.A.; Davis, R.L.; Soller, I.J.; Lubin, F.D. Environmental Enrichment Reverses Histone Methylation Changes in the Aged Hippocampus and Restores Age-Related Memory Deficits. Biology 2015, 4, 298–313. [Google Scholar] [CrossRef]
  59. Neelamegam, R.; Ricq, E.L.; Malvaez, M.; Patnaik, D.; Norton, S.; Carlin, S.M.; Hill, I.T.; Wood, M.A.; Haggarty, S.J.; Hooker, J.M. Brain-Penetrant LSD1 Inhibitors Can Block Memory Consolidation. ACS Chem. Neurosci. 2012, 3, 120–128. [Google Scholar] [CrossRef]
  60. Korb, E.; Herre, M.; Zucker-Scharff, I.; Darnell, R.B.; Allis, C.D. BET Protein Brd4 Activates Transcription in Neurons and BET Inhibitor Jq1 Blocks Memory in Mice. Nat. Neurosci. 2015, 18, 1464–1473. [Google Scholar] [CrossRef]
  61. Li, J.; Ma, J.; Meng, G.; Lin, H.; Wu, S.; Wang, J.; Luo, J.; Xu, X.; Tough, D.; Lindon, M.; et al. BET Bromodomain Inhibition Promotes Neurogenesis While Inhibiting Gliogenesis in Neural Progenitor Cells. Stem Cell. Res. 2016, 17, 212–221. [Google Scholar] [CrossRef]
  62. Yu, L.; Wang, Z.; Zhang, Z.; Ren, X.; Lu, X.; Ding, K. Small-Molecule BET Inhibitors in Clinical and Preclinical Development and Their Therapeutic Potential. CTMC 2015, 15, 776–794. [Google Scholar] [CrossRef]
  63. Mota, M.; Porrini, V.; Parrella, E.; Benarese, M.; Bellucci, A.; Rhein, S.; Schwaninger, M.; Pizzi, M. Neuroprotective Epi-Drugs Quench the Inflammatory Response and Microglial/Macrophage Activation in a Mouse Model of Permanent Brain Ischemia. J. Neuroinflammation 2020, 17, 361. [Google Scholar] [CrossRef]
  64. Nuñez, J.K.; Chen, J.; Pommier, G.C.; Cogan, J.Z.; Replogle, J.M.; Adriaens, C.; Ramadoss, G.N.; Shi, Q.; Hung, K.L.; Samelson, A.J.; et al. Genome-Wide Programmable Transcriptional Memory by CRISPR-Based Epigenome Editing. Cell 2021, 184, 2503–2519.e17. [Google Scholar] [CrossRef]
  65. Peng, Q.; Weng, K.; Li, S.; Xu, R.; Wang, Y.; Wu, Y. A Perspective of Epigenetic Regulation in Radiotherapy. Front. Cell. Dev. Biol. 2021, 9, 624312. [Google Scholar] [CrossRef] [PubMed]
  66. Topper, M.J.; Vaz, M.; Marrone, K.A.; Brahmer, J.R.; Baylin, S.B. The Emerging Role of Epigenetic Therapeutics in Immuno-Oncology. Nat. Rev. Clin. Oncol. 2020, 17, 75–90. [Google Scholar] [CrossRef] [PubMed]
  67. Abballe, L.; Miele, E. Epigenetic Modulators for Brain Cancer Stem Cells: Implications for Anticancer Treatment. World J. Stem Cells 2021, 13, 670–684. [Google Scholar] [CrossRef] [PubMed]
  68. Bukowinski, A.; Chang, B.; Reid, J.M.; Liu, X.; Minard, C.G.; Trepel, J.B.; Lee, M.-J.; Fox, E.; Weigel, B.J. A Phase 1 Study of Entinostat in Children and Adolescents with Recurrent or Refractory Solid Tumors, Including CNS Tumors: Trial ADVL1513, Pediatric Early Phase-Clinical Trial Network (PEP-CTN). Pediatr. Blood Cancer 2021, 68, e28892. [Google Scholar] [CrossRef]
  69. Lee, E.Q.; Reardon, D.A.; Schiff, D.; Drappatz, J.; Muzikansky, A.; Grimm, S.A.; Norden, A.D.; Nayak, L.; Beroukhim, R.; Rinne, M.L.; et al. Phase II Study of Panobinostat in Combination with Bevacizumab for Recurrent Glioblastoma and Anaplastic Glioma. Neuro Oncol. 2015, 17, 862–867. [Google Scholar] [CrossRef] [PubMed]
  70. van Tilburg, C.M.; Witt, R.; Heiss, M.; Pajtler, K.W.; Plass, C.; Poschke, I.; Platten, M.; Harting, I.; Sedlaczek, O.; Freitag, A.; et al. INFORM2 NivEnt: The First Trial of the INFORM2 Biomarker Driven Phase I/II Trial Series: The Combination of Nivolumab and Entinostat in Children and Adolescents with Refractory High-Risk Malignancies. BMC Cancer 2020, 20, 523. [Google Scholar] [CrossRef]
  71. Buyandelger, B.; Bar, E.E.; Hung, K.-S.; Chen, R.-M.; Chiang, Y.-H.; Liou, J.-P.; Huang, H.-M.; Wang, J.-Y. Histone Deacetylase Inhibitor MPT0B291 Suppresses Glioma Growth in Vitro and in Vivo Partially through Acetylation of P53. Int. J. Biol. Sci. 2020, 16, 3184–3199. [Google Scholar] [CrossRef]
  72. Yin, C.; Li, P. Growth Suppression of Glioma Cells Using HDAC6 Inhibitor, Tubacin. Open. Med. 2018, 13, 221–226. [Google Scholar] [CrossRef]
  73. Auzmendi-Iriarte, J.; Saenz-Antoñanzas, A.; Mikelez-Alonso, I.; Carrasco-Garcia, E.; Tellaetxe-Abete, M.; Lawrie, C.H.; Sampron, N.; Cortajarena, A.L.; Matheu, A. Characterization of a New Small-Molecule Inhibitor of HDAC6 in Glioblastoma. Cell. Death Dis. 2020, 11, 417. [Google Scholar] [CrossRef]
  74. Nguyen, T.T.T.; Zhang, Y.; Shang, E.; Shu, C.; Torrini, C.; Zhao, J.; Bianchetti, E.; Mela, A.; Humala, N.; Mahajan, A.; et al. HDAC Inhibitors Elicit Metabolic Reprogramming by Targeting Super-Enhancers in Glioblastoma Models. J. Clin. Investig. 2020, 130, 3699–3716. [Google Scholar] [CrossRef]
  75. Funck-Brentano, E.; Vizlin-Hodzic, D.; Nilsson, J.A.; Nilsson, L.M. BET Bromodomain Inhibitor HMBA Synergizes with MEK Inhibition in Treatment of Malignant Glioma. Epigenetics 2021, 16, 54–63. [Google Scholar] [CrossRef] [PubMed]
  76. Wiese, M.; Hamdan, F.H.; Kubiak, K.; Diederichs, C.; Gielen, G.H.; Nussbaumer, G.; Carcaboso, A.M.; Hulleman, E.; Johnsen, S.A.; Kramm, C.M. Combined Treatment with CBP and BET Inhibitors Reverses Inadvertent Activation of Detrimental Super Enhancer Programs in DIPG Cells. Cell. Death Dis. 2020, 11, 673. [Google Scholar] [CrossRef]
  77. Lam, F.C.; Morton, S.W.; Wyckoff, J.; Vu Han, T.-L.; Hwang, M.K.; Maffa, A.; Balkanska-Sinclair, E.; Yaffe, M.B.; Floyd, S.R.; Hammond, P.T. Enhanced Efficacy of Combined Temozolomide and Bromodomain Inhibitor Therapy for Gliomas Using Targeted Nanoparticles. Nat. Commun. 2018, 9, 1991. [Google Scholar] [CrossRef] [PubMed]
  78. Estekizadeh, A.; Landázuri, N.; Pantalone, M.R.; Davoudi, B.; Hu, L.-F.; Nawaz, I.; Stragliotto, G.; Ekström, T.J.; Rahbar, A. 5-Azacytidine Treatment Results in Nuclear Exclusion of DNA Methyltransferase-1, as Well as Reduced Proliferation and Invasion in Human Cytomegalovirus-infected Glioblastoma Cells. Oncol. Rep. 2019, 41, 2927–2936. [Google Scholar] [CrossRef] [PubMed]
  79. Turcan, S.; Fabius, A.W.M.; Borodovsky, A.; Pedraza, A.; Brennan, C.; Huse, J.; Viale, A.; Riggins, G.J.; Chan, T.A. Efficient Induction of Differentiation and Growth Inhibition in IDH1 Mutant Glioma Cells by the DNMT Inhibitor Decitabine. Oncotarget 2013, 4, 1729–1736. [Google Scholar] [CrossRef] [PubMed]
  80. Gupta, R.; Ambasta, R.K.; Kumar, P. Pharmacological Intervention of Histone Deacetylase Enzymes in the Neurodegenerative Disorders. Life Sci. 2020, 243, 117278. [Google Scholar] [CrossRef] [PubMed]
  81. Kim, H.J.; Rowe, M.; Ren, M.; Hong, J.-S.; Chen, P.-S.; Chuang, D.-M. Histone Deacetylase Inhibitors Exhibit Anti-Inflammatory and Neuroprotective Effects in a Rat Permanent Ischemic Model of Stroke: Multiple Mechanisms of Action. J. Pharmacol. Exp. Ther. 2007, 321, 892–901. [Google Scholar] [CrossRef]
  82. Qu, X.; Neuhoff, C.; Cinar, M.U.; Pröll, M.; Tholen, E.; Tesfaye, D.; Hölker, M.; Schellander, K.; Uddin, M.J. Epigenetic Modulation of TLR4 Expression by Sulforaphane Increases Anti-Inflammatory Capacity in Porcine Monocyte-Derived Dendritic Cells. Biology 2021, 10, 490. [Google Scholar] [CrossRef]
  83. Wang, P.; Zhang, Y.; Gong, Y.; Yang, R.; Chen, Z.; Hu, W.; Wu, Y.; Gao, M.; Xu, X.; Qin, Y.; et al. Sodium Butyrate Triggers a Functional Elongation of Microglial Process via Akt-Small RhoGTPase Activation and HDACs Inhibition. Neurobiol. Dis. 2018, 111, 12–25. [Google Scholar] [CrossRef]
  84. Wu, Y.; Dou, J.; Wan, X.; Leng, Y.; Liu, X.; Chen, L.; Shen, Q.; Zhao, B.; Meng, Q.; Hou, J. Histone Deacetylase Inhibitor MS-275 Alleviates Postoperative Cognitive Dysfunction in Rats by Inhibiting Hippocampal Neuroinflammation. Neuroscience 2019, 417, 70–80. [Google Scholar] [CrossRef]
  85. Zhang, B.; West, E.J.; Van, K.C.; Gurkoff, G.G.; Zhou, J.; Zhang, X.-M.; Kozikowski, A.P.; Lyeth, B.G. HDAC Inhibitor Increases Histone H3 Acetylation and Reduces Microglia Inflammatory Response Following Traumatic Brain Injury in Rats. Brain Res. 2008, 1226, 181–191. [Google Scholar] [CrossRef] [PubMed]
  86. Sartor, G.C.; Malvezzi, A.M.; Kumar, A.; Andrade, N.S.; Wiedner, H.J.; Vilca, S.J.; Janczura, K.J.; Bagheri, A.; Al-Ali, H.; Powell, S.K.; et al. Enhancement of BDNF Expression and Memory by HDAC Inhibition Requires BET Bromodomain Reader Proteins. J. Neurosci. 2019, 39, 612–626. [Google Scholar] [CrossRef]
  87. Seo, M.K.; Kim, Y.H.; McIntyre, R.S.; Mansur, R.B.; Lee, Y.; Carmona, N.E.; Choi, A.J.; Kim, G.-M.; Lee, J.G.; Park, S.W. Effects of Antipsychotic Drugs on the Epigenetic Modification of Brain-Derived Neurotrophic Factor Gene Expression in the Hippocampi of Chronic Restraint Stress Rats. Neural Plast. 2018, 2018, 2682037. [Google Scholar] [CrossRef] [PubMed]
  88. Kao, C.-Y.; Hsu, Y.-C.; Liu, J.-W.; Lee, D.-C.; Chung, Y.-F.; Chiu, I.-M. The Mood Stabilizer Valproate Activates Human FGF1 Gene Promoter through Inhibiting HDAC and GSK-3 Activities. J. Neurochem. 2013, 126, 4–18. [Google Scholar] [CrossRef] [PubMed]
  89. Shukla, S.; Chaurasiya, N.; Walker, L.; Tekwani, B. Vorinostat, a Pan Histone Deacetylase Inhibitor, Stimulates Neuritic Outgrowth in PC12 Cells via Activation of MAPK and PI3K Signal Transduction Pathways. Planta Med. 2013, 79, P90. [Google Scholar] [CrossRef]
  90. Ye, J. Improving Insulin Sensitivity with HDAC Inhibitor. Diabetes 2013, 62, 685–687. [Google Scholar] [CrossRef] [PubMed]
  91. Levine, A.; Worrell, T.R.; Zimnisky, R.; Schmauss, C. Early Life Stress Triggers Sustained Changes in Histone Deacetylase Expression and Histone H4 Modifications That Alter Responsiveness to Adolescent Antidepressant Treatment. Neurobiol. Dis. 2012, 45, 488–498. [Google Scholar] [CrossRef]
  92. Singh, H.; Wray, N.; Schappi, J.M.; Rasenick, M.M. Disruption of Lipid-Raft Localized Gαs/Tubulin Complexes by Antidepressants: A Unique Feature of HDAC6 Inhibitors, SSRI and Tricyclic Compounds. Neuropsychopharmacology 2018, 43, 1481–1491. [Google Scholar] [CrossRef]
  93. Wu, D.; Yan, Y.; Wei, T.; Ye, Z.; Xiao, Y.; Pan, Y.; Orme, J.J.; Wang, D.; Wang, L.; Ren, S.; et al. An Acetyl-Histone Vulnerability in PI3K/AKT Inhibition-Resistant Cancers Is Targetable by Both BET and HDAC Inhibitors. Cell. Rep. 2021, 34, 108744. [Google Scholar] [CrossRef]
  94. Leng, Y.; Chuang, D.-M. Endogenous Alpha-Synuclein Is Induced by Valproic Acid through Histone Deacetylase Inhibition and Participates in Neuroprotection against Glutamate-Induced Excitotoxicity. J. Neurosci. 2006, 26, 7502–7512. [Google Scholar] [CrossRef]
  95. Lu, J.; Frerich, J.M.; Turtzo, L.C.; Li, S.; Chiang, J.; Yang, C.; Wang, X.; Zhang, C.; Wu, C.; Sun, Z.; et al. Histone Deacetylase Inhibitors Are Neuroprotective and Preserve NGF-Mediated Cell Survival Following Traumatic Brain Injury. Proc. Natl. Acad. Sci. USA 2013, 110, 10747–10752. [Google Scholar] [CrossRef] [PubMed]
  96. Ryu, H.; Lee, J.; Olofsson, B.A.; Mwidau, A.; Dedeoglu, A.; Escudero, M.; Flemington, E.; Azizkhan-Clifford, J.; Ferrante, R.J.; Ratan, R.R.; et al. Histone Deacetylase Inhibitors Prevent Oxidative Neuronal Death Independent of Expanded Polyglutamine Repeats via an Sp1-Dependent Pathway. Proc. Natl. Acad. Sci. USA 2003, 100, 4281–4286. [Google Scholar] [CrossRef] [PubMed]
  97. Uittenbogaard, M.; Brantner, C.A.; Chiaramello, A. Epigenetic Modifiers Promote Mitochondrial Biogenesis and Oxidative Metabolism Leading to Enhanced Differentiation of Neuroprogenitor Cells. Cell. Death Dis. 2018, 9, 360. [Google Scholar] [CrossRef]
  98. Silva, M.R.; Correia, A.O.; Dos Santos, G.C.A.; Parente, L.L.T.; de Siqueira, K.P.; Lima, D.G.S.; Moura, J.A.; da Silva Ribeiro, A.E.; Costa, R.O.; Lucetti, D.L.; et al. Neuroprotective Effects of Valproic Acid on Brain Ischemia Are Related to Its HDAC and GSK3 Inhibitions. Pharmacol. Biochem. Behav. 2018, 167, 17–28. [Google Scholar] [CrossRef]
  99. Balasubramaniyan, V.; Boddeke, E.; Bakels, R.; Küst, B.; Kooistra, S.; Veneman, A.; Copray, S. Effects of Histone Deacetylation Inhibition on Neuronal Differentiation of Embryonic Mouse Neural Stem Cells. Neuroscience 2006, 143, 939–951. [Google Scholar] [CrossRef]
  100. Hsieh, J.; Nakashima, K.; Kuwabara, T.; Mejia, E.; Gage, F.H. Histone Deacetylase Inhibition-Mediated Neuronal Differentiation of Multipotent Adult Neural Progenitor Cells. Proc. Natl. Acad. Sci. USA 2004, 101, 16659–16664. [Google Scholar] [CrossRef] [PubMed]
  101. Lawlor, L.; Yang, X.B. Harnessing the HDAC–Histone Deacetylase Enzymes, Inhibitors and How These Can Be Utilised in Tissue Engineering. Int. J. Oral. Sci. 2019, 11, 20. [Google Scholar] [CrossRef]
  102. Medvedev, S.P.; Pokushalov, E.A.; Zakian, S.M. Epigenetics of Pluripotent Cells. Acta Nat. 2012, 4, 28–46. [Google Scholar] [CrossRef]
  103. Park, G.; Tan, J.; Garcia, G.; Kang, Y.; Salvesen, G.; Zhang, Z. Regulation of Histone Acetylation by Autophagy in Parkinson Disease. J. Biol. Chem. 2016, 291, 3531–3540. [Google Scholar] [CrossRef]
  104. Bardai, F.H.; Price, V.; Zaayman, M.; Wang, L.; D’Mello, S.R. Histone Deacetylase-1 (HDAC1) Is a Molecular Switch between Neuronal Survival and Death. J. Biol. Chem. 2012, 287, 35444–35453. [Google Scholar] [CrossRef]
  105. Konsoula, Z.; Barile, F.A. Epigenetic Histone Acetylation and Deacetylation Mechanisms in Experimental Models of Neurodegenerative Disorders. J. Pharmacol. Toxicol. Methods 2012, 66, 215–220. [Google Scholar] [CrossRef]
  106. Allain, H.; Denmat, J.; Bentue-Ferrer, D.; Milon, D.; Pignol, P.; Reymann, J.M.; Pape, D.; Sabouraud, O.; Driessche, J. RANDOMIZED, DOUBLE-BLIND TRIAL OF EXIFONE VERSUS COGNITIVE PROBLEMS IN PARKINSON’S DISEASE. Fundam. Clin. Pharmacol. 1988, 2, 1–12. [Google Scholar] [CrossRef] [PubMed]
  107. Pao, P.-C.; Patnaik, D.; Watson, L.A.; Gao, F.; Pan, L.; Wang, J.; Adaikkan, C.; Penney, J.; Cam, H.P.; Huang, W.-C.; et al. HDAC1 Modulates OGG1-Initiated Oxidative DNA Damage Repair in the Aging Brain and Alzheimer’s Disease. Nat. Commun. 2020, 11, 2484. [Google Scholar] [CrossRef] [PubMed]
  108. Puttagunta, R.; Tedeschi, A.; Sória, M.G.; Hervera, A.; Lindner, R.; Rathore, K.I.; Gaub, P.; Joshi, Y.; Nguyen, T.; Schmandke, A.; et al. PCAF-Dependent Epigenetic Changes Promote Axonal Regeneration in the Central Nervous System. Nat. Commun. 2014, 5, 3527. [Google Scholar] [CrossRef] [PubMed]
  109. Finelli, M.J.; Wong, J.K.; Zou, H. Epigenetic Regulation of Sensory Axon Regeneration after Spinal Cord Injury. J. Neurosci. 2013, 33, 19664–19676. [Google Scholar] [CrossRef] [PubMed]
  110. Bayraktar, G.; Kreutz, M.R. Neuronal DNA Methyltransferases: Epigenetic Mediators between Synaptic Activity and Gene Expression? Neuroscientist 2018, 24, 171–185. [Google Scholar] [CrossRef]
  111. Morris, M.J.; Monteggia, L.M. Role of DNA Methylation and the DNA Methyltransferases in Learning and Memory. Dialogues Clin. Neurosci. 2014, 16, 359–371. [Google Scholar] [CrossRef]
  112. Baek, M.; Yoo, E.; Choi, H.I.; An, G.Y.; Chai, J.C.; Lee, Y.S.; Jung, K.H.; Chai, Y.G. The BET Inhibitor Attenuates the Inflammatory Response and Cell Migration in Human Microglial HMC3 Cell Line. Sci. Rep. 2021, 11, 8828. [Google Scholar] [CrossRef]
  113. DeMars, K.M.; Yang, C.; Candelario-Jalil, E. Neuroprotective Effects of Targeting BET Proteins for Degradation with DBET1 in Aged Mice Subjected to Ischemic Stroke. Neurochem. Int. 2019, 127, 94–102. [Google Scholar] [CrossRef]
  114. DeMars, K.M.; Yang, C.; Castro-Rivera, C.I.; Candelario-Jalil, E. Selective Degradation of BET Proteins with DBET1, a Proteolysis-Targeting Chimera, Potently Reduces pro-Inflammatory Responses in Lipopolysaccharide-Activated Microglia. Biochem. Biophys. Res. Commun. 2018, 497, 410–415. [Google Scholar] [CrossRef]
  115. Zhou, Y.; Gu, Y.; Liu, J. BRD4 Suppression Alleviates Cerebral Ischemia-Induced Brain Injury by Blocking Glial Activation via the Inhibition of Inflammatory Response and Pyroptosis. Biochem. Biophys. Res. Commun. 2019, 519, 481–488. [Google Scholar] [CrossRef] [PubMed]
  116. Sánchez-Ventura, J.; Amo-Aparicio, J.; Navarro, X.; Penas, C. BET Protein Inhibition Regulates Cytokine Production and Promotes Neuroprotection after Spinal Cord Injury. J. Neuroinflammation 2019, 16, 124. [Google Scholar] [CrossRef] [PubMed]
  117. Francelle, L.; Outeiro, T.F.; Rappold, G.A. Inhibition of HDAC6 Activity Protects Dopaminergic Neurons from Alpha-Synuclein Toxicity. Sci. Rep. 2020, 10, 6064. [Google Scholar] [CrossRef] [PubMed]
  118. Yeh, H.H.; Young, D.; Gelovani, J.G.; Robinson, A.; Davidson, Y.; Herholz, K.; Mann, D.M.A. Histone Deacetylase Class II and Acetylated Core Histone Immunohistochemistry in Human Brains with Huntington’s Disease. Brain Res. 2013, 1504, 16–24. [Google Scholar] [CrossRef]
  119. Huang, M.; Lou, D.; Charli, A.; Kong, D.; Jin, H.; Zenitsky, G.; Anantharam, V.; Kanthasamy, A.; Wang, Z.; Kanthasamy, A.G. Mitochondrial Dysfunction-Induced H3K27 Hyperacetylation Perturbs Enhancers in Parkinson’s Disease. JCI Insight 2021, 6, e138088. [Google Scholar] [CrossRef]
  120. Toker, L.; Tran, G.T.; Sundaresan, J.; Tysnes, O.-B.; Alves, G.; Haugarvoll, K.; Nido, G.S.; Dölle, C.; Tzoulis, C. Genome-Wide Histone Acetylation Analysis Reveals Altered Transcriptional Regulation in the Parkinson’s Disease Brain. Mol. Neurodegener. 2021, 16, 31. [Google Scholar] [CrossRef]
  121. Mahady, L.; Nadeem, M.; Malek-Ahmadi, M.; Chen, K.; Perez, S.E.; Mufson, E.J. Frontal Cortex Epigenetic Dysregulation During the Progression of Alzheimer’s Disease. J. Alzheimers Dis. 2018, 62, 115–131. [Google Scholar] [CrossRef]
  122. Nativio, R.; Donahue, G.; Berson, A.; Lan, Y.; Amlie-Wolf, A.; Tuzer, F.; Toledo, J.B.; Gosai, S.J.; Gregory, B.D.; Torres, C.; et al. Dysregulation of the Epigenetic Landscape of Normal Aging in Alzheimer’s Disease. Nat. Neurosci. 2018, 21, 497–505. [Google Scholar] [CrossRef]
  123. Zhang, K.; Schrag, M.; Crofton, A.; Trivedi, R.; Vinters, H.; Kirsch, W. Targeted Proteomics for Quantification of Histone Acetylation in Alzheimer’s Disease. Proteomics 2012, 12, 1261–1268. [Google Scholar] [CrossRef]
  124. Bennett, S.A.; Tanaz, R.; Cobos, S.N.; Torrente, M.P. Epigenetics in Amyotrophic Lateral Sclerosis: A Role for Histone Post-Translational Modifications in Neurodegenerative Disease. Transl. Res. 2019, 204, 19–30. [Google Scholar] [CrossRef]
  125. Chuang, D.-M.; Leng, Y.; Marinova, Z.; Kim, H.-J.; Chiu, C.-T. Multiple Roles of HDAC Inhibition in Neurodegenerative Conditions. Trends Neurosci. 2009, 32, 591–601. [Google Scholar] [CrossRef] [PubMed]
  126. Pirooznia, S.K.; Elefant, F. Targeting Specific HATs for Neurodegenerative Disease Treatment: Translating Basic Biology to Therapeutic Possibilities. Front. Cell. Neurosci. 2013, 7, 30. [Google Scholar] [CrossRef] [PubMed]
  127. Santana, D.A.; Smith, M.d.A.C.; Chen, E.S. Histone Modifications in Alzheimer’s Disease. Genes 2023, 14, 347. [Google Scholar] [CrossRef]
  128. Tejido, C.; Pakravan, D.; Bosch, L.V.D. Potential Therapeutic Role of HDAC Inhibitors in FUS-ALS. Front. Mol. Neurosci. 2021, 14, 686995. [Google Scholar] [CrossRef] [PubMed]
  129. Ricobaraza, A.; Cuadrado-Tejedor, M.; Pérez-Mediavilla, A.; Frechilla, D.; Del Río, J.; García-Osta, A. Phenylbutyrate Ameliorates Cognitive Deficit and Reduces Tau Pathology in an Alzheimer’s Disease Mouse Model. Neuropsychopharmacology 2009, 34, 1721–1732. [Google Scholar] [CrossRef] [PubMed]
  130. Han, Y.; Chen, L.; Guo, Y.; Wang, C.; Zhang, C.; Kong, L.; Ma, H. Class I HDAC Inhibitor Improves Synaptic Proteins and Repairs Cytoskeleton Through Regulating Synapse-Related Genes In Vitro and In Vivo. Front. Aging Neurosci. 2020, 12, 619866. [Google Scholar] [CrossRef] [PubMed]
  131. Huang, H.-J.; Huang, H.-Y.; Hsieh-Li, H.M. MGCD0103, a Selective Histone Deacetylase Inhibitor, Coameliorates Oligomeric Aβ25-35 -Induced Anxiety and Cognitive Deficits in a Mouse Model. CNS Neurosci. Ther. 2019, 25, 175–186. [Google Scholar] [CrossRef] [PubMed]
  132. Dresselhaus, E.; Duerr, J.M.; Vincent, F.; Sylvain, E.K.; Beyna, M.; Lanyon, L.F.; LaChapelle, E.; Pettersson, M.; Bales, K.R.; Ramaswamy, G. Class I HDAC Inhibition Is a Novel Pathway for Regulating Astrocytic ApoE Secretion. PLoS ONE 2018, 13, e0194661. [Google Scholar] [CrossRef]
  133. Corder, E.H.; Saunders, A.M.; Strittmatter, W.J.; Schmechel, D.E.; Gaskell, P.C.; Small, G.W.; Roses, A.D.; Haines, J.L.; Pericak-Vance, M.A. Gene Dose of Apolipoprotein E Type 4 Allele and the Risk of Alzheimer’s Disease in Late Onset Families. Science 1993, 261, 921–923. [Google Scholar] [CrossRef]
  134. Harrison, I.F.; Smith, A.D.; Dexter, D.T. Pathological Histone Acetylation in Parkinson’s Disease: Neuroprotection and Inhibition of Microglial Activation through SIRT 2 Inhibition. Neurosci. Lett. 2018, 666, 48–57. [Google Scholar] [CrossRef]
  135. 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 α-Synuclein-Induced Degeneration in Cellular Models of Parkinson’s Disease. Mol. Cell. Neurosci. 2021, 115, 103642. [Google Scholar] [CrossRef] [PubMed]
  136. Chopra, V.; Quinti, L.; Khanna, P.; Paganetti, P.; Kuhn, R.; Young, A.B.; Kazantsev, A.G.; Hersch, S. LBH589, A Hydroxamic Acid-Derived HDAC Inhibitor, Is Neuroprotective in Mouse Models of Huntington’s Disease. J. Huntingt. Dis. 2016, 5, 347–355. [Google Scholar] [CrossRef] [PubMed]
  137. Naia, L.; Cunha-Oliveira, T.; Rodrigues, J.; Rosenstock, T.R.; Oliveira, A.; Ribeiro, M.; Carmo, C.; Oliveira-Sousa, S.I.; Duarte, A.I.; Hayden, M.R.; et al. Histone Deacetylase Inhibitors Protect Against Pyruvate Dehydrogenase Dysfunction in Huntington’s Disease. J. Neurosci. 2017, 37, 2776–2794. [Google Scholar] [CrossRef] [PubMed]
  138. Suelves, N.; Kirkham-McCarthy, L.; Lahue, R.S.; Ginés, S. A Selective Inhibitor of Histone Deacetylase 3 Prevents Cognitive Deficits and Suppresses Striatal CAG Repeat Expansions in Huntington’s Disease Mice. Sci. Rep. 2017, 7, 6082. [Google Scholar] [CrossRef] [PubMed]
  139. 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. [Google Scholar] [CrossRef] [PubMed]
  140. Yazar, V.; Dawson, V.L.; Dawson, T.M.; Kang, S.-U. DNA Methylation Signature of Aging: Potential Impact on the Pathogenesis of Parkinson’s Disease. J. Park. Dis. 2023, 13, 145–164. [Google Scholar] [CrossRef]
  141. Wang, S.-C.; Oelze, B.; Schumacher, A. Age-Specific Epigenetic Drift in Late-Onset Alzheimer’s Disease. PLoS ONE 2008, 3, e2698. [Google Scholar] [CrossRef]
  142. Available online: https://www.alzforum.org/therapeutics/vafidemstat (accessed on 31 December 2022).
  143. Clinical Trials for Ory-2001. Available online: https://www.clinicaltrialsregister.eu/ctr-search/search?query=ory-2001 (accessed on 31 December 2022).
  144. Magistri, M.; Velmeshev, D.; Makhmutova, M.; Patel, P.; Sartor, G.C.; Volmar, C.-H.; Wahlestedt, C.; Faghihi, M.A. The BET-Bromodomain Inhibitor JQ1 Reduces Inflammation and Tau Phosphorylation at Ser396 in the Brain of the 3xTg Model of Alzheimer’s Disease. Curr. Alzheimer Res. 2016, 13, 985–995. [Google Scholar] [CrossRef]
  145. Benito, E.; Ramachandran, B.; Schroeder, H.; Schmidt, G.; Urbanke, H.; Burkhardt, S.; Capece, V.; Dean, C.; Fischer, A. The BET/BRD Inhibitor JQ1 Improves Brain Plasticity in WT and APP Mice. Transl. Psychiatry 2017, 7, e1239. [Google Scholar] [CrossRef]
  146. Cummings, J.; Schwartz, G.G.; Nicholls, S.J.; Khan, A.; Halliday, C.; Toth, P.P.; Sweeney, M.; Johansson, J.O.; Wong, N.C.W.; Kulikowski, E.; et al. Cognitive Effects of the BET Protein Inhibitor Apabetalone: A Prespecified Montreal Cognitive Assessment Analysis Nested in the BETonMACE Randomized Controlled Trial. J. Alzheimers Dis. 2021, 83, 1703–1715. [Google Scholar] [CrossRef]
  147. Carlson-Stevermer, J.; Kelso, R.; Kadina, A.; Joshi, S.; Rossi, N.; Walker, J.; Stoner, R.; Maures, T. CRISPRoff Enables Spatio-Temporal Control of CRISPR Editing. Nat. Commun. 2020, 11, 5041. [Google Scholar] [CrossRef] [PubMed]
  148. Moussa, H.F.; Angstman, J.F.; Khalil, A.S. Here to Stay: Writing Lasting Epigenetic Memories. Cell 2021, 184, 2281–2283. [Google Scholar] [CrossRef] [PubMed]
  149. Pardridge, W.M. Brain Delivery of Nanomedicines: Trojan Horse Liposomes for Plasmid DNA Gene Therapy of the Brain. Front. Med. Technol. 2020, 2, 602236. [Google Scholar] [CrossRef]
  150. Wilkinson, M.B.; Xiao, G.; Kumar, A.; LaPlant, Q.; Renthal, W.; Sikder, D.; Kodadek, T.J.; Nestler, E.J. Imipramine Treatment and Resiliency Exhibit Similar Chromatin Regulation in the Mouse Nucleus Accumbens in Depression Models. J. Neurosci. 2009, 29, 7820–7832. [Google Scholar] [CrossRef] [PubMed]
  151. Jochim, J.; Rifkin-Zybutz, R.P.; Geddes, J.; Cipriani, A. Valproate for Acute Mania. Cochrane Database Syst. Rev. 2019, 10, CD004052. [Google Scholar] [CrossRef]
  152. Wang, Y.; Xia, J.; Helfer, B.; Li, C.; Leucht, S. Valproate for Schizophrenia. Cochrane Database Syst. Rev. 2016, 11, CD004028. [Google Scholar] [CrossRef] [PubMed]
  153. Boks, M.P.; de Jong, N.M.; Kas, M.J.H.; Vinkers, C.H.; Fernandes, C.; Kahn, R.S.; Mill, J.; Ophoff, R.A. Current Status and Future Prospects for Epigenetic Psychopharmacology. Epigenetics 2012, 7, 20–28. [Google Scholar] [CrossRef]
  154. Gassen, N.C.; Fries, G.R.; Zannas, A.S.; Hartmann, J.; Zschocke, J.; Hafner, K.; Carrillo-Roa, T.; Steinbacher, J.; Preißinger, S.N.; Hoeijmakers, L.; et al. Chaperoning Epigenetics: FKBP51 Decreases the Activity of DNMT1 and Mediates Epigenetic Effects of the Antidepressant Paroxetine. Sci. Signal. 2015, 8, ra119. [Google Scholar] [CrossRef]
  155. Glover, M.E.; McCoy, C.R.; Shupe, E.A.; Unroe, K.A.; Jackson, N.L.; Clinton, S.M. Perinatal Exposure to the SSRI Paroxetine Alters the Methylome Landscape of the Developing Dentate Gyrus. Eur. J. Neurosci. 2019, 50, 1843–1870. [Google Scholar] [CrossRef]
  156. Toth, M. Epigenetic Neuropharmacology: Drugs Affecting the Epigenome in the Brain. Annu. Rev. Pharmacol. Toxicol. 2021, 61, 181–201. [Google Scholar] [CrossRef]
  157. Fukuchi, M.; Nii, T.; Ishimaru, N.; Minamino, A.; Hara, D.; Takasaki, I.; Tabuchi, A.; Tsuda, M. Valproic Acid Induces Up- or down-Regulation of Gene Expression Responsible for the Neuronal Excitation and Inhibition in Rat Cortical Neurons through Its Epigenetic Actions. Neurosci. Res. 2009, 65, 35–43. [Google Scholar] [CrossRef] [PubMed]
  158. Kurita, M.; Holloway, T.; García-Bea, A.; Kozlenkov, A.; Friedman, A.K.; Moreno, J.L.; Heshmati, M.; Golden, S.A.; Kennedy, P.J.; Takahashi, N.; et al. HDAC2 Regulates Atypical Antipsychotic Responses through the Modulation of MGlu2 Promoter Activity. Nat. Neurosci. 2012, 15, 1245–1254. [Google Scholar] [CrossRef]
  159. Agudelo, M.; Yoo, C.; Nair, M.P. Alcohol-Induced Serotonergic Modulation: The Role of Histone Deacetylases. Alcohol 2012, 46, 635–642. [Google Scholar] [CrossRef] [PubMed]
  160. Phi van, D.K.; Mühlbauer, E.; Phi-van, L. Histone Deacetylase HDAC1 Downregulates Transcription of the Serotonin Transporter (5-HTT) Gene in Tumor Cells. Biochim. Biophys. Acta 2015, 1849, 909–918. [Google Scholar] [CrossRef] [PubMed]
  161. Seo, M.K.; Lee, J.G.; Park, S.W. Early Life Stress Induces Age-Dependent Epigenetic Changes in P11 Gene Expression in Male Mice. Sci. Rep. 2021, 11, 10663. [Google Scholar] [CrossRef] [PubMed]
  162. Onishchenko, N.; Karpova, N.; Sabri, F.; Castrén, E.; Ceccatelli, S. Long-Lasting Depression-like Behavior and Epigenetic Changes of BDNF Gene Expression Induced by Perinatal Exposure to Methylmercury. J. Neurochem. 2008, 106, 1378–1387. [Google Scholar] [CrossRef] [PubMed]
  163. Sales, A.J.; Biojone, C.; Terceti, M.S.; Guimarães, F.S.; Gomes, M.V.M.; Joca, S.R.L. Antidepressant-like Effect Induced by Systemic and Intra-Hippocampal Administration of DNA Methylation Inhibitors. Br. J. Pharmacol. 2011, 164, 1711–1721. [Google Scholar] [CrossRef]
  164. Wang, J.; Hodes, G.E.; Zhang, H.; Zhang, S.; Zhao, W.; Golden, S.A.; Bi, W.; Menard, C.; Kana, V.; Leboeuf, M.; et al. Epigenetic Modulation of Inflammation and Synaptic Plasticity Promotes Resilience against Stress in Mice. Nat. Commun. 2018, 9, 477. [Google Scholar] [CrossRef]
  165. Matosin, N.; Halldorsdottir, T.; Binder, E.B. Understanding the Molecular Mechanisms Underpinning Gene by Environment Interactions in Psychiatric Disorders: The FKBP5 Model. Biol. Psychiatry 2018, 83, 821–830. [Google Scholar] [CrossRef]
  166. Yasuda, S.; Liang, M.-H.; Marinova, Z.; Yahyavi, A.; Chuang, D.-M. The Mood Stabilizers Lithium and Valproate Selectively Activate the Promoter IV of Brain-Derived Neurotrophic Factor in Neurons. Mol. Psychiatry 2009, 14, 51–59. [Google Scholar] [CrossRef]
  167. Alexander, N.; Kirschbaum, C.; Wankerl, M.; Stauch, B.J.; Stalder, T.; Steudte-Schmiedgen, S.; Muehlhan, M.; Miller, R. Glucocorticoid Receptor Gene Methylation Moderates the Association of Childhood Trauma and Cortisol Stress Reactivity. Psychoneuroendocrinology 2018, 90, 68–75. [Google Scholar] [CrossRef] [PubMed]
  168. Farrell, C.; Doolin, K.; O’ Leary, N.; Jairaj, C.; Roddy, D.; Tozzi, L.; Morris, D.; Harkin, A.; Frodl, T.; Nemoda, Z.; et al. DNA Methylation Differences at the Glucocorticoid Receptor Gene in Depression Are Related to Functional Alterations in Hypothalamic-Pituitary-Adrenal Axis Activity and to Early Life Emotional Abuse. Psychiatry Res. 2018, 265, 341–348. [Google Scholar] [CrossRef] [PubMed]
  169. Wadji, D.L.; Tandon, T.; Ketcha Wanda, G.J.M.; Wicky, C.; Dentz, A.; Hasler, G.; Morina, N.; Martin-Soelch, C. Child Maltreatment and NR3C1 Exon 1F Methylation, Link with Deregulated Hypothalamus-Pituitary-Adrenal Axis and Psychopathology: A Systematic Review. Child. Abuse Negl. 2021, 122, 105304. [Google Scholar] [CrossRef]
  170. Oliveira, A.M.M.; Hemstedt, T.J.; Bading, H. Rescue of Aging-Associated Decline in Dnmt3a2 Expression Restores Cognitive Abilities. Nat. Neurosci. 2012, 15, 1111–1113. [Google Scholar] [CrossRef]
  171. Argyrousi, E.K.; de Nijs, L.; Lagatta, D.C.; Schlütter, A.; Weidner, M.T.; Zöller, J.; van Goethem, N.P.; Joca, S.R.L.; van den Hove, D.L.A.; Prickaerts, J. Effects of DNA Methyltransferase Inhibition on Pattern Separation Performance in Mice. Neurobiol. Learn. Mem. 2019, 159, 6–15. [Google Scholar] [CrossRef] [PubMed]
  172. Maddox, S.A.; Schafe, G.E. Epigenetic Alterations in the Lateral Amygdala Are Required for Reconsolidation of a Pavlovian Fear Memory. Learn. Mem. 2011, 18, 579–593. [Google Scholar] [CrossRef]
  173. Miller, C.A.; Gavin, C.F.; White, J.A.; Parrish, R.R.; Honasoge, A.; Yancey, C.R.; Rivera, I.M.; Rubio, M.D.; Rumbaugh, G.; Sweatt, J.D. Cortical DNA Methylation Maintains Remote Memory. Nat. Neurosci. 2010, 13, 664–666. [Google Scholar] [CrossRef]
  174. Miller, C.A.; Campbell, S.L.; Sweatt, J.D. DNA Methylation and Histone Acetylation Work in Concert to Regulate Memory Formation and Synaptic Plasticity. Neurobiol. Learn. Mem. 2008, 89, 599–603. [Google Scholar] [CrossRef]
  175. Miller, C.A.; Sweatt, J.D. Covalent Modification of DNA Regulates Memory Formation. Neuron 2007, 53, 857–869. [Google Scholar] [CrossRef]
  176. Duan, Q.; Huang, F.-L.; Li, S.-J.; Chen, K.-Z.; Gong, L.; Qi, J.; Yang, Z.-H.; Yang, T.-L.; Li, F.; Li, C.-Q. BET Proteins Inhibitor JQ-1 Impaired the Extinction of Remote Auditory Fear Memory: An Effect Mediated by Insulin like Growth Factor 2. Neuropharmacology 2020, 177, 108255. [Google Scholar] [CrossRef]
  177. Wang, W.; Wang, R.; Jiang, Z.; Li, H.; Zhu, Z.; Khalid, A.; Liu, D.; Pan, F. Inhibiting Brd4 Alleviated PTSD-like Behaviors and Fear Memory through Regulating Immediate Early Genes Expression and Neuroinflammation in Rats. J. Neurochem. 2021, 158, 912–927. [Google Scholar] [CrossRef] [PubMed]
  178. Hoffmann, A.; Spengler, D. Chromatin Remodeling Complex NuRD in Neurodevelopment and Neurodevelopmental Disorders. Front. Genet. 2019, 10, 682. [Google Scholar] [CrossRef]
  179. Kozikowski, A.P.; Shen, S.; Pardo, M.; Tavares, M.T.; Szarics, D.; Benoy, V.; Zimprich, C.A.; Kutil, Z.; Zhang, G.; Bařinka, C.; et al. Brain Penetrable Histone Deacetylase 6 Inhibitor SW-100 Ameliorates Memory and Learning Impairments in a Mouse Model of Fragile X Syndrome. ACS Chem. Neurosci. 2019, 10, 1679–1695. [Google Scholar] [CrossRef] [PubMed]
  180. Jamal, I.; Kumar, V.; Vatsa, N.; Shekhar, S.; Singh, B.K.; Sharma, A.; Jana, N.R. Rescue of Altered HDAC Activity Recovers Behavioural Abnormalities in a Mouse Model of Angelman Syndrome. Neurobiol. Dis. 2017, 105, 99–108. [Google Scholar] [CrossRef] [PubMed]
  181. Lamar, K.-M.J.; Carvill, G.L. Free as a BRD4: Bromodomain Inhibition Ameliorates Disease Phenotypes in a Model of MECP2 Deficiency and Is a Potential Therapy for Rett Syndrome. Epilepsy Curr. 2020, 20, 390–392. [Google Scholar] [CrossRef]
  182. Sullivan, J.M.; Badimon, A.; Schaefer, U.; Ayata, P.; Gray, J.; Chung, C.; von Schimmelmann, M.; Zhang, F.; Garton, N.; Smithers, N.; et al. Autism-like Syndrome Is Induced by Pharmacological Suppression of BET Proteins in Young Mice. J. Exp. Med. 2015, 212, 1771–1781. [Google Scholar] [CrossRef]
  183. Gowen, A.M.; Odegaard, K.E.; Hernandez, J.; Chand, S.; Koul, S.; Pendyala, G.; Yelamanchili, S.V. Role of MicroRNAs in the Pathophysiology of Addiction. Wiley Interdiscip. Rev. RNA 2021, 12, e1637. [Google Scholar] [CrossRef]
  184. Ju, C.; Fiori, L.M.; Belzeaux, R.; Theroux, J.-F.; Chen, G.G.; Aouabed, Z.; Blier, P.; Farzan, F.; Frey, B.N.; Giacobbe, P.; et al. Integrated Genome-Wide Methylation and Expression Analyses Reveal Functional Predictors of Response to Antidepressants. Transl. Psychiatry 2019, 9, 254. [Google Scholar] [CrossRef]
  185. Lisoway, A.J.; Chen, C.C.; Zai, C.C.; Tiwari, A.K.; Kennedy, J.L. Toward Personalized Medicine in Schizophrenia: Genetics and Epigenetics of Antipsychotic Treatment. Schizophr. Res. 2021, 232, 112–124. [Google Scholar] [CrossRef]
  186. McKenna, B.G.; Hendrix, C.L.; Brennan, P.A.; Smith, A.K.; Stowe, Z.N.; Newport, D.J.; Knight, A.K. Maternal Prenatal Depression and Epigenetic Age Deceleration: Testing Potentially Confounding Effects of Prenatal Stress and SSRI Use. Epigenetics 2021, 16, 327–337. [Google Scholar] [CrossRef]
  187. Oh, G.; Wang, S.-C.; Pal, M.; Chen, Z.F.; Khare, T.; Tochigi, M.; Ng, C.; Yang, Y.A.; Kwan, A.; Kaminsky, Z.A.; et al. DNA Modification Study of Major Depressive Disorder: Beyond Locus-by-Locus Comparisons. Biol. Psychiatry 2015, 77, 246–255. [Google Scholar] [CrossRef] [PubMed]
  188. Robison, A.J.; Nestler, E.J. Transcriptional and Epigenetic Mechanisms of Addiction. Nat. Rev. Neurosci. 2011, 12, 623–637. [Google Scholar] [CrossRef] [PubMed]
  189. Saad, L.; Zwiller, J.; Kalsbeek, A.; Anglard, P. Epigenetic Regulation of Circadian Clocks and Its Involvement in Drug Addiction. Genes 2021, 12, 1263. [Google Scholar] [CrossRef]
  190. Werner, C.T.; Altshuler, R.D.; Shaham, Y.; Li, X. Epigenetic Mechanisms in Drug Relapse. Biol. Psychiatry 2021, 89, 331–338. [Google Scholar] [CrossRef] [PubMed]
  191. Doke, M.; Pendyala, G.; Samikkannu, T. Psychostimulants and Opioids Differentially Influence the Epigenetic Modification of Histone Acetyltransferase and Histone Deacetylase in Astrocytes. PLoS ONE 2021, 16, e0252895. [Google Scholar] [CrossRef] [PubMed]
  192. Arndt, D.L.; Wukitsch, T.J.; Garcia, E.J.; Cain, M. Histone Deacetylase Inhibition Differentially Attenuates Cue-Induced Reinstatement: An Interaction of Environment and AcH3K9 Expression in the Dorsal Striatum. Behav. Neurosci. 2019, 133, 478–488. [Google Scholar] [CrossRef] [PubMed]
  193. Chen, W.-S.; Xu, W.-J.; Zhu, H.-Q.; Gao, L.; Lai, M.-J.; Zhang, F.-Q.; Zhou, W.-H.; Liu, H.-F. Effects of Histone Deacetylase Inhibitor Sodium Butyrate on Heroin Seeking Behavior in the Nucleus Accumbens in Rats. Brain Res. 2016, 1652, 151–157. [Google Scholar] [CrossRef] [PubMed]
  194. Sartor, G.C.; Powell, S.K.; Brothers, S.P.; Wahlestedt, C. Epigenetic Readers of Lysine Acetylation Regulate Cocaine-Induced Plasticity. J. Neurosci. 2015, 35, 15062–15072. [Google Scholar] [CrossRef] [PubMed]
  195. Sharma, C.; Oh, Y.J.; Park, B.; Lee, S.; Jeong, C.-H.; Lee, S.; Seo, J.H.; Seo, Y.H. Development of Thiazolidinedione-Based HDAC6 Inhibitors to Overcome Methamphetamine Addiction. Int. J. Mol. Sci. 2019, 20, 6213. [Google Scholar] [CrossRef]
  196. Simon-O’Brien, E.; Alaux-Cantin, S.; Warnault, V.; Buttolo, R.; Naassila, M.; Vilpoux, C. The Histone Deacetylase Inhibitor Sodium Butyrate Decreases Excessive Ethanol Intake in Dependent Animals. Addict. Biol. 2015, 20, 676–689. [Google Scholar] [CrossRef]
  197. Zhang, X.; Sun, L.; Wang, L.; Wang, M.; Lu, G.; Wang, Y.; Li, Q.; Li, C.; Zhou, J.; Ma, H.; et al. The Effects of Histone Deacetylase Inhibitors on the Attentional Set-Shifting Task Performance of Alcohol-Dependent Rats. Brain Res. Bull. 2019, 149, 208–215. [Google Scholar] [CrossRef] [PubMed]
  198. Kong, F.-C.; Ma, C.-L.; Zhong, M.-K. Epigenetic Effects Mediated by Antiepileptic Drugs and Their Potential Application. Curr. Neuropharmacol. 2020, 18, 153–166. [Google Scholar] [CrossRef] [PubMed]
  199. Koob, G.F.; Volkow, N.D. Neurobiology of Addiction: A Neurocircuitry Analysis. Lancet Psychiatry 2016, 3, 760–773. [Google Scholar] [CrossRef] [PubMed]
  200. Hamilton, P.J.; Nestler, E.J. Epigenetics and Addiction. Curr. Opin. Neurobiol. 2019, 59, 128–136. [Google Scholar] [CrossRef]
  201. Younus, I.; Reddy, D.S. Epigenetic Interventions for Epileptogenesis: A New Frontier for Curing Epilepsy. Pharmacol. Ther. 2017, 177, 108–122. [Google Scholar] [CrossRef]
  202. Tao, H.; Chen, Z.; Wu, J.; Chen, J.; Chen, Y.; Fu, J.; Sun, C.; Zhou, H.; Zhong, W.; Zhou, X.; et al. DNA Methylation Signature of Epileptic Encephalopathy-Related Pathogenic Genes Encoding Ion Channels in Temporal Lobe Epilepsy. Front. Neurol. 2021, 12, 692412. [Google Scholar] [CrossRef]
  203. Liu, K.; Wu, H.; Gao, R.; Han, G. DNA Methylation May Be Involved in the Analgesic Effect of Hyperbaric Oxygen via Regulating FUNDC1. Pain Res. Manag. 2020, 2020, 1528362. [Google Scholar] [CrossRef]
  204. He, X.-T.; Hu, X.-F.; Zhu, C.; Zhou, K.-X.; Zhao, W.-J.; Zhang, C.; Han, X.; Wu, C.-L.; Wei, Y.-Y.; Wang, W.; et al. Suppression of Histone Deacetylases by SAHA Relieves Bone Cancer Pain in Rats via Inhibiting Activation of Glial Cells in Spinal Dorsal Horn and Dorsal Root Ganglia. J. Neuroinflammation 2020, 17, 125. [Google Scholar] [CrossRef]
  205. Matsushita, Y.; Araki, K.; Omotuyi, O.i.; Mukae, T.; Ueda, H. HDAC Inhibitors Restore C-Fibre Sensitivity in Experimental Neuropathic Pain Model. Br. J. Pharmacol. 2013, 170, 991–998. [Google Scholar] [CrossRef]
  206. Uchida, H.; Matsushita, Y.; Araki, K.; Mukae, T.; Ueda, H. Histone Deacetylase Inhibitors Relieve Morphine Resistance in Neuropathic Pain after Peripheral Nerve Injury. J. Pharmacol. Sci. 2015, 128, 208–211. [Google Scholar] [CrossRef]
  207. Alegría-Torres, J.A.; Baccarelli, A.; Bollati, V. Epigenetics and Lifestyle. Epigenomics 2011, 3, 267–277. [Google Scholar] [CrossRef] [PubMed]
  208. Klemp, I.; Hoffmann, A.; Müller, L.; Hagemann, T.; Horn, K.; Rohde-Zimmermann, K.; Tönjes, A.; Thiery, J.; Löffler, M.; Burkhardt, R.; et al. DNA Methylation Patterns Reflect Individual’s Lifestyle Independent of Obesity. Clin. Transl. Med. 2022, 12, e851. [Google Scholar] [CrossRef]
  209. Hashioka, S.; Inoue, K.; Hayashida, M.; Wake, R.; Oh-Nishi, A.; Miyaoka, T. Implications of Systemic Inflammation and Periodontitis for Major Depression. Front. Neurosci. 2018, 12, 483. [Google Scholar] [CrossRef] [PubMed]
  210. de Haan, P.; Klein, H.C.; ’t Hart, B.A. Autoimmune Aspects of Neurodegenerative and Psychiatric Diseases: A Template for Innovative Therapy. Front. Psychiatry 2017, 8, 46. [Google Scholar] [CrossRef] [PubMed]
  211. Cunningham, C.; Campion, S.; Lunnon, K.; Murray, C.L.; Woods, J.F.C.; Deacon, R.M.J.; Rawlins, J.N.P.; Perry, V.H. Systemic Inflammation Induces Acute Behavioral and Cognitive Changes and Accelerates Neurodegenerative Disease. Biol. Psychiatry 2009, 65, 304–312. [Google Scholar] [CrossRef] [PubMed]
  212. Ellulu, M.S.; Patimah, I.; Khaza’ai, H.; Rahmat, A.; Abed, Y. Obesity and Inflammation: The Linking Mechanism and the Complications. Arch. Med. Sci. 2017, 13, 851–863. [Google Scholar] [CrossRef] [PubMed]
  213. Guillemot-Legris, O.; Muccioli, G.G. Obesity-Induced Neuroinflammation: Beyond the Hypothalamus. Trends Neurosci. 2017, 40, 237–253. [Google Scholar] [CrossRef]
  214. Milaneschi, Y.; Simmons, W.K.; van Rossum, E.F.C.; Penninx, B.W. Depression and Obesity: Evidence of Shared Biological Mechanisms. Mol. Psychiatry 2019, 24, 18–33. [Google Scholar] [CrossRef]
  215. Fasshauer, M.; Blüher, M. Adipokines in Health and Disease. Trends Pharmacol. Sci. 2015, 36, 461–470. [Google Scholar] [CrossRef]
  216. Ouchi, N.; Parker, J.L.; Lugus, J.J.; Walsh, K. Adipokines in Inflammation and Metabolic Disease. Nat. Rev. Immunol. 2011, 11, 85–97. [Google Scholar] [CrossRef]
  217. Chen, Y.; Xia, K.; Chen, L.; Fan, D. Increased Interleukin-6 Levels in the Astrocyte-Derived Exosomes of Sporadic Amyotrophic Lateral Sclerosis Patients. Front. Neurosci. 2019, 13, 574. [Google Scholar] [CrossRef] [PubMed]
  218. Jung, Y.J.; Tweedie, D.; Scerba, M.T.; Greig, N.H. Neuroinflammation as a Factor of Neurodegenerative Disease: Thalidomide Analogs as Treatments. Front. Cell Dev. Biol. 2019, 7, 313. [Google Scholar] [CrossRef] [PubMed]
  219. Lin, C.; Chen, K.; Yu, J.; Feng, W.; Fu, W.; Yang, F.; Zhang, X.; Chen, D. Relationship between TNF-α Levels and Psychiatric Symptoms in First-Episode Drug-Naïve Patients with Schizophrenia before and after Risperidone Treatment and in Chronic Patients. BMC Psychiatry 2021, 21, 561. [Google Scholar] [CrossRef]
  220. Silva, N.M.L.E.; Gonçalves, R.A.; Pascoal, T.A.; Lima-Filho, R.A.S.; de Paula França Resende, E.; Vieira, E.L.M.; Teixeira, A.L.; de Souza, L.C.; Peny, J.A.; Fortuna, J.T.S.; et al. Pro-Inflammatory Interleukin-6 Signaling Links Cognitive Impairments and Peripheral Metabolic Alterations in Alzheimer’s Disease. Transl. Psychiatry 2021, 11, 251. [Google Scholar] [CrossRef] [PubMed]
  221. Ting, E.Y.-C.; Yang, A.C.; Tsai, S.-J. Role of Interleukin-6 in Depressive Disorder. Int. J. Mol. Sci. 2020, 21, 2194. [Google Scholar] [CrossRef] [PubMed]
  222. Wang, A.K.; Miller, B.J. Meta-Analysis of Cerebrospinal Fluid Cytokine and Tryptophan Catabolite Alterations in Psychiatric Patients: Comparisons between Schizophrenia, Bipolar Disorder, and Depression. Schizophr. Bull. 2018, 44, 75–83. [Google Scholar] [CrossRef]
  223. Yao, L.; Pan, L.; Qian, M.; Sun, W.; Gu, C.; Chen, L.; Tang, X.; Hu, Y.; Xu, L.; Wei, Y.; et al. Tumor Necrosis Factor-α Variations in Patients with Major Depressive Disorder before and after Antidepressant Treatment. Front. Psychiatry 2020, 11, 518837. [Google Scholar] [CrossRef]
  224. Ladd-Acosta, C.; Pevsner, J.; Sabunciyan, S.; Yolken, R.H.; Webster, M.J.; Dinkins, T.; Callinan, P.A.; Fan, J.-B.; Potash, J.B.; Feinberg, A.P. DNA Methylation Signatures within the Human Brain. Am. J. Hum. Genet. 2007, 81, 1304–1315. [Google Scholar] [CrossRef]
  225. Ligthart, S.; Marzi, C.; Aslibekyan, S.; Mendelson, M.M.; Conneely, K.N.; Tanaka, T.; Colicino, E.; Waite, L.L.; Joehanes, R.; Guan, W.; et al. DNA Methylation Signatures of Chronic Low-Grade Inflammation Are Associated with Complex Diseases. Genome Biol. 2016, 17, 255. [Google Scholar] [CrossRef]
  226. Gonzalez-Jaramillo, V.; Portilla-Fernandez, E.; Glisic, M.; Voortman, T.; Ghanbari, M.; Bramer, W.; Chowdhury, R.; Nijsten, T.; Dehghan, A.; Franco, O.H.; et al. Epigenetics and Inflammatory Markers: A Systematic Review of the Current Evidence. Int. J. Inflamm. 2019, 2019, e6273680. [Google Scholar] [CrossRef]
  227. Christ, A.; Günther, P.; Lauterbach, M.A.R.; Duewell, P.; Biswas, D.; Pelka, K.; Scholz, C.J.; Oosting, M.; Haendler, K.; Baßler, K.; et al. Western Diet Triggers NLRP3-Dependent Innate Immune Reprogramming. Cell 2018, 172, 162–175.e14. [Google Scholar] [CrossRef] [PubMed]
  228. Manna, P.; Jain, S.K. Obesity, Oxidative Stress, Adipose Tissue Dysfunction, and the Associated Health Risks: Causes and Therapeutic Strategies. Metab. Syndr. Relat. Disord. 2015, 13, 423–444. [Google Scholar] [CrossRef] [PubMed]
  229. Kietzmann, T.; Petry, A.; Shvetsova, A.; Gerhold, J.M.; Görlach, A. The Epigenetic Landscape Related to Reactive Oxygen Species Formation in the Cardiovascular System. Br. J. Pharmacol. 2017, 174, 1533–1554. [Google Scholar] [CrossRef] [PubMed]
  230. Arora, I.; Sharma, M.; Sun, L.Y.; Tollefsbol, T.O. The Epigenetic Link between Polyphenols, Aging and Age-Related Diseases. Genes 2020, 11, 1094. [Google Scholar] [CrossRef] [PubMed]
  231. Melnik, B.C.; Stremmel, W.; Weiskirchen, R.; John, S.M.; Schmitz, G. Exosome-Derived MicroRNAs of Human Milk and Their Effects on Infant Health and Development. Biomolecules 2021, 11, 851. [Google Scholar] [CrossRef]
  232. Cannataro, R.; Cione, E. Diet and MiRNA: Epigenetic Regulator or a New Class of Supplements? Microrna 2022, 11, 89–90. [Google Scholar] [CrossRef]
  233. Donohoe, D.R.; Bultman, S.J. Metaboloepigenetics: Interrelationships between Energy Metabolism and Epigenetic Control of Gene Expression. J. Cell. Physiol. 2012, 227, 3169–3177. [Google Scholar] [CrossRef]
  234. Clapier, C.R.; Iwasa, J.; Cairns, B.R.; Peterson, C.L. Mechanisms of Action and Regulation of ATP-Dependent Chromatin-Remodelling Complexes. Nat. Rev. Mol. Cell. Biol. 2017, 18, 407–422. [Google Scholar] [CrossRef]
  235. Bradshaw, P.C. Acetyl-CoA Metabolism and Histone Acetylation in the Regulation of Aging and Lifespan. Antioxidants 2021, 10, 572. [Google Scholar] [CrossRef]
  236. Schwer, B.; North, B.J.; Frye, R.A.; Ott, M.; Verdin, E. The Human Silent Information Regulator (Sir)2 Homologue HSIRT3 Is a Mitochondrial Nicotinamide Adenine Dinucleotide-Dependent Deacetylase. J. Cell. Biol. 2002, 158, 647–657. [Google Scholar] [CrossRef]
  237. López-Otín, C.; Blasco, M.A.; Partridge, L.; Serrano, M.; Kroemer, G. The Hallmarks of Aging. Cell 2013, 153, 1194–1217. [Google Scholar] [CrossRef] [PubMed]
  238. Alirezaei, M.; Kemball, C.C.; Flynn, C.T.; Wood, M.R.; Whitton, J.L.; Kiosses, W.B. Short-Term Fasting Induces Profound Neuronal Autophagy. Autophagy 2010, 6, 702–710. [Google Scholar] [CrossRef] [PubMed]
  239. Liu, Y.; Wang, R.; Zhao, Z.; Dong, W.; Zhang, X.; Chen, X.; Ma, L. Short-Term Caloric Restriction Exerts Neuroprotective Effects Following Mild Traumatic Brain Injury by Promoting Autophagy and Inhibiting Astrocyte Activation. Behav. Brain Res. 2017, 331, 135–142. [Google Scholar] [CrossRef] [PubMed]
  240. Loos, B.; Klionsky, D.J.; Wong, E. Augmenting Brain Metabolism to Increase Macro- and Chaperone-Mediated Autophagy for Decreasing Neuronal Proteotoxicity and Aging. Prog. Neurobiol. 2017, 156, 90–106. [Google Scholar] [CrossRef]
  241. Cunha-Santos, J.; Duarte-Neves, J.; Carmona, V.; Guarente, L.; Pereira de Almeida, L.; Cavadas, C. Caloric Restriction Blocks Neuropathology and Motor Deficits in Machado-Joseph Disease Mouse Models through SIRT1 Pathway. Nat. Commun. 2016, 7, 11445. [Google Scholar] [CrossRef] [PubMed]
  242. Green, C.L.; Lamming, D.W.; Fontana, L. Molecular Mechanisms of Dietary Restriction Promoting Health and Longevity. Nat. Rev. Mol. Cell. Biol. 2022, 23, 56–73. [Google Scholar] [CrossRef] [PubMed]
  243. Spanaki, C.; Rodopaios, N.E.; Koulouri, A.; Pliakas, T.; Papadopoulou, S.K.; Vasara, E.; Skepastianos, P.; Serafeim, T.; Boura, I.; Dermitzakis, E.; et al. The Christian Orthodox Church Fasting Diet Is Associated with Lower Levels of Depression and Anxiety and a Better Cognitive Performance in Middle Life. Nutrients 2021, 13, 627. [Google Scholar] [CrossRef] [PubMed]
  244. Thapar, A.; Riglin, L. The Importance of a Developmental Perspective in Psychiatry: What Do Recent Genetic-Epidemiological Findings Show? Mol. Psychiatry 2020, 25, 1631–1639. [Google Scholar] [CrossRef] [PubMed]
  245. Heijmans, B.T.; Tobi, E.W.; Stein, A.D.; Putter, H.; Blauw, G.J.; Susser, E.S.; Slagboom, P.E.; Lumey, L.H. Persistent Epigenetic Differences Associated with Prenatal Exposure to Famine in Humans. Proc. Natl. Acad. Sci. USA 2008, 105, 17046–17049. [Google Scholar] [CrossRef] [PubMed]
  246. Kereliuk, S.M.; Brawerman, G.M.; Dolinsky, V.W. Maternal Macronutrient Consumption and the Developmental Origins of Metabolic Disease in the Offspring. Int. J. Mol. Sci. 2017, 18, 1451. [Google Scholar] [CrossRef]
  247. Bygren, L.O. Intergenerational Health Responses to Adverse and Enriched Environments. Annu. Rev. Public Health 2013, 34, 49–60. [Google Scholar] [CrossRef]
  248. Susser, E.S.; Lin, S.P. Schizophrenia after Prenatal Exposure to the Dutch Hunger Winter of 1944–1945. Arch. Gen. Psychiatry 1992, 49, 983–988. [Google Scholar] [CrossRef] [PubMed]
  249. Gawliński, D.; Gawlińska, K.; Smaga, I. Maternal High-Fat Diet Modulates Cnr1 Gene Expression in Male Rat Offspring. Nutrients 2021, 13, 2885. [Google Scholar] [CrossRef] [PubMed]
  250. Glendining, K.A.; Jasoni, C.L. Maternal High Fat Diet-Induced Obesity Modifies Histone Binding and Expression of Oxtr in Offspring Hippocampus in a Sex-Specific Manner. Int. J. Mol. Sci. 2019, 20, 329. [Google Scholar] [CrossRef] [PubMed]
  251. Urbonaite, G.; Knyzeliene, A.; Bunn, F.S.; Smalskys, A.; Neniskyte, U. The Impact of Maternal High-Fat Diet on Offspring Neurodevelopment. Front. Neurosci. 2022, 16, 909762. [Google Scholar] [CrossRef] [PubMed]
  252. Burdge, G.C.; Lillycrop, K.A. Fatty Acids and Epigenetics. Curr. Opin. Clin. Nutr. Metab. Care 2014, 17, 156–161. [Google Scholar] [CrossRef]
  253. Ershad Sarabi, S.; Han, Q.; Romme, A.G.L.; de Vries, B.; Wendling, L. Wendling Key Enablers of and Barriers to the Uptake and Implementation of Nature-Based Solutions in Urban Settings: A Review. Resources 2019, 8, 121. [Google Scholar] [CrossRef]
  254. Cinquina, V.; Calvigioni, D.; Farlik, M.; Halbritter, F.; Fife-Gernedl, V.; Shirran, S.L.; Fuszard, M.A.; Botting, C.H.; Poullet, P.; Piscitelli, F.; et al. Life-Long Epigenetic Programming of Cortical Architecture by Maternal ‘Western’ Diet during Pregnancy. Mol. Psychiatry 2020, 25, 22–36. [Google Scholar] [CrossRef]
  255. Ishii, T.; Shimpo, Y.; Matsuoka, Y.; Kinoshita, K. Anti-Apoptotic Effect of Acetyl-l-Carnitine and I-Carnitine in Primary Cultured Neurons. Jpn. J. Pharmacol. 2000, 83, 119–124. [Google Scholar] [CrossRef]
  256. Nasca, C.; Xenos, D.; Barone, Y.; Caruso, A.; Scaccianoce, S.; Matrisciano, F.; Battaglia, G.; Mathé, A.A.; Pittaluga, A.; Lionetto, L.; et al. L-Acetylcarnitine Causes Rapid Antidepressant Effects through the Epigenetic Induction of MGlu2 Receptors. Proc. Natl. Acad. Sci. USA 2013, 110, 4804–4809. [Google Scholar] [CrossRef] [PubMed]
  257. Schaevitz, L.R.; Nicolai, R.; Lopez, C.M.; D’Iddio, S.; Iannoni, E.; Berger-Sweeney, J.E. Acetyl-L-Carnitine Improves Behavior and Dendritic Morphology in a Mouse Model of Rett Syndrome. PLoS ONE 2012, 7, e51586. [Google Scholar] [CrossRef] [PubMed]
  258. Sakamoto, A.; Terui, Y.; Uemura, T.; Igarashi, K.; Kashiwagi, K. Polyamines Regulate Gene Expression by Stimulating Translation of Histone Acetyltransferase MRNAs. J. Biol. Chem. 2020, 295, 8736–8745. [Google Scholar] [CrossRef] [PubMed]
  259. Abdul, Q.A.; Yu, B.P.; Chung, H.Y.; Jung, H.A.; Choi, J.S. Epigenetic Modifications of Gene Expression by Lifestyle and Environment. Arch. Pharm. Res. 2017, 40, 1219–1237. [Google Scholar] [CrossRef]
  260. Fini, L.; Selgrad, M.; Fogliano, V.; Graziani, G.; Romano, M.; Hotchkiss, E.; Daoud, Y.A.; De Vol, E.B.; Boland, C.R.; Ricciardiello, L. Annurca Apple Polyphenols Have Potent Demethylating Activity and Can Reactivate Silenced Tumor Suppressor Genes in Colorectal Cancer Cells. J. Nutr. 2007, 137, 2622–2628. [Google Scholar] [CrossRef] [PubMed]
  261. Link, A.; Balaguer, F.; Goel, A. Cancer Chemoprevention by Dietary Polyphenols: Promising Role for Epigenetics. Biochem. Pharmacol. 2010, 80, 1771–1792. [Google Scholar] [CrossRef] [PubMed]
  262. Russo, G.L.; Vastolo, V.; Ciccarelli, M.; Albano, L.; Macchia, P.E.; Ungaro, P. Dietary Polyphenols and Chromatin Remodeling. Crit. Rev. Food Sci. Nutr. 2017, 57, 2589–2599. [Google Scholar] [CrossRef]
  263. Morris, G.; Gamage, E.; Travica, N.; Berk, M.; Jacka, F.N.; O’Neil, A.; Puri, B.K.; Carvalho, A.F.; Bortolasci, C.C.; Walder, K.; et al. Polyphenols as Adjunctive Treatments in Psychiatric and Neurodegenerative Disorders: Efficacy, Mechanisms of Action, and Factors Influencing Inter-Individual Response. Free. Radic. Biol. Med. 2021, 172, 101–122. [Google Scholar] [CrossRef]
  264. Ntie-Kang, F.; Karaman Mayack, B.; Valente, S.; Battistelli, C. Editorial: Natural Product Epigenetic Modulators and Inhibitors. Front. Pharmacol. 2021, 12, 651395. [Google Scholar] [CrossRef]
  265. Soares Romeiro, L.A.; da Costa Nunes, J.L.; de Oliveira Miranda, C.; Simões Heyn Roth Cardoso, G.; de Oliveira, A.S.; Gandini, A.; Kobrlova, T.; Soukup, O.; Rossi, M.; Senger, J.; et al. Novel Sustainable-by-Design HDAC Inhibitors for the Treatment of Alzheimer’s Disease. ACS Med. Chem. Lett. 2019, 10, 671–676. [Google Scholar] [CrossRef]
  266. Kitahara, M.; Inoue, T.; Mani, H.; Takamatsu, Y.; Ikegami, R.; Tohyama, H.; Maejima, H. Exercise and Pharmacological Inhibition of Histone Deacetylase Improves Cognitive Function Accompanied by an Increase of Gene Expressions Crucial for Neuronal Plasticity in the Hippocampus. Neurosci. Lett. 2021, 749, 135749. [Google Scholar] [CrossRef]
  267. Liang, J.; Wang, H.; Zeng, Y.; Qu, Y.; Liu, Q.; Zhao, F.; Duan, J.; Jiang, Y.; Li, S.; Ying, J.; et al. Physical Exercise Promotes Brain Remodeling by Regulating Epigenetics, Neuroplasticity and Neurotrophins. Rev. Neurosci. 2021, 32, 615–629. [Google Scholar] [CrossRef] [PubMed]
  268. Maejima, H.; Kanemura, N.; Kokubun, T.; Murata, K.; Takayanagi, K. Exercise Enhances Cognitive Function and Neurotrophin Expression in the Hippocampus Accompanied by Changes in Epigenetic Programming in Senescence-Accelerated Mice. Neurosci. Lett. 2018, 665, 67–73. [Google Scholar] [CrossRef]
  269. Sellami, M.; Bragazzi, N.; Prince, M.S.; Denham, J.; Elrayess, M. Regular, Intense Exercise Training as a Healthy Aging Lifestyle Strategy: Preventing DNA Damage, Telomere Shortening and Adverse DNA Methylation Changes Over a Lifetime. Front. Genet. 2021, 12, 652497. [Google Scholar] [CrossRef] [PubMed]
  270. Spindler, C.; Segabinazi, E.; de Meireles, A.L.F.; Piazza, F.V.; Mega, F.; Dos Santos Salvalaggio, G.; Achaval, M.; Elsner, V.R.; Marcuzzo, S. Paternal Physical Exercise Modulates Global DNA Methylation Status in the Hippocampus of Male Rat Offspring. Neural Regen. Res. 2019, 14, 491–500. [Google Scholar] [CrossRef] [PubMed]
  271. Xu, M.; Zhu, J.; Liu, X.-D.; Luo, M.-Y.; Xu, N.-J. Roles of Physical Exercise in Neurodegeneration: Reversal of Epigenetic Clock. Transl. Neurodegener. 2021, 10, 30. [Google Scholar] [CrossRef] [PubMed]
  272. Dougherty, R.J.; Jonaitis, E.M.; Gaitán, J.M.; Lose, S.R.; Mergen, B.M.; Johnson, S.C.; Okonkwo, O.C.; Cook, D.B. Cardiorespiratory Fitness Mitigates Brain Atrophy and Cognitive Decline in Adults at Risk for Alzheimer’s Disease. Alzheimers Dement. 2021, 13, e12212. [Google Scholar] [CrossRef]
  273. Won, J.; Callow, D.D.; Pena, G.S.; Jordan, L.S.; Arnold-Nedimala, N.A.; Nielson, K.A.; Smith, J.C. Hippocampal Functional Connectivity and Memory Performance after Exercise Intervention in Older Adults with Mild Cognitive Impairment. J. Alzheimers Dis. 2021, 82, 1015–1031. [Google Scholar] [CrossRef]
  274. Herbsleb, M.; Schumann, A.; Lehmann, L.; Gabriel, H.H.W.; Bär, K.-J. Cardio-Respiratory Fitness and Autonomic Function in Patients with Major Depressive Disorder. Front. Psychiatry 2019, 10, 980. [Google Scholar] [CrossRef]
  275. Ito, S.; D’Alessio, A.C.; Taranova, O.V.; Hong, K.; Sowers, L.C.; Zhang, Y. Role of Tet Proteins in 5mC to 5hmC Conversion, ES-Cell Self-Renewal and Inner Cell Mass Specification. Nature 2010, 466, 1129–1133. [Google Scholar] [CrossRef]
  276. Scourzic, L.; Mouly, E.; Bernard, O.A. TET Proteins and the Control of Cytosine Demethylation in Cancer. Genome Med. 2015, 7, 9. [Google Scholar] [CrossRef]
  277. MacArthur, I.C.; Dawlaty, M.M. TET Enzymes and 5-Hydroxymethylcytosine in Neural Progenitor Cell Biology and Neurodevelopment. Front. Cell. Dev. Biol. 2021, 9, 645335. [Google Scholar] [CrossRef] [PubMed]
  278. Kriaucionis, S.; Heintz, N. The Nuclear DNA Base 5-Hydroxymethylcytosine Is Present in Purkinje Neurons and the Brain. Science 2009, 324, 929–930. [Google Scholar] [CrossRef]
  279. Jakubek, M.; Kejík, Z.; Kaplánek, R.; Antonyová, V.; Hromádka, R.; Šandriková, V.; Sýkora, D.; Martásek, P.; Král, V. Hydrazones as Novel Epigenetic Modulators: Correlation between TET 1 Protein Inhibition Activity and Their Iron(II) Binding Ability. Bioorganic Chem. 2019, 88, 102809. [Google Scholar] [CrossRef] [PubMed]
  280. Singh, A.K.; Zhao, B.; Liu, X.; Wang, X.; Li, H.; Qin, H.; Wu, X.; Ma, Y.; Horne, D.; Yu, X. Selective Targeting of TET Catalytic Domain Promotes Somatic Cell Reprogramming. Proc. Natl. Acad. Sci. USA 2020, 117, 3621–3626. [Google Scholar] [CrossRef]
  281. Winkle, M.; El-Daly, S.M.; Fabbri, M.; Calin, G.A. Noncoding RNA Therapeutics—Challenges and Potential Solutions. Nat. Rev. Drug. Discov. 2021, 20, 629–651. [Google Scholar] [CrossRef]
  282. Hanna, J.; Hossain, G.S.; Kocerha, J. The Potential for MicroRNA Therapeutics and Clinical Research. Front. Genet. 2019, 10, 478. [Google Scholar] [CrossRef] [PubMed]
  283. Bian, Q.; Chen, J.; Wu, J.; Ding, F.; Li, X.; Ma, Q.; Zhang, L.; Zou, X.; Chen, J. Bioinformatics Analysis of a TF-MiRNA-LncRNA Regulatory Network in Major Depressive Disorder. Psychiatry Res. 2021, 299, 113842. [Google Scholar] [CrossRef]
  284. Bavamian, S.; Mellios, N.; Lalonde, J.; Fass, D.M.; Wang, J.; Sheridan, S.D.; Madison, J.M.; Zhou, F.; Rueckert, E.H.; Barker, D.; et al. Dysregulation of MiR-34a Links Neuronal Development to Genetic Risk Factors for Bipolar Disorder. Mol. Psychiatry 2015, 20, 573–584. [Google Scholar] [CrossRef]
  285. Dwivedi, Y. Evidence Demonstrating Role of MicroRNAs in the Etiopathology of Major Depression. J. Chem. Neuroanat. 2011, 42, 142–156. [Google Scholar] [CrossRef]
  286. Huaying, C.; Xing, J.; Luya, J.; Linhui, N.; Di, S.; Xianjun, D. A Signature of Five Long Non-Coding RNAs for Predicting the Prognosis of Alzheimer’s Disease Based on Competing Endogenous RNA Networks. Front. Aging Neurosci. 2020, 12, 598606. [Google Scholar] [CrossRef]
  287. Maffioletti, E.; Cattaneo, A.; Rosso, G.; Maina, G.; Maj, C.; Gennarelli, M.; Tardito, D.; Bocchio-Chiavetto, L. Peripheral Whole Blood MicroRNA Alterations in Major Depression and Bipolar Disorder. J. Affect. Disord. 2016, 200, 250–258. [Google Scholar] [CrossRef] [PubMed]
  288. Song, J.; Kim, Y.-K. Identification of the Role of MiR-142-5p in Alzheimer’s Disease by Comparative Bioinformatics and Cellular Analysis. Front. Mol. Neurosci. 2017, 10, 227. [Google Scholar] [CrossRef]
  289. Zhou, L.; Zhu, Y.; Chen, W.; Tang, Y. Emerging Role of MicroRNAs in Major Depressive Disorder and Its Implication on Diagnosis and Therapeutic Response. J. Affect. Disord. 2021, 286, 80–86. [Google Scholar] [CrossRef]
  290. Ling, H. Non-Coding RNAs: Therapeutic Strategies and Delivery Systems. Adv. Exp. Med. Biol. 2016, 937, 229–237. [Google Scholar] [CrossRef] [PubMed]
  291. MacDiarmid, J.A.; Brahmbhatt, H. Minicells: Versatile Vectors for Targeted Drug or Si/ShRNA Cancer Therapy. Curr. Opin. Biotechnol. 2011, 22, 909–916. [Google Scholar] [CrossRef] [PubMed]
  292. HIN Clinical Trials Database. Available online: https://clinicaltrials.gov/ct2/home (accessed on 31 December 2022).
  293. EHSAN Effect of Hydralazine on Alzheimer’s Disease (EHSAN). Available online: https://clinicaltrials.gov/ct2/show/NCT04842552 (accessed on 31 December 2022).
  294. Baker, G.B.; Coutts, R.T.; McKenna, K.F.; Sherry-McKenna, R.L. Insights into the Mechanisms of Action of the MAO Inhibitors Phenelzine and Tranylcypromine: A Review. J. Psychiatry Neurosci. 1992, 17, 206–214. [Google Scholar]
  295. Giallongo, S.; Longhitano, L.; Denaro, S.; D’Aprile, S.; Torrisi, F.; La Spina, E.; Giallongo, C.; Mannino, G.; Lo Furno, D.; Zappalà, A.; et al. The Role of Epigenetics in Neuroinflammatory-Driven Diseases. Int. J. Mol. Sci. 2022, 23, 15218. [Google Scholar] [CrossRef]
  296. Ortega, M.A.; Fraile-Martínez, Ó.; García-Montero, C.; Alvarez-Mon, M.A.; Lahera, G.; Monserrat, J.; Llavero-Valero, M.; Mora, F.; Rodríguez-Jiménez, R.; Fernandez-Rojo, S.; et al. Nutrition, Epigenetics, and Major Depressive Disorder: Understanding the Connection. Front. Nutr. 2022, 9, 867150. [Google Scholar] [CrossRef]
  297. Rasheed, M.; Liang, J.; Wang, C.; Deng, Y.; Chen, Z. Epigenetic Regulation of Neuroinflammation in Parkinson’s Disease. Int. J. Mol. Sci. 2021, 22, 4956. [Google Scholar] [CrossRef]
Figure 1. Chromatin structure. In eukaryotes, DNA (deoxyribonucleic acid) molecules in the cell nucleus are packed into chromosomes; the integral substance of which is called chromatin. Back in 1928, Emil Heitz noticed that some chromatin regions (chromosome territories) are more compact than others and described the chromosomal substance as unfolded eu- and compact heterochromatin [1]. Loosely coiled chromatin contains transcriptionally accessible DNA regions, whereas tightly coiled chromatin comprises transcriptionally inactive DNA regions [2]. One long DNA molecule is wound around globules of histone proteins (like beads on a string) and folded into a compact structure along with other proteins (architectural) and RNA (ribonucleic acid) molecules. Negatively charged 147 base pairs piece of the DNA molecule, which is wrapped around a positively charged octamer of a histone protein core [3], called a nucleosome. Nucleosomes are arranged regularly: each of the two molecules of histones 2A (H2A), 2B (H2B), 3 (H3), and 4 (H4) is folded into a compact structure in such a way that a tetramer from histones H3 and H4 is “sandwiched” between two dimers of H2A and H2B. Such packaging allows the DNA molecule, 2 m long, to be packed into a nucleus of several micrometers in size, and to be effectively subjected to the complex processes of gene reading and expression. Created with BioRender.com (accessed on 2 August 2022).
Figure 1. Chromatin structure. In eukaryotes, DNA (deoxyribonucleic acid) molecules in the cell nucleus are packed into chromosomes; the integral substance of which is called chromatin. Back in 1928, Emil Heitz noticed that some chromatin regions (chromosome territories) are more compact than others and described the chromosomal substance as unfolded eu- and compact heterochromatin [1]. Loosely coiled chromatin contains transcriptionally accessible DNA regions, whereas tightly coiled chromatin comprises transcriptionally inactive DNA regions [2]. One long DNA molecule is wound around globules of histone proteins (like beads on a string) and folded into a compact structure along with other proteins (architectural) and RNA (ribonucleic acid) molecules. Negatively charged 147 base pairs piece of the DNA molecule, which is wrapped around a positively charged octamer of a histone protein core [3], called a nucleosome. Nucleosomes are arranged regularly: each of the two molecules of histones 2A (H2A), 2B (H2B), 3 (H3), and 4 (H4) is folded into a compact structure in such a way that a tetramer from histones H3 and H4 is “sandwiched” between two dimers of H2A and H2B. Such packaging allows the DNA molecule, 2 m long, to be packed into a nucleus of several micrometers in size, and to be effectively subjected to the complex processes of gene reading and expression. Created with BioRender.com (accessed on 2 August 2022).
Cells 12 01464 g001
Figure 3. Diversity of epidrug action. Epidrugs interfere with fundamental cellular processes and provide effective maintenance of cells: on the tissue level, they affect complex phenomena of neuroprotection, neuroinflammation, and cellular integration (including neuroplasticity); systemically, they affect not only developmental and neoplastic processes but also complex biological programs. Such a spectrum of target functions suggests antiaging, anti-inflammatory, metabolic, and rehabilitative effects of epidrugs.
Figure 3. Diversity of epidrug action. Epidrugs interfere with fundamental cellular processes and provide effective maintenance of cells: on the tissue level, they affect complex phenomena of neuroprotection, neuroinflammation, and cellular integration (including neuroplasticity); systemically, they affect not only developmental and neoplastic processes but also complex biological programs. Such a spectrum of target functions suggests antiaging, anti-inflammatory, metabolic, and rehabilitative effects of epidrugs.
Cells 12 01464 g003
Table 1. Analytical methods of epigenomics.
Table 1. Analytical methods of epigenomics.
Epigenetic
Phenomenon
TargetsAssay TypesAnalytic MethodsResolutions
DNA methylationMethylome analysis or
determination of DNA
methylation/demethylation levels in CpG DNA islands
Microarray-basedDigestion-based
(CHARM, DMH, MCAM, MDRE,
MSRE, and HELP)
Bulk, regional, gene-specific
Affinity enrichment
(MBD and MeDIP)
Bulk, regional
Bisulphite conversion
(HumanMethylation and
EPICBeadChip)
Bulk, single CpG
Sequencing-basedDigestion-based
(HELP-NGS)
Bulk, multiple
genomic regions
Affinity enrichment
(MeDIP and MIRA-seq)
Bulk, regional
Bisulphite conversion
(BS-seq, BSPP, OxBS-seq, RRBS, TAB-seq, Truseq EPIC, and sWGBS)
Bulk, single-cell,
genome-wide,
multiple genomic
regions
Chromatin accessibility and histone
modifications
Chromatin proteomic profiling technologies allow for
investigation of DNA–protein interactions, chromatin
accessibility analysis,
nucleosome positioning
(nucleosome mapping), transcription factor (TF)
binding analysis, etc.
Sequencing-basedAffinity enrichment (ChIP-seq and its
variations, CUT&RUN, CUT&TAG, and DamID)
Bulk, single-cell, multiple genomic
regions
Open chromatin
digestion-based
(DNase-seq and
ATAC-seq)
Bulk, single-cell, multiple genomic
regions
Nuclear organizationMethods of chromosome
conformation capture and
genome 3D structure analysis, including spatial interactions within topologically
associating domains (TADs),
the activity of polycomb-group proteins and helicases, and the assessment of the
architecture of the cell nucleus
Chromatin
conformation
capture
3C/4C/5C, Hi-C,
ChIA-PET, HiChIP, and
PLAC-seq
Bulk, single-cell,
genome-wide,
multiple genomic
regions
Regulatory RNAs and epigenetic enzymesTranscriptome analysis with RNA sequencing and
expression profiling of non-coding ribonucleic acids (ncRNAs), such as microRNA (miRNA) and long non-coding RNA (lncRNA), which
regulate transcription and translation; molecular
biological assessment of
expression and activities of
the epigenetic enzymes
Sequencing- and multi-OMICS-basedRNA-seq, RIP-seq,
NOMe-seq, and
EpiMethylTag
Bulk, single-cell, multiple genomic
regions
Table 3. Fields of epidrug studies with respect to their effects in the CNS.
Table 3. Fields of epidrug studies with respect to their effects in the CNS.
Epidrug ClusterInstancesSome PK/PD FeaturesRepresentative Preclinical
and Clinical Studies
First-generation DNMTis5-azacytidine
5-aza-dioxycytidine
antimetabolitespotentiation of neurotoxic effects in
the culture of dopamine neurons [43]
Second-generation DNMTishydralazine
zebularine
RG108
procainamide
selective inhibition of DNMT isoforms and
inhibition of DNMTs
and cytidine deaminase
interference with β-amyloid production [44];
enhancement of neurotoxicity [45]; and
suppression of apoptosis in motor
neurons [46]
First-generation
HDACis
trichostatin A
vorinostat (SAHA)
romidepsin
reversible binding to
Zn2+ in the HDAC
catalytic center selective for HDAC I, II class
enzymes
prevention of stress-related
behavioral changes in mice [47,48]
Second-generation
HDACis
belinostat
panobinostat
benzamides (e.g., MS-275)
carboxylic acid derivatives (e.g., valproic acid)
improved bioavailability and less toxicityadjuvant therapy for glioma [49];
prevention of stress-related
behavioral changes in mice,
antidepressant effect [50,51]; and
reduction in anxiety and
panic attacks in humans [52,53,54]
HATisexifonegenome stabilizationneuroprotection [55]
HMTistazemetostat
JNJ-64619178
duality of action: the effect is determined by the
position of the
methylated lysine
therapy for glioma [56]
LSDistranylcypromine ORY-1001FAD-dependent inhibition (similar to inhibition of homologous LSD1 and LSD2 monoamine oxidase MAO)change in Bdnf transcription,
antidepressant effect in mice [57]; and
alteration of Bdnf transcription and memory suppression [58,59]
BETisRVX-208
JQ-1
diazepine derivatives
high-affinity BET
inhibitors (antitumor and
expression-regulating ApoA1 and HDL activity)
positive effect on neurogenesis in vitro,
modulation of memory and learning
mechanisms in mice [60,61]; and
improving cognitive performance
in humans [62]
Table 4. Clinical trials of epidrugs with primary epigenetic activity for treating CNS diseases. A + C represents the number of drug-respective trials which are active or completed; T + W represents the number of drug-respective trials which are terminated + withdrawn (January 2023); *—Evaluation of the safety of chronic use of vorinostat at a dose of 100–400 mg per day for the treatment of Alzheimer’s disease.
Table 4. Clinical trials of epidrugs with primary epigenetic activity for treating CNS diseases. A + C represents the number of drug-respective trials which are active or completed; T + W represents the number of drug-respective trials which are terminated + withdrawn (January 2023); *—Evaluation of the safety of chronic use of vorinostat at a dose of 100–400 mg per day for the treatment of Alzheimer’s disease.
CompoundBrain CancerNeurodegenerative DisordersPsychopathological SyndromesOthers (Epilepsy;
Pain Therapy)
A + CT + WA + CT + WA + CT + WA + CT + W
5-azacytidine30000000
belinostat10000000
vorinostat1931 *00000
decitabine01000000
panobinostat53000000
romidepsin10000000
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gladkova, M.G.; Leidmaa, E.; Anderzhanova, E.A. Epidrugs in the Therapy of Central Nervous System Disorders: A Way to Drive on? Cells 2023, 12, 1464. https://doi.org/10.3390/cells12111464

AMA Style

Gladkova MG, Leidmaa E, Anderzhanova EA. Epidrugs in the Therapy of Central Nervous System Disorders: A Way to Drive on? Cells. 2023; 12(11):1464. https://doi.org/10.3390/cells12111464

Chicago/Turabian Style

Gladkova, Marina G., Este Leidmaa, and Elmira A. Anderzhanova. 2023. "Epidrugs in the Therapy of Central Nervous System Disorders: A Way to Drive on?" Cells 12, no. 11: 1464. https://doi.org/10.3390/cells12111464

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop