Skip to main content
Genes & Diseases logoLink to Genes & Diseases
. 2024 Dec 4;12(4):101481. doi: 10.1016/j.gendis.2024.101481

Epigenetic regulation of nuclear receptors: Implications for endocrine-related diseases and therapeutic strategies

Yixin Song a,1, Kexin Zhang a,1, Jingwen Zhang a, Qinying Li a, Na Huang a, Yujie Ma a, Ningning Hou a, Fang Han b, Chengxia Kan a,⁎⁎, Xiaodong Sun a,
PMCID: PMC12022648  PMID: 40290121

Abstract

The expression and function of the receptor are controlled by epigenetic changes, such as DNA methylation, histone modification, and noncoding RNAs. These modifications play a pivotal role in receptor activity and can lead to or exacerbate endocrine-related diseases. This review examines the epigenetic alterations of nuclear receptors and their significant impact on conditions such as diabetes, thyroid disorders, and endocrine-related tumors. It highlights current therapies targeting these epigenetic mechanisms, including gene editing, epigenetic drugs, and various other therapeutic approaches. This review offers fresh insight into the mechanisms of endocrine-associated disorders, highlighting the latest progress in the development of novel epigenetic therapies that can be used to address receptor–endocrine interactions.

Keywords: Androgen receptor, Breast cancer, DNA methylation, Epigenetic modulation, Estrogen receptor, Histone modification, Nuclear receptor

Introduction

The endocrine system, a complex network of glands like the pituitary, thyroid, adrenal glands, and pancreas, regulates various bodily functions by releasing hormones directly into the bloodstream.1,2 These hormones bind to specific receptors, initiating biological effects crucial for metabolism, growth, stress response, immune function, and mental health.3,4 Hormones and their receptors are pivotal components of this system.1 Recent research underscores the significance of epigenetic control in regulating hormone receptor function.5

Hormones and their receptors primarily function through genomic and non-genomic signaling pathways. Upon ligand binding, nuclear receptors (NRs) undergo conformational changes, leading to their migration into the nucleus where they bind to DNA.6 Subsequently, accessory regulators, chromatin remodeling factors, and normal transcription mechanisms regulate the expression of NR target genes. Steroid hormones often elicit non-genomic signals, wherein their interaction with hormone receptors typically triggers diverse protein kinase pathways. These pathways can indirectly influence gene expression through the phosphorylation of transcription factors.7 Steroid hormones are ligand-induced transcription factors that achieve synchronized and precise functional responses by regulating the expression of genetic programs.8 Epigenetics, which modifies DNA without altering coding areas, influences receptor expression and functionality through DNA methylation, histone modification, chromatin restructuring, and non-coding RNA (ncRNA) regulation.9, 10, 11 These mechanisms can affect, in a stable and potentially heritable manner, NRs that regulate transcription and cell surface receptors that modulate transcription factor responsiveness by altering chromatin structure.12,13

NRs serve as sensors for intrinsic cellular changes that impact the transcription of genes involved in key biological processes, such as inflammation, proliferation, apoptosis, and susceptibility to chronic conditions like diabetes.12 The transcriptional activity of receptor genes and the functionality of enzymes related to receptor epigenetics are regulated by four major epigenetic mechanisms, which together control receptor expression levels and influence the onset and progression of endocrine-related diseases.14,15 This review seeks to elucidate the epigenetic control processes involved in various receptor types in endocrine diseases, providing essential scientific insights that may inform the development of innovative therapeutic approaches. Understanding the complex interplay between epigenetics and hormone receptors is crucial for unraveling the pathogenesis of these diseases and advancing treatment strategies.

Epigenetic regulation of receptors

Epigenetic regulation is pivotal in controlling receptor expression, particularly NRs, which are vital for diverse physiological processes. Epigenetic regulation usually includes DNA methylation, histone modifications, RNA-based mechanisms, and chromatin remodeling (Fig. 1). Aberrant epigenetic changes are implicated in diseases like breast cancer (BC) and osteoporosis. Understanding these epigenetic mechanisms can provide insights into receptor-associated diseases and offer potential therapeutic avenues for intervention.

Figure 1.

Figure 1

Epigenetic regulation of receptors. ATP, adenosine triphosphate; ADP, adenosine diphosphate; AC, acetylation; DNMT3A, DNA methyltransferases 3A; DNMT3B, DNA methyltransferases 3B; FTO, the fat mass and obesity associated gene; HDAC, histone deacetylase; HMT, histone methyltransferase; ISWI, imitation switch; ME, methylation; NFR, nucleosome-free region; RSC, remodel the structure of chromatin; SWI/SNF, switch/sucrose non fermentable; TET, the ten-eleven translocation.

Effect of DNA methylation on epigenetic regulation of receptor

DNA methylation is a common epigenetic mechanism in eukaryotic cells, occurring at cytosine residues within cytosine-guanine (CpG) dinucleotides.16 In mammals, DNA methyltransferases (DNMTs), including DNMT1, DNMT3A, and DNMT3B, regulate methylation patterns by adding a methyl group to the fifth carbon of cytosine.17 In sporadic breast tumors, BRCA1 promoter hypermethylation is partly linked to DNMT3B overexpression.18 DNA methylation can influence hormone receptor levels, as seen in the NF-κB ligand–osteoprotegerin axis. For instance, hypermethylation of RANKL and osteoprotegerin gene promoters has been observed in osteoporosis fracture patients, indicating a potential pathogenic role in primary osteoporosis.19

Epigenetic regulation of hormone receptors is crucial for detecting intrinsic changes and controlling the transcription of relevant genes12 (Table 1). Aberrant DNA methylation of the type 1A angiotensin II receptor gene is linked to fetal development and later onset of salt-sensitive hypertension.20 Elevated estrogen receptor (ER) and glucocorticoid receptor (GR) activity in offspring exposed to intense maternal care is associated with promoter DNA methylation of these receptors.21 Although DNA methylation has been implicated in various diseases, few studies have explored its role in regulating NR expression at the promoter level.

Table 1.

Epigenetic regulation and effects of receptors in endocrine-related diseases.

NR Diseases Epigenetic biomarkers Receptor expression levels Related epigenetic regulation Epigenetic effects in diseases Reference
ER BC ESR Methylation of ESR2 promoter; H3K27 residue methylation suppresses transcription Promotes BC development 62,65
KMT2D KMT2D enhances ER-α activity in BYL719-treated PIK3CA mutant BC Promotes PIK3CA-mutant BC development 63
HDAC1, PRC2, NurD Facilitates CpG island methylation during ERE formation Promotes BC cell proliferation 64
GATA1 Prevent histone deacetylation and obstructing ER-α DNA binding at ERE sites Promotes BC cell development 66
SRC3, p300/CBP, CARM1, P160P300 Activates ER transcription Promotes BC cell proliferation 67,68
FOXA1, GRHL2, GATA3 Regulates ER-⍺ signaling via enhancer elements Causes endocrine resistance in ER-positive BC 71, 72, 73
HOTAIR lncRNA HOTAIR associated with PRC2 and LSD1 Promotes tumor growth and metastasis 74
GT3-INCP Regulated by ER and GATA3 Up-regulated in ER-positive BC; drives tumor growth 76
Endometriosis SF-1 Acetylation enrichment of H3 and H4 identified in the SF-1 promoter Causes endometriosis 78,79
GATA 6 GATA6 hypomethylation elevated in proliferating stromal cells Causes endometriosis 80,81
ESR1 The ESR 1 promoter shows partial hypoacetylation at H3 and H4 Increases endometriotic cell proliferation 82
SRA1 SRA1 inhibits ER-α levels Induces apoptosis in endometrial stromal cells 84,85
PR Endometriosis miR-196a miR-196a overexpression in stromal cells suppresses PR function Aggravates endometriosis 83
AR PC FOXA1, NSD2 FOXA1 reprograms AR; NSD2 promotes mutation-driven AR/FOXA1 in PCs Accelerates oncogenic transcription processes 87
miR-194 Affects FOXA1 and stimulates ERK signaling AR-targeted therapy promotes PC cell metastasis 90
HOTAIR HOTAIR-AR interaction inhibits HOTAIR ubiquitination and degradation lnRNA induces tolerance to male deprivation therapy and drug resistance to deprivation in tumor cells 94
CTBP1-AS Induces AR conduction to prevent miRNA-regulated AR degradation 95
miR-193a-5p Disrupts AR-STAT3 connection; triggers ROS Crucial for cornin-induced apoptosis 96
ARLNC1 Stabilizes AR transcripts 97
ER Ovarian cancer ESR1 Abundant DNA methylation at ESR1 promoter Maintains cells in a secretory state 100
KDM1A KDM1A recruitment inhibits ER-β promoter Elevates ovarian cancer metastasis risk 101
miR-193a E2F6 targets epigenetic silencing of miR-193a Promotes ovarian cancer development 102
SMYD2 Inhibits transactivation through ER methylation Enhances cell vitality in CCOC 103,104
JMJD2C, LSD1 The monomethylation and dimethylation of H3K4 and H3K9 inhibits transcription Induces cancer proliferation 53,101
GR Chronic stress NEDD 4 Chronic stress induces NEDD4 expression, downregulating AMPA receptors via GR regulation Potential therapeutic targets for chronic stress-induced cognitive issues 115
TR Thyroid cancer TRβ Hypermethylation in thyroid carcinoma 107
TSHR AITD IFNα Induces genome-wide H3K4me1 modification in thyroid cells TSHR mRNA expression enables TSHR T cells to escape tolerance 109
PPAR Obesity MOF MOF-induced acetylation of H4K16 serves as a modulator by interacting with PPARγ Promotes glucose uptake, lipid storage, and obesity 14
JHDM2A/KDM3A Decreased levels of H3K9me2 through PPAR 111
Bladder cancer PPARγ activation promotes cell proliferation, survival and migration Supports tumor growth and metastasis 117,118
VDR T2DM BAF complex Acetylation of lysine 91 (K91Ac) in VDR is the docking site of ATP-dependent chromatin remodeling complex (BAF complex) BAF complex crucial in T2DM 113
Adrenal cortical carcinoma Hypermethylation of cytosine nucleotide in CpG island of adrenal VDR promoter Leads to loss of VDR protection 112
LXR T2DM Hypermethylation on LXR promoter Disrupts lipid metabolism 65,116
IGF1 Dwarfism CG-137 Methylation of CG cluster in IGF1 P2 promoter IGF-1 methylation inversely affects GH response 23
PGC-1 T2DM PGC-1 DNA methylation of PGC-1α gene promoter 110

Note: AR, androgen receptor; AITD, autoimmune thyroid diseases; AMPA, alpha-amino-3-hydroxy-5-methyl-4-isoxazole propionic acid; BAF, the BRG1/BRM associated factors complex; BC, breast cancer; CCOC, clear cell ovarian carcinoma; CpG, cytosine-guanine; ERK, extracellular signal-regulated kinase; ESC, endometrial stromal cells; ERE, estrogen response element; ER-α, estrogen receptor alpha; FOXA1, Forkhead box A1; GH, growth hormone; GR, glucocorticoid receptor; GATA3, GATA-binding protein 3; GATA 6, GATA-binding factor-6; GRHL2, Grainyhead like 2; HDAC1, histone deacetylase; H3K27, histone-3 lysine-27; IGF-1, Insulin like growth factor 1; IFNα, interferon alpha; JMJD2C, the Jumonji C domain-containing histone demethylases 2C; KDM, histone lysine demethylase; LXR, liver X receptor; LSD1, lysine-specific demethylase1; NurD, the nucleosome-remodeling and deacetylase; OPN, osteopontin; PR, progesterone receptor; PRC2, Polycomb repressive complex 2; PPAR, peroxisome proliferator-activated receptor; PGC-1α, proliferator-activated receptor gamma coactivator-1 alpha; SRA1, steroid receptor RNA activator1; STAT3, signal transducer and activator of transcription 3; SRC3, steroid receptor coactivator 3; SF-1, steroidogenic factor 1; TR, thyroid hormone; TET, the ten-eleven translocation; VDR, vitamin D receptor.

Effect of histone modification on epigenetic regulation of receptor

Histone octamers frequently undergo post-translational modifications, which are key to transcriptional regulation.22 Histone acetylation, methylation, phosphorylation, ubiquitination, and glycosylation regulate gene activity by modifying both histones and transcription factors.23 NRs form corepressor complexes with elements containing histone deacetylase (HDAC), histone methyltransferase, histone demethylase (KDM), and phosphatase functions, either in the absence of a ligand or upon NRs' attachment to an antagonist.24 For instance, alterations in deacetylation can trigger gene activity, while changes in methylation can suppress it. Histone alterations, whether added or removed, contribute to precise control over gene expression.25

Ligand binding induces structural changes in NRs, which control the recruitment of coactivator and co-repressor complexes essential for chromatin modification, thereby affecting transcriptional access to DNA.26 Corepressor complexes involved in gene silencing, such as HDAC3, suppressor interacting 3a, and nuclear receptor co-repressor, maintain target genes in a repressed state by associating with HDAC.27 Specific histone modifications by NR corepressors, like H3K9me3, H3K27me3, and H3K20me3, are markers of heterochromatin suppression. Epigenetic NR suppression is further aided by ubiquitin-mediated degradation of co-repressors.28 Ligand activation at NR sites, including AR, ER, peroxisome proliferator-activated receptor gamma (PPARγ), and vitamin D receptor (VDR), is linked to increased histone acetylation at adjacent regions.29, 30, 31 Proliferator-activated receptor gamma coactivator-1 alpha (PGC-1α) plays a key role in tissue metabolism by activating transcription factors for inflammation and mitochondrial genes, maintaining oxidative capacity with PPARα and ER-α.32 NR agonists trigger co-repressor removal, recruiting coactivators like CREB-binding protein and p300 to H3 and H4 acetylated promoters.33,34

Effect of RNA-based mechanisms on epigenetic regulation of receptor

MicroRNAs (miRNAs) act as epigenetic regulators, impacting about 30% of the mammalian genome by modulating protein levels of target mRNAs without changing genetic sequences.35 miRNAs regulate epigenetic processes by targeting key enzymes like HDACs and DNMTs, thereby influencing DNA methylation and altering the genome's methylation patterns.36,37 For instance, Denis et al38 demonstrated that the KDM5B is regulated by miR-138 in BC. Additionally, the addition of an HDAC inhibitor, OBP-801, to a cell line inhibits the activity of the male receptor by increasing miRNA levels after transcription, thus impeding the development of tumor-like tumors.39

M6A, the most prevalent mRNA modification, significantly impacts miRNA expression post-transcription.40 ncRNAs aid DNA methylation and serve as scaffolds for histone modification complexes.41,42 miR-29b targets DNMTs and ten-eleven translocation (TET) to modulate DNA methylation.43 lncRNAs can recruit or repel DNA modifiers to specific genes, acting as protein scaffolds to drive DNMT degradation via ubiquitin, thereby influencing gene expression in conditions like obesity-induced beta cell activity and hepatocellular carcinoma.44,45 By altering S-adenosyl-l-methionine levels, lncRNAs regulate DNMT activity, impacting DNMT and TET expression at various stages.46 Thus, detecting miRNA methylation and related enzymes may aid disease diagnosis and prognosis. Epigenetic modulators like 5-AZA, LBH589, and GSK-J4 can influence the miRNA-epigenetic feedback loop.47

Effect of chromatin remodeling on epigenetic regulation of receptor

Chromatin remodeling profoundly influences gene transcription by modifying chromatin accessibility to transcription complexes, thereby altering cell phenotype.48 Epigenetic processes can be triggered by DNA methylation and specific histone 3 modifications (such as H3K9 or H3K27), leading to changes in chromatin architecture and heterochromatin alterations.49 NRs interact with histone-modifying proteins that regulate transcriptional activation or inhibition. These remodelers modify the chromatin structure of target gene promoters through post-translational modifications of histone tails and DNA-histone interactions.50 Ligand-bound NRs facilitate transcription by recruiting remodeling enzymes to make promoters accessible. In the absence of ligands, some NRs promote closed chromatin for gene silencing.51 For instance, inhibiting lysine-specific demethylase 1 (LSD1) disrupts androgen receptor (AR)-dependent gene expression by blocking H3K9 demethylation.52 KDM2C independently binds to prostate-specific antigen promoter chromatin and, along with ligand-activated AR, is essential for H3K9 demethylation. It interacts with LSD1 to enhance AR-dependent transactivation.53

Epigenetic regulation of receptors in endocrine-related diseases

Epigenetic regulation of BC induced by ER

The ER plays a crucial role in BC and is encoded by two genes, ESR1 and ESR2 (estrogen receptor 1/2), which produce ER-α and ER-β, respectively.54,55 In BC cases involving ESR1, genes like progesterone receptor (PR), epoxy hydrolase 2, lipocalin 2, and interferon α-inducible protein 27 are silenced due to CpG island methylation.55, 56, 57 Recent research suggests that DNA methylation regulates ER-mediated intercellular adhesion genes, potentially promoting metastatic BC involvement.58 ER-α contributes to DNA methylation processes by recruiting DNMTs, which alter transcription initiation and drive methylation at specific sites.59 In BC, DNA demethylation is a primary trigger for gene transcription, with ER-α actively removing methyl groups from cytosine residues to promote hypomethylation in ER-α-positive BC cells.60,61 ESR2 promoter methylation has been noted in Chinese female ER-β BC patients, suggesting a role in ER-β regulation and BC pathology.62 Histone modifications also influence ER-α activity. The epigenetic regulator KMT2D (lysine methyltransferase 2D), a H3K4-methyltransferase, is associated with increased ER-α activity in PIK3CA mutant BC.63

As a transcription factor, ER synergizes with co-regulators and co–regulatory complexes, producing diverse epigenetic effects that variably impact BC development. ER-α recruits corepressors, such as HDAC1 and the polycomb repressive complex 2, and utilizes complexes like the nucleosome-remodeling and deacetylase complex to facilitate CpG island methylation during estrogen response element formation.64 However, this methylation alone is insufficient for complete transcriptional repression. ER-α further silences gene transcription by altering chromatin structure, leading to H3K27me, a marker of transcriptional repression.65 Conversely, GATA binding 1 acts as a transcriptional repressor by modulating ER-α interactions with histone modification complexes, preventing histone deacetylation and obstructing ER-α DNA binding at estrogen response element sites66 (Fig. 2A).

Figure 2.

Figure 2

Epigenetic regulation of receptors in endocrine-related diseases. (A) Epigenetic mechanisms of estrogen receptor in endocrine-related diseases. (B) Epigenetic mechanisms of androgen receptor in endocrine-related diseases. (C) Epigenetic mechanisms of other hormone receptors in endocrine-related diseases. AC, acetylation; AP-1, activator protein-1; ARE, androgen response element; AR, androgen receptor; APM, antigen-processing machinery; DNMT, DNA methyltransferase; ERE, estrogen response element; EZH2, enhancer of zeste homolog 2; ESR1, estrogen receptor 1; FOXA1, Forkhead box A1; FXRE, FXR response element; HDAC, histone deacetylase; HATs, histone acetyltransferases; HDMs, histone demethylases; HMTs, histone methyltransferases; IRF, interferon regulatory factor; IFNα, interferon alpha; lncRNA, long noncoding RNA; GRHL2, Grainyhead like 2; GATA6, GATA-binding factor 6; GATA3, GATA-binding factor 3; GRE, glucocorticoid response element; LSD-1, lysine-specific histone demethylase 1; TRPS1, trichorhinophalangeal syndrome-1; MBD, methyl-CpG-binding domain; MeCP2, methyl-CpG binding protein 2; ME, methylation; NF-kB, nuclear transcription factor-kappa B; NCOR, nuclear receptor co-repressor; NurD, the nucleosome-remodeling and deacetylase; PRC2, Polycomb repressive complex 2; PGR, progesterone receptor; PGC-1, peroxisome proliferator-activated receptor-gamma coactivator-1; SRC-1, steroid receptor coactivator 1; SRA1, steroid receptor RNA activator 1; SF-1, steroidogenic factor 1; TET, ten-eleven translocation; TSHR, thyrotropin receptor; Sin3a, suppressor interacting 3a; 5 mC, 5-methylcytosine; 5hmC, 5-hydroxymethylcytosin.

In addition, ER recruits the primary steroid receptor coactivator 3 (SRC3), along with secondary coactivators p300/CBP and CARM1, to activate transcription.67 ER-α also interacts with histone acetyltransferases, such as p160 and p300, which support transcriptional activation by acetylating histones.68 Aberrant DNA methylation, histone modifications, and chromatin remodeling arising from this complex interplay of ER-α activity and associated receptor complexes provide insights into BC mechanisms, offering potential targets for therapeutic interventions.

Forkhead box A1 (FOXA1) and Grainyhead-like 2 (GRHL2) function as pioneer factors that facilitate chromatin accessibility and regulate ER-⍺ signaling in hormone receptor-positive BC. Elevated FOXA1 levels are linked to metastasis in endocrine therapy-resistant BC, as it promotes a pro-metastatic secretome.69,70 Together with GRHL2, FOXA1 collaborates with GATA-binding protein 3 and ER-⍺ to utilize enhancer elements in modulating ER-⍺ transcription and contribute to therapeutic resistance.71, 72, 73

Non-coding RNAs, including miRNAs, play a role in regulating metastatic niche formation and metabolic reprogramming by modulating target genes, thereby influencing BC initiation and progression.74 The lncRNA HOTAIR, associated with polycomb repressive complex 2 and LSD1, induces epigenetic changes that promote tumor growth and metastasis.74 In ER-positive BC, 28 functional open reading frames encoded by cryptic lncRNAs are up-regulated, particularly in ductal BC.75 The peptide GT3-INCP, encoded by LINC00992 and regulated by estrogen and ER, promotes tumor growth by modulating susceptibility and risk genes through the transcription factor GATA-binding protein 3.76 Exosome-derived miRNAs also show potential for clinical diagnosis and treatment by influencing tumor growth, progression, and organ-specific targeting77 (Fig. 2A).

Epigenetic regulation of endometriosis induced by ER and PR

The epigenetic control of receptor expression could be crucial in the development of endometriosis. Steroidogenic factor 1 (SF-1), a crucial transcription factor, activates numerous steroid synthesis genes to facilitate steroid production. Differential methylation of the SF-1 gene promoter and CpG islands surrounding the exon I region modulate its expression.78 In proliferative stromal cells, the methyl-CpG-binding protein 2 fails to bind to the unmethylated SF-1 promoter. Moreover, increased acetylation of H3 and H4 is observed in the SF-1 promoter, potentially leading to SF-1 overexpression79 (Fig. 2A).

GATA-binding factor-6 (GATA6) serves as a biomarker for endometriosis induction. Notably, in proliferative stromal cells, hypomethylation in both the promoter and coding region of GATA6 is significantly elevated.80 SF-1 and GATA6 are pivotal for initiating cascade reactions linked to steroidogenic proteins and enzymes, ultimately contributing to the development of endometriosis.81 DNA methylation levels also directly influence the expression of ESR2, ESR1, and PR. Reduced acetylation at the H3 and H4 sites of the ESR1 promoter, particularly at H3, may result in decreased expression of ESR1.82 Furthermore, miR-196a is overexpressed in endometrial stromal cells in endometriosis, inhibiting PR expression and exacerbating the condition.83 Additionally, steroid receptor RNA activator1 lncRNAs regulate ER expression, where suppression of ER-α via this mechanism hinders cell growth and promotes apoptosis in endometrial stromal cells as endometriosis progresses84,85 (Fig. 2A).

Epigenetic regulation of prostate tumor induced by AR

Recent research in China found that 41 % of primary prostate cancers (PC) harbor FOXA1 mutations, which are critical for AR signaling regulation during prostate development and transformation. FOXA1 reprograms AR binding, accelerating carcinogenic transcription.86,87 The NSD2 subunit is essential for AR/FOXA1 neo-enhancer-driven prostate tumors.88 miR-194, initially a circulating marker for post-surgical PC recurrence, also targets FOXA1, activating ERK signaling and potentially contributing to resistance against AR-targeted therapies and metastasis.89,90

AR regulates the expression of TET proteins, which convert DNA 5-methylcytosine into 5-hydroxymethylcytosine and are involved in DNMT production and function.91 Various lncRNAs also play significant roles in PC progression by modulating both AR-dependent and -independent pathways.92 Notably, lncRNAs like PCA3, HOTAIR, and CTBP1-AS are linked to AR pathway stimulation,93 with HOTAIR promoting cancer growth by directly interacting with AR, inhibiting its ubiquitination and degradation, and enhancing AR target gene expression independently of androgens.94 Furthermore, CBR3-AS1 can support AR signaling by preventing miRNA-mediated AR degradation.95 Recent findings also indicate that microRNA-193a-5p-induced reactive oxygen species production, along with the disrupted colocalization of STAT3 and AR, is crucial for cornin-induced apoptosis.96 Finally, it has been demonstrated that AR-regulated lncRNA 1 stabilizes AR transcripts and promotes the translation of AR mRNA transcripts, thereby increasing AR expression.97 The results indicate that lncRNAs might be key targets for epigenetic medications aimed at reducing tolerance to androgen deprivation therapy and resistance to prostate cancer in cancerous cells98 (Fig. 2B).

Epigenetic regulation of ovarian cancer induced by ER and AR

The role of ER epigenetic regulation is recognized in various ovarian cancer cases.99 In clear cell ovarian carcinoma, DNA methylation at the ESR1 promoter is particularly enriched, potentially locking cells in a secretory state.100 Lysine-specific histone demethylase 1A (KDM1A) acts as a key epigenetic modulator and fundamental regulator of steroid hormone receptors. By recruiting KDM1A to the ERβ promoter, its expression is suppressed, which may increase the risk of ovarian cancer metastasis.101 Additionally, E2F6, a significant target of ER, plays a crucial role in the epigenetic silencing of miR-193a, a mechanism implicated in ovarian cancer development.102 Another histone modifier, SMYD2 (SET and MYND domain containing 2), a histone methyltransferase, enhances cell vitality in clear cell ovarian carcinoma through ER methylation, whereas inhibiting SMYD2 induces apoptosis in these cancer cells.103,104 In relation to AR regulation, LSD1 interacts with AR, leading to the monomethylation and dimethylation of H3K4 and H3K9. This interaction prevents demethylation of these markers, thereby inhibiting the transcription of AR-related genes. LSD1's involvement has also been linked to the development of ovarian cancer.101

Epigenetic regulation of autoimmune thyroid diseases and thyroid cancer induced by thyrotropin receptor and thyroid hormone

A number of genetic single nucleotide polymorphisms in genes regulated by DNA methylation may lead to malfunction and irregular DNA methylation, heightening the host's vulnerability to autoimmune thyroid diseases.105 Methylation of gene promoters in patients with autoimmune thyroid diseases correlates with alterations in chromatin architecture, resulting in the suppression of gene activity.106 Furthermore, hypermethylation of the TRβ (thyroid hormone receptor beta) gene is prevalent in thyroid cancer, serving as an alternative mechanism for gene silencing.107 miRNAs contribute to the activation of nuclear transcription factor-kappa B (NF-kB), interferon regulatory factor (IRF), and activator protein-1 (AP-1), which in turn promote the production of regulatory genes and autoimmune antibodies108 (Fig. 2C).

At the molecular level, the epigenetic process governing the regulation of thyrotropin receptor gene expression involves a single-nucleotide polymorphism within intron 1. Interferon alpha triggers widespread alterations in the monomethylation on H3K4me within thyroid cells, coinciding with a pair of adjacent single nucleotide polymorphisms within thyrotropin receptor intron 1 (rs1210255 and rs12101261).109 Individuals harboring the Graves' disease risk T allele of the rs12101261 exhibit reduced thyrotropin receptor mRNA expression, promoting thyrotropin receptor-responsive T cell escape from central tolerance (Fig. 2C).

Epigenetic regulation of endocrine-related diseases induced by other hormone receptors

In the islets of type 2 diabetes mellitus, the promoter of PGC-1α gene exhibits a two-fold increase in DNA methylation rate.110 Increasing evidence points to a connection between histone acetylation and obesity. MOF, a lysine acetyltransferase, is involved in H4K16ac. Activation of H4K16ac by MOF modulates glucose absorption and fat storage in adipocytes through its interaction with PPARγ, exacerbating obesity progression.14 Eliminating the histone demethylase JHDM2A/KDM3A suggests that vulnerability to obesity and metabolic syndrome might result in disruptions in fat build-up and glucose processing.111 This reduces H3K9me2 through the binding of Jhdm2a to PPAR-responsive elements.

Research has shown increased cytosine nucleotide methylation in the CpG island of the adrenal VDR promoter in patients with adrenal cortical carcinoma, resulting in diminished protective effects of the VDR protein against malignant tumors.112 Acetylation of lysine 91 in VDR serves as a docking site for the ATP-dependent chromatin remodeling complex (the BRG1/BRM associated factors complex), which is crucial in type 2 diabetes mellitus.113 There is an inverse relationship between the methylation level of the growth presumptive promoter and both the growth rate and mRNA expression levels of growth hormone.114 Additionally, recent findings indicate that methylation of CG in the P2 promoter of the insulin-like growth factor-1 gene (CG-137) correlates negatively with the response of growth and circulation of insulin-like growth factor-1 23 (Fig. 2C).

Persistent stress regulates the expression of the epigenetic writer-induced ubiquitin ligase NEDD4 and reduces AMPA receptor activity via GR-dependent control. These proteins could serve as therapeutic targets for alleviating chronic stress and/or glucocorticoid-related cognitive impairment.115 Observations show hypermethylation at NR's hepatic X receptor, crucial in controlling cholesterol and fatty acid metabolism, linked to changes in the expression of genes targeted by the hepatic X receptor.65,116

PPARγ activation can promote bladder cancer by enhancing cell proliferation, survival, and migration.117 This occurs through the transcriptional regulation of genes involved in inflammation and cell cycle control, thereby contributing to tumor growth and metastatic potential.118 Additionally, PPARγ′s modulation of the tumor microenvironment and angiogenesis further supports cancer progression.119 Treatment with the PPARγ agonist rosiglitazone effectively suppresses tumor growth, and when combined with trametinib (a MEK inhibitor), it induces apoptosis, leading to a significant reduction in tumor size.120 However, therapeutic targeting of PPARγ in bladder cancer remains controversial, as its effects may vary depending on the tumor context and the specific PPARγ ligands used.121

Receptor-based epigenetic therapy for endocrine-related diseases

ER-induced epigenetic therapy for BC

Tamoxifen was the first ER-α targeted medication to receive clinical approval. It functions selectively by competing with E2 to bind to ER-α, thereby hindering the recruitment of coactivators mediated by the ER-α ligand-binding domain.122 Additionally, tamoxifen facilitates the activation of the AF1 domain independently of ligands, leading to weak transcriptional activation in E2-deficient scenarios and partial inhibition under E2 conditions in vitro.123 Furthermore, tamoxifen can trigger interactions between ER and SRC3, thereby regulating the transcriptional activation of ER.124

Dysregulation of the phosphatidylinositol-3-kinase (PI3K)/protein kinase B (Akt) pathway, including PIK3CA activation mutations, is frequent in BC.125 PI3K pathway regulates ER-dependent transcription in BC through AKT phosphorylation of lysine methyltransferase KMT2D.126 The combination of the PI3Kα inhibitor alpelisib with an anti-ER inhibitor has been approved for treatment.127 Recently, a methylation site has also been identified on KMT2D, catalyzed by the lysine methyltransferase SMYD2. SMYD2 deletion attenuated alpelisib-induced KMT2D chromatin binding and alpelisib-mediated changes in gene expression, including ER-dependent transcription.128 This opens up the possibility of SMYD2 inhibitors in combination with PI3Kα/AKT inhibitors for the treatment of ER-positive PIK3CA-mutant BC (Fig. 3).

Figure 3.

Figure 3

Mechanisms of epigenetic therapy in ER-positive PIK3CA-mutant breast cancer. AKT, protein kinase B; FOXA1, Forkhead box protein A1; mTOR, mechanistic target of rapamycin; PI3K, phosphatidylinositol-3-kinase; PBX1, Pre-B-cell leukemia homeobox transcription factor 1.

Epigenetic modifications play a critical role in ER-positive BC, especially concerning endocrine therapy resistance. Small molecule inhibitors, such as HDAC inhibitors (entinostat, vorinostat) and DNA hypomethylation agents (decitabine, 5-azacytidine), have been investigated as re-sensitizing agents.129,130 Approximately 20% of patients with hypermethylation of the ESR1 promoter exhibit reduced ER-α expression, exacerbating disease progression.131 By restoring ER-α expression, HDAC and aromatase inhibitors show promise in reversing resistance to endocrine therapy.132 Decitabine significantly inhibited preclinical metastasis of ER-positive BC, resulting in a notable reduction in tumor growth in the xenograft model.133 Additionally, combination therapies involving estrogens, HDAC inhibitors, and tamoxifen have been effective in re-establishing endocrine sensitivity.134

Recent studies highlight novel approaches to overcoming endocrine resistance in ER-positive BC. Fulvestrant, an ER degrader, has shown efficacy in sensitizing cells to ferroptosis.135 For patients with tamoxifen-resistant infiltrating lobular carcinoma, the FOXA1-ER pathway is associated with this resistance.136 Moreover, targeting LYPD3 (LY6/PLAUR domain containing 3), which is regulated downstream of FOXA1 and GRHL2, can reduce tumor proliferation in cases that resist endocrine treatments. Thus, LYPD3 represents a viable target for endocrine-resistant BC.72 KAT6A and KAT6B (lysine acetyltransferase 6A/B) are histone acetyltransferases with oncogenic roles in BC. The KAT6A/KAT6B inhibitor CTx-648 (PF-9363) blocked histone tail modifications, demonstrating anti-tumor activity in ER-positive BC, including cases resistant to endocrine therapy.69 To overcome endocrine resistance in ER-positive BC, therapies targeting epigenetic modifications during the dormant state post-endocrine therapy and subsequent reactivation have shown potential.137 Differential DNA methylation of estrogen response enhancers is associated with endocrine sensitivity.138

Several new therapeutic strategies have shown promise for ER-positive BC. Large tumor suppressor kinase (LATS) inhibitors, such as VT02956, target the Hippo pathway to suppress ESR1 expression, particularly in endocrine therapy-resistant cases.139 Additionally, cyclin-dependent kinase 4/6 (CDK4/6) inhibitors (e.g., abemaciclib) and AKT inhibitors (e.g., capivasertib) show efficacy when combined with anti-ER agents like fulvestrant.140,141 A novel therapeutic agent, ERX-11, binds to ER and modulates ER co–regulator interactions to inhibit the proliferation of BC cells that are resistant to endocrine therapy and CDK4/6 inhibitors.142 Targeting these epigenetic modifications can effectively counteract endocrine resistance, thereby enhancing the efficacy of endocrine resistance.131

AR-induced epigenetic therapy for PC

Chromatin profiling has identified various castration-resistant prostate cancer (CRPC) subtypes, unveiling potential therapeutic targets tailored to specific epigenetic landscapes.143 AR function in PC, particularly in CRPC, is significantly influenced by a range of enzymatic epigenetic co-regulators, including KDMs, bromodomain-containing protein 4 (BRD4), and enhancer of zeste homolog 2 (EZH2).144 KDMs, especially KDM4 and KDM1, have emerged as critical players in apoptosis regulation, with KDM-targeting inhibitors showing effectiveness in inducing apoptosis in PC cells.145 This positions KDMs as promising therapeutic targets for epigenetic treatments in PC.

BRD4, a member of the bromodomain and extra-terminal domain subfamily, is another key player. By binding to acetylated lysine residues on histones, BRD4 recruits RNA polymerase and facilitates transcription.146 In CRPC, BRD4 collaborates with AR to drive AR-mediated transcriptional activity. Inhibiting BRD4 disrupts AR's ability to bind to its target genes, making it a compelling therapeutic target for advanced PC. Additionally, some CRPC cases exhibit acquired GR activity, which activates the cAMP/PKA signaling pathway, influencing downstream gene expression.147

EZH2, a methyltransferase, acts as an AR co-activator in CRPC by binding to the AR promoter region, enhancing AR signaling.148 Notably, inhibiting EZH2 can counteract resistance to enzalutamide, a widely used anti-androgen medication in CRPC.149 This highlights the therapeutic potential of EZH2 inhibitors in combating drug resistance. Moreover, methylation and degradation of EZH2 by SET domain-containing 2 (SETD2) have been linked to metastasis prevention, further underscoring the therapeutic implications of targeting EZH2.150

Pioneer factors such as FOXA1 and GRHL2 play crucial roles in PC progression and endocrine resistance. Enzalutamide-induced FOXA1 activity, for instance, can open chromatin and drive AR-driven transcription.151 Targeting FOXA1 to inhibit AR activity represents a promising therapeutic avenue in CRPC. Irregular FOXA1 activity is also implicated in PC progression through its effects on the AR cis-regulatory network.152

Other emerging therapeutic approaches include targeting heat shock protein 70 (Hsp70), which binds to the AR N-terminal domain, reducing endogenous AR expression and inhibiting its transcriptional activity.151 This could be particularly useful in overcoming resistance to enzalutamide in CRPC. Recently, ARNTL, a circadian clock component, has been proposed as a novel therapeutic target, while TET2 inhibitors show promise in overcoming resistance to AR-targeted therapies in ZNF397-deficient tumors.153 As the understanding of these epigenetic regulators deepens, epigenetic therapies could provide new strategies for managing drug-resistant PC and other endocrine-related diseases. These insights offer a path to developing more effective treatments for CRPC, addressing resistance mechanisms, and potentially extending these strategies to other forms of PC and endocrine-related conditions.154

ER-induced epigenetic therapy for endometriosis

In interventions for endometriosis, drug research has consistently targeted enzymes that regulate epigenetic alterations. Targeted suppression of prostaglandin E2 receptors EP2 and EP4 has been shown to involve DNMT3A and DNMT3B, with no effect on DNMT1 expression.155 Researchers propose that focusing on EP2 and EP4 receptors could serve as a non-steroidal treatment for active endometrial lesions in women.155 Additionally, demethylating agents have been found to increase ER-β mRNA levels in endometrial tissue, indicating a potential target for epigenetic therapy.156

Other nuclear hormone-induced epigenetic therapy for type 2 diabetes mellitus

Known alternatively as 5-aza-2′-deoxycytidine, decitabine stands as the most potent DNMT inhibitor. Additionally, decitabine suppresses the methylation of PPARγ1 promoter DNA, enhancing macrophage activation and reducing insulin resistance in overweight individuals.157 Furthermore, class I HDAC inhibitors, like MS-275, enhance hyperglycemia and body mass in diabetic mice with insulin-induced obesity.158 This is accomplished through the regulation of mitochondrial function transcription factors and cofactors like PGC-1α and PPARγ, in addition to gene expression linked to glucose and lipid metabolism, encompassing glucose transporter.159

Research also suggests that targeting DNA methylation could be an effective strategy in treating diabetic osteoporosis. Decitabine aids in the osteogenic differentiation of adipose-derived stromal cells by reducing the degree of methylation in osteogenic genes like osteopontin (OPN) and Runt-related transcription factor 2 (RUNX2).160 Vorinostat, by acetylating histone 4, enhances the regulation and phosphorylation of insulin receptor β, AKT, and the forkhead box O1 (FOXO1).161 Additionally, C646 acts as a targeted blocker of P300 acetyltransferase, obstructing insulin receptor substrate 1/2 (IRS1/2) acetylation and facilitating IRS1/2's movement across membranes, resulting in the stimulation of the insulin pathway.162 Methylation of 12 CpG at the start point of the glucagon-like peptide 1 (GLP1) gene transcription has been observed in type 2 diabetes mellitus human islets. The expression of GLP1 receptor can be managed using the GLP1 receptor agonist rivenatide.163

While a range of medications targeting DNA methylation and histone acetylation has been formulated, their application in clinical medical practice is infrequent.164 Despite numerous investigations into ncRNAs and various histone alterations, there has been no authorization for an epigenetic medication targeting metabolic disorders.165 Furthermore, the adverse impacts of these epigenetic medications could stem from modifications in genes that are not targeted.166 Consequently, delving deeper into the function of epigenetic medications in metabolic disorders is of utmost significance (Table 2).

Table 2.

Receptor-based epigenetic therapy and mechanisms for endocrine-related diseases.

Epigenetic therapeutic targets Epigenetics drugs Therapeutic mechanisms Reference
AF1 Tamoxifen Disrupts receptors' competition with E2 for ERα binding and hinders ERα-LBD coactivator recruitment 123
KMT2D
SMYD2
PI3Kα/AKT inhibitors
SMYD2 inhibitors
Sensitizes BC to PI3K/AKT inhibition and endocrine therapy, in part through KMT2D K1330 methylation 125,128
DNMT Decitabine, 5-azacytidine Suppressed the ethylation of PPARγ1 promoter DNA, enhancing macrophage activation 129,130
Enterestat and letrozole Restored ER-α and enzyme expression in ER-BC cell lines, leading to growth inhibition 132
Fulvestrant Sensitized ER-positive BC cells to ferroptosis through down-regulating MBOAT1 135
FOXA1/GRHL2 LYPD3 inhibitors Inhibits proliferation of endocrine-resistant tumors 72
KAT6A/KAT6B KAT6A/6B HAT inhibitors Blocks histone modification, showing anti-tumor activity in ER-positve BC 69
LATS LATS inhibitors (VT02956) Suppresses ESR1, controlling ER-positve BC growth via Hippo pathway 139
CDK4/6
AKT
CDK4/6 inhibitors (abemaciclib)
AKT inhibitors (capivasertib)
Reverses endocrine resistance with fulvestrant 140,141
ER coregulator ERX-11 Binds to ER, modulates coregulator interactions, and inhibits BC proliferation 142
DMA Increases ER-β mRNA levels in endometrium 156
IL-1β, TGF-β Inhibits DNMT1 +32204 GG genotype remission in patients with Graves' disease 105
KDM4 and KDM1 Induces PC cells apoptosis 145
BRD4 Obstructs AR binding and transcriptional activity 146
EZH2 SETD2 Blocks PC metastasis by methylation and EZH2 degradation (SETD2) 150
FOXA1 Provides treatment strategy for CRPC via AR cis-antigenome impact 152
Hsp70 Hsp70 inhibitors Binds to AR N-terminal domain and reduces AR expression and transcriptional activity 151
TET2 TET2 inhibitors Eliminates resistance in ZNF397-deficient tumors to AR therapy 153
HDAC MS-275 Regulates genes in glucose and lipid metabolism 158
OPN and RUNX2 Decitabine Promotes osteogenic differentiation by reducing methylation 160
Histone 4 Vorinostat Enhances regulation and phosphorylation of insulin receptor β, AKT, and FOXO1 161
P300 acetyltransferase C646 Blocks IRS1/2 acetylation and aids IRS1/2 membrane movement 162
GLP1R Rivenatide Suppresses 12 CpG sites methylated at GLP1 transcription start 163

Note: AKT, protein kinase B; AR, androgen receptor; BRD4, bromodomain-containing protein 4; CDK, cyclin-dependent kinases; DNMT, DNA methyltransferases; DMA, demethylating agent; ERXs, estrogen receptor coregulator binding modulators; EZH2, enhancer of zeste homolog 2; ER, estrogen receptor; ERα-LBD, estrogen receptor-α ligand binding domain; FOXO1, Forkhead box O1; FOXA1, Forkhead box A1; GRHL2, Grainyhead like 2; GLP1, glucagon-like peptide-1; GLP1R, glucagon-like peptide-1 receptor; HDAC, histone deacetylase; IL-1β, interleukin-1beta; IRS1/2, insulin receptor substrate 1/2; KDM, histone lysine demethylase; LYPD3, LY6/PLAUR domain containing 3; LATS, large tumor suppressor kinase; MBOAT1, membrane-bound O-acyltransferase domain-containing 1; PPARγ1, peroxisome proliferator-activated receptor gamma 1; RUNX2, Runt-related transcription factor 2; SETD2, SET domain-containing 2; TGF-β, transforming growth factor-beta.

It is well-recognized that demethylation of CpG sites in the promoter region enhances gene expression.167 The ability of Graves' disease patients with the +32204 GG genotype to fully methylate their DNA may correlate with the methylation levels of interleukin-1beta (IL-1β) and transforming growth factor-beta (TGF-β) promoter regions, impacting IL-1β and TGF-β production.105 This mechanism could potentially impede the remission initiation of patients with Graves' disease who have the DNMT1 +32204 GG genotype.

Perspectives and conclusions

Epigenetics research mainly focuses on changes in gene expression, centering on processes that regulate gene expression rather than altering the DNA sequence. These mechanisms include DNA methylation, histone modification, ncRNA regulation, and modulation of receptor gene expression, which collectively influence the functionality of internal systems. Receptors are crucial in controlling diverse internal mechanisms, with their expression and functionality intricately governed by epigenetic processes. This review explores the epigenetic mechanisms involved in receptor-related diseases and potential treatment strategies. For instance, certain drugs can inhibit the activity of DNMT, thereby affecting receptor expression. Furthermore, ncRNAs are recognized for their role in controlling receptor expression and functionality. Despite the progress made in epigenetic therapy for receptors, several challenges remain. Firstly, there is a need to deepen our understanding of the specific epigenetic mechanisms that govern receptor expression and function. Secondly, more effective drugs are required to modulate receptor expression and function through epigenetic mechanisms. Finally, the efficacy and safety of these treatments need to be thoroughly validated through clinical trials.

In summary, epigenetic therapy targeting receptors holds promise as a therapeutic approach for endocrine-related diseases. However, further research and practical implementation are necessary to refine and optimize this treatment strategy to better serve patients.

Funding

This work was supported by the National Natural Science Foundation of China (No. 82170865, 82370856, 32101029), the Natural Science Foundation of Shandong Province, China (No. ZR2020QB164), Shandong Province Medical and Health Science and Technology Development Project of China (No. 202203060805, 202203060917), and Taishan Scholars Project of Shandong Province, China (No. tsqn202211365).

CRediT authorship contribution statement

Yixin Song: Conceptualization, Methodology, Visualization, Writing – original draft, Formal analysis. Kexin Zhang: Conceptualization, Data curation, Methodology, Visualization, Formal analysis. Jingwen Zhang: Data curation, Investigation, Methodology. Qinying Li: Data curation, Funding acquisition, Investigation. Na Huang: Data curation, Formal analysis, Investigation, Methodology. Yujie Ma: Data curation, Investigation. Ningning Hou: Data curation, Investigation, Supervision. Fang Han: Data curation, Investigation. Chengxia Kan: Conceptualization, Data curation, Formal analysis, Methodology, Supervision, Writing – review & editing, Investigation. Xiaodong Sun: Conceptualization, Data curation, Funding acquisition, Supervision, Validation, Writing – review & editing.

Conflict of interests

None competing interests to declare in this study.

Footnotes

Peer review under the responsibility of the Genes & Diseases Editorial Office, in alliance with the Association of Chinese Americans in Cancer Research (ACACR, Baltimore, MD, USA).

Contributor Information

Chengxia Kan, Email: fykanchengxia@sdsmu.edu.cn.

Xiaodong Sun, Email: xiaodong.sun@sdsmu.edu.cn.

References

  • 1.Leblebicioglu B., Connors J., Mariotti A. Principles of endocrinology. Periodontol 2000. 2013;61(1):54–68. doi: 10.1111/j.1600-0757.2011.00440.x. [DOI] [PubMed] [Google Scholar]
  • 2.Jeon S., Lee Y.S., Oh S.R., et al. Recent advances in endocrine organoids for therapeutic application. Adv Drug Deliv Rev. 2023;199 doi: 10.1016/j.addr.2023.114959. [DOI] [PubMed] [Google Scholar]
  • 3.Cox R.M., Hale M.D., Wittman T.N., Robinson C.D., Cox C.L. Evolution of hormone-phenotype couplings and hormone-genome interactions. Horm Behav. 2022;144 doi: 10.1016/j.yhbeh.2022.105216. [DOI] [PubMed] [Google Scholar]
  • 4.Clemente-Suárez V.J., Redondo-Flórez L., Rubio-Zarapuz A., Martín-Rodríguez A., Tornero-Aguilera J.F. Microbiota implications in endocrine-related diseases: from development to novel therapeutic approaches. Biomedicines. 2024;12(1):221. doi: 10.3390/biomedicines12010221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Ozgyin L., Erdős E., Bojcsuk D., Balint B.L. Nuclear receptors in transgenerational epigenetic inheritance. Prog Biophys Mol Biol. 2015;118(1–2):34–43. doi: 10.1016/j.pbiomolbio.2015.02.012. [DOI] [PubMed] [Google Scholar]
  • 6.Shiota M., Fujimoto N., Kashiwagi E., Eto M. The role of nuclear receptors in prostate cancer. Cells. 2019;8(6):602. doi: 10.3390/cells8060602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Vrtačnik P., Ostanek B., Mencej-Bedrač S., Marc J. The many faces of estrogen signaling. Biochem Med. 2014;24(3):329–342. doi: 10.11613/BM.2014.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Malbeteau L., Pham H.T., Eve L., Stallcup M.R., Poulard C., Le Romancer M. How protein methylation regulates steroid receptor function. Endocr Rev. 2022;43(1):160–197. doi: 10.1210/endrev/bnab014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Deeb S.S., Bisset D., Fu L. Epigenetic control of expression of the human L- and M- pigment genes. Ophthalmic Physiol Opt. 2010;30(5):446–453. doi: 10.1111/j.1475-1313.2010.00735.x. [DOI] [PubMed] [Google Scholar]
  • 10.Tammen S.A., Friso S., Choi S.W. Epigenetics: the link between nature and nurture. Mol Aspect Med. 2013;34(4):753–764. doi: 10.1016/j.mam.2012.07.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kovács T., Szabó-Meleg E., Ábrahám I.M. Estradiol-induced epigenetically mediated mechanisms and regulation of gene expression. Int J Mol Sci. 2020;21(9):3177. doi: 10.3390/ijms21093177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Sever R., Glass C.K. Signaling by nuclear receptors. Cold Spring Harbor Perspect Biol. 2013;5(3) doi: 10.1101/cshperspect.a016709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Baumbach J.L., Zovkic I.B. Hormone-epigenome interactions in behavioural regulation. Horm Behav. 2020;118 doi: 10.1016/j.yhbeh.2020.104680. [DOI] [PubMed] [Google Scholar]
  • 14.Pessoa Rodrigues C., Chatterjee A., Wiese M., et al. Histone H4 lysine 16 acetylation controls central carbon metabolism and diet-induced obesity in mice. Nat Commun. 2021;12(1):6212. doi: 10.1038/s41467-021-26277-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Vasan N., Toska E., Scaltriti M. Overview of the relevance of PI3K pathway in HR-positive breast cancer. Ann Oncol. 2019;30(Suppl_10):x3–x11. doi: 10.1093/annonc/mdz281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Dawson M.A., Kouzarides T. Cancer epigenetics: from mechanism to therapy. Cell. 2012;150(1):12–27. doi: 10.1016/j.cell.2012.06.013. [DOI] [PubMed] [Google Scholar]
  • 17.Zhang W., Xu J. DNA methyltransferases and their roles in tumorigenesis. Biomark Res. 2017;5:1. doi: 10.1186/s40364-017-0081-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Butcher D.T., Rodenhiser D.I. Epigenetic inactivation of BRCA1 is associated with aberrant expression of CTCF and DNA methyltransferase (DNMT3B) in some sporadic breast tumours. Eur J Cancer. 2007;43(1):210–219. doi: 10.1016/j.ejca.2006.09.002. [DOI] [PubMed] [Google Scholar]
  • 19.Wang P., Cao Y., Zhan D., et al. Influence of DNA methylation on the expression of OPG/RANKL in primary osteoporosis. Int J Med Sci. 2018;15(13):1480–1485. doi: 10.7150/ijms.27333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Kawarazaki W., Fujita T. Kidney and epigenetic mechanisms of salt-sensitive hypertension. Nat Rev Nephrol. 2021;17(5):350–363. doi: 10.1038/s41581-021-00399-2. [DOI] [PubMed] [Google Scholar]
  • 21.Champagne F.A., Weaver I.C.G., Diorio J., Dymov S., Szyf M., Meaney M.J. Maternal care associated with methylation of the estrogen receptor-alpha1b promoter and estrogen receptor-alpha expression in the medial preoptic area of female offspring. Endocrinology. 2006;147(6):2909–2915. doi: 10.1210/en.2005-1119. [DOI] [PubMed] [Google Scholar]
  • 22.Tessarz P., Kouzarides T. Histone core modifications regulating nucleosome structure and dynamics. Nat Rev Mol Cell Biol. 2014;15(11):703–708. doi: 10.1038/nrm3890. [DOI] [PubMed] [Google Scholar]
  • 23.Ouni M., Belot M.P., Castell A.L., Fradin D., Bougnères P. The P2 promoter of the IGF1 gene is a major epigenetic locus for GH responsiveness. Pharmacogenomics J. 2016;16(1):102–106. doi: 10.1038/tpj.2015.26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Feng Y.Q., Desprat R., Fu H., et al. DNA methylation supports intrinsic epigenetic memory in mammalian cells. PLoS Genet. 2006;2(4) doi: 10.1371/journal.pgen.0020065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Zhong S., Fields C.R., Su N., Pan Y.X., Robertson K.D. Pharmacologic inhibition of epigenetic modifications, coupled with gene expression profiling, reveals novel targets of aberrant DNA methylation and histone deacetylation in lung cancer. Oncogene. 2007;26(18):2621–2634. doi: 10.1038/sj.onc.1210041. [DOI] [PubMed] [Google Scholar]
  • 26.Green C.D., Han J.D.J. Epigenetic regulation by nuclear receptors. Epigenomics. 2011;3(1):59–72. doi: 10.2217/epi.10.75. [DOI] [PubMed] [Google Scholar]
  • 27.Wen Y.D., Perissi V., Staszewski L.M., et al. The histone deacetylase-3 complex contains nuclear receptor corepressors. Proc Natl Acad Sci U S A. 2000;97(13):7202–7207. doi: 10.1073/pnas.97.13.7202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Yoon H.G., Chan D.W., Reynolds A.B., Qin J., Wong J. N-CoR mediates DNA methylation-dependent repression through a methyl CpG binding protein Kaiso. Mol Cell. 2003;12(3):723–734. doi: 10.1016/j.molcel.2003.08.008. [DOI] [PubMed] [Google Scholar]
  • 29.Kininis M., Chen B.S., Diehl A.G., et al. Genomic analyses of transcription factor binding, histone acetylation, and gene expression reveal mechanistically distinct classes of estrogen-regulated promoters. Mol Cell Biol. 2007;27(14):5090–5104. doi: 10.1128/MCB.00083-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Takayama K., Kaneshiro K., Tsutsumi S., et al. Identification of novel androgen response genes in prostate cancer cells by coupling chromatin immunoprecipitation and genomic microarray analysis. Oncogene. 2007;26(30):4453–4463. doi: 10.1038/sj.onc.1210229. [DOI] [PubMed] [Google Scholar]
  • 31.Steger D.J., Grant G.R., Schupp M., et al. Propagation of adipogenic signals through an epigenomic transition state. Genes Dev. 2010;24(10):1035–1044. doi: 10.1101/gad.1907110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Wang Y., Fang F., Wong C.W. Troglitazone is an estrogen-related receptor alpha and gamma inverse agonist. Biochem Pharmacol. 2010;80(1):80–85. doi: 10.1016/j.bcp.2010.03.013. [DOI] [PubMed] [Google Scholar]
  • 33.Nishihara E., O'Malley B.W., Xu J. Nuclear receptor coregulators are new players in nervous system development and function. Mol Neurobiol. 2004;30(3):307–325. doi: 10.1385/MN:30:3:307. [DOI] [PubMed] [Google Scholar]
  • 34.Wang D., Xia X., Weiss R.E., Refetoff S., Yen P.M. Distinct and histone-specific modifications mediate positive versus negative transcriptional regulation of TSHalpha promoter. PLoS One. 2010;5(3) doi: 10.1371/journal.pone.0009853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Bartel D.P. MicroRNAs: target recognition and regulatory functions. Cell. 2009;136(2):215–233. doi: 10.1016/j.cell.2009.01.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Silva-Llanes I., Shin C.H., Jiménez-Villegas J., Gorospe M., Lastres-Becker I. The transcription factor NRF2 has epigenetic regulatory functions modulating HDACs, DNMTs, and miRNA biogenesis. Antioxidants. 2023;12(3):641. doi: 10.3390/antiox12030641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Gu Y., Ci C., Zhang X., et al. Prediction of circRNAs based on the DNA methylation-mediated feature sponge function in breast cancer. Front Bioeng Biotechnol. 2019;7:365. doi: 10.3389/fbioe.2019.00365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Denis H., Van Grembergen O., Delatte B., et al. MicroRNAs regulate KDM5 histone demethylases in breast cancer cells. Mol Biosyst. 2016;12(2):404–413. doi: 10.1039/c5mb00513b. [DOI] [PubMed] [Google Scholar]
  • 39.Sato S., Katsushima K., Shinjo K., et al. Histone deacetylase inhibition in prostate cancer triggers miR-320-mediated suppression of the androgen receptor. Cancer Res. 2016;76(14):4192–4204. doi: 10.1158/0008-5472.CAN-15-3339. [DOI] [PubMed] [Google Scholar]
  • 40.Alarcón C.R., Goodarzi H., Lee H., Liu X., Tavazoie S., Tavazoie S.F. HNRNPA2B1 is a mediator of m(6)A-dependent nuclear RNA processing events. Cell. 2015;162(6):1299–1308. doi: 10.1016/j.cell.2015.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Sun T.T., He J., Liang Q., et al. LncRNA GClnc1 promotes gastric carcinogenesis and may act as a modular scaffold of WDR5 and KAT2A complexes to specify the histone modification pattern. Cancer Discov. 2016;6(7):784–801. doi: 10.1158/2159-8290.CD-15-0921. [DOI] [PubMed] [Google Scholar]
  • 42.Yang Z., Xu F., Wang H., Teschendorff A.E., Xie F., He Y. Pan-cancer characterization of long non-coding RNA and DNA methylation mediated transcriptional dysregulation. EBioMedicine. 2021;68 doi: 10.1016/j.ebiom.2021.103399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Fabbri M., Garzon R., Cimmino A., et al. MicroRNA-29 family reverts aberrant methylation in lung cancer by targeting DNA methyltransferases 3A and 3B. Proc Natl Acad Sci USA. 2007;104(40):15805–15810. doi: 10.1073/pnas.0707628104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Zhang Z., Cao Y., Zhai Y., et al. MicroRNA-29b regulates DNA methylation by targeting Dnmt3a/3b and Tet1/2/3 in porcine early embryo development. Dev Growth Differ. 2018;60(4):197–204. doi: 10.1111/dgd.12537. [DOI] [PubMed] [Google Scholar]
  • 45.Zhang F., Yang Y., Chen X., et al. The long non-coding RNA βFaar regulates islet β-cell function and survival during obesity in mice. Nat Commun. 2021;12(1):3997. doi: 10.1038/s41467-021-24302-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Loaeza-Loaeza J., Beltran A.S., Hernández-Sotelo D. DNMTs and impact of CpG content, transcription factors, consensus motifs, lncRNAs, and histone marks on DNA methylation. Genes. 2020;11(11):1336. doi: 10.3390/genes11111336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Yao Q., Chen Y., Zhou X. The roles of microRNAs in epigenetic regulation. Curr Opin Chem Biol. 2019;51:11–17. doi: 10.1016/j.cbpa.2019.01.024. [DOI] [PubMed] [Google Scholar]
  • 48.Lorch Y., Kornberg R.D. Chromatin-remodeling for transcription. Q Rev Biophys. 2017;50 doi: 10.1017/S003358351700004X. [DOI] [PubMed] [Google Scholar]
  • 49.Lee S.C., Adams D.W., Ipsaro J.J., et al. Chromatin remodeling of histone H3 variants by DDM1 underlies epigenetic inheritance of DNA methylation. Cell. 2023;186(19):4100–4116. doi: 10.1016/j.cell.2023.08.001. e15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Magnani L., Lupien M. Chromatin and epigenetic determinants of estrogen receptor alpha (ESR1) signaling. Mol Cell Endocrinol. 2014;382(1):633–641. doi: 10.1016/j.mce.2013.04.026. [DOI] [PubMed] [Google Scholar]
  • 51.Buranapramest M., Chakravarti D. Chromatin remodeling and nuclear receptor signaling. Prog Mol Biol Transl Sci. 2009;87:193–234. doi: 10.1016/S1877-1173(09)87006-3. [DOI] [PubMed] [Google Scholar]
  • 52.Metzger E., Wissmann M., Yin N., et al. LSD1 demethylates repressive histone marks to promote androgen-receptor-dependent transcription. Nature. 2005;437(7057):436–439. doi: 10.1038/nature04020. [DOI] [PubMed] [Google Scholar]
  • 53.Wissmann M., Yin N., Müller J.M., et al. Cooperative demethylation by JMJD2C and LSD1 promotes androgen receptor-dependent gene expression. Nat Cell Biol. 2007;9(3):347–353. doi: 10.1038/ncb1546. [DOI] [PubMed] [Google Scholar]
  • 54.Mosselman S., Polman J., Dijkema R. ER beta: identification and characterization of a novel human estrogen receptor. FEBS Lett. 1996;392(1):49–53. doi: 10.1016/0014-5793(96)00782-x. [DOI] [PubMed] [Google Scholar]
  • 55.Prabhu J.S., Wahi K., Korlimarla A., et al. The epigenetic silencing of the estrogen receptor (ER) by hypermethylation of the ESR1 promoter is seen predominantly in triple-negative breast cancers in Indian women. Tumor Biol. 2012;33(2):315–323. doi: 10.1007/s13277-012-0343-1. [DOI] [PubMed] [Google Scholar]
  • 56.Stone A., Valdés-Mora F., Gee J.M., et al. Tamoxifen-induced epigenetic silencing of oestrogen-regulated genes in anti-hormone resistant breast cancer. PLoS One. 2012;7(7) doi: 10.1371/journal.pone.0040466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Ariazi E.A., Taylor J.C., Black M.A., et al. A new role for ERα: silencing via DNA methylation of basal, stem cell, and EMT genes. Mol Cancer Res. 2017;15(2):152–164. doi: 10.1158/1541-7786.MCR-16-0283. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Garcia-Recio S., Hinoue T., Wheeler G.L., et al. Multiomics in primary and metastatic breast tumors from the AURORA US network finds microenvironment and epigenetic drivers of metastasis. Nat Can (Ott) 2023;4(1):128–147. doi: 10.1038/s43018-022-00491-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Zhao Z., Fan L., Fortress A.M., Boulware M.I., Frick K.M. Hippocampal histone acetylation regulates object recognition and the estradiol-induced enhancement of object recognition. J Neurosci. 2012;32(7):2344–2351. doi: 10.1523/JNEUROSCI.5819-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Ung M., Ma X., Johnson K.C., Christensen B.C., Cheng C. Effect of estrogen receptor α binding on functional DNA methylation in breast cancer. Epigenetics. 2014;9(4):523–532. doi: 10.4161/epi.27688. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Wang L., Ozark P.A., Smith E.R., et al. TET2 coactivates gene expression through demethylation of enhancers. Sci Adv. 2018;4(11) doi: 10.1126/sciadv.aau6986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Zhao L., Yu Z., Li Y., et al. Clinical implications of ERβ methylation on sporadic breast cancers in Chinese women. Med Oncol. 2012;29(3):1569–1575. doi: 10.1007/s12032-011-0107-6. [DOI] [PubMed] [Google Scholar]
  • 63.Koren S., Bentires-Alj M. Tackling resistance to PI3K inhibition by targeting the epigenome. Cancer Cell. 2017;31(5):616–618. doi: 10.1016/j.ccell.2017.04.010. [DOI] [PubMed] [Google Scholar]
  • 64.Gurevich I., Flores A.M., Aneskievich B.J. Corepressors of agonist-bound nuclear receptors. Toxicol Appl Pharmacol. 2007;223(3):288–298. doi: 10.1016/j.taap.2007.05.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Li R., Sun X., Li P., et al. GLP-1-induced AMPK activation inhibits PARP-1 and promotes LXR-mediated ABCA1 expression to protect pancreatic β-cells against cholesterol-induced toxicity through cholesterol efflux. Front Cell Dev Biol. 2021;9 doi: 10.3389/fcell.2021.646113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Serandour A.A., Mohammed H., Miremadi A., Mulder K.W., Carroll J.S. TRPS1 regulates oestrogen receptor binding and histone acetylation at enhancers. Oncogene. 2018;37(39):5281–5291. doi: 10.1038/s41388-018-0312-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Yi P., Wang Z., Feng Q., et al. Structural and functional impacts of ER coactivator sequential recruitment. Mol Cell. 2017;67(5):733–743. doi: 10.1016/j.molcel.2017.07.026. e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Zwart W., Theodorou V., Kok M., Canisius S., Linn S., Carroll J.S. Oestrogen receptor-co-factor-chromatin specificity in the transcriptional regulation of breast cancer. EMBO J. 2011;30(23):4764–4776. doi: 10.1038/emboj.2011.368. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Sharma S., Chung C.Y., Uryu S., et al. Discovery of a highly potent, selective, orally bioavailable inhibitor of KAT6A/B histone acetyltransferases with efficacy against KAT6A-high ER+ breast cancer. Cell Chem Biol. 2023;30(10):1191–1210.e20. doi: 10.1016/j.chembiol.2023.07.005. [DOI] [PubMed] [Google Scholar]
  • 70.Fu X., Pereira R., Liu C.C., et al. High FOXA1 levels induce ER transcriptional reprogramming, a pro-metastatic secretome, and metastasis in endocrine-resistant breast cancer. Cell Rep. 2023;42(8) doi: 10.1016/j.celrep.2023.112821. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Wang Z., Coban B., Wu H., et al. GRHL2-controlled gene expression networks in luminal breast cancer. Cell Commun Signal. 2023;21(1):15. doi: 10.1186/s12964-022-01029-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Cocce K.J., Jasper J.S., Desautels T.K., et al. The lineage determining factor GRHL2 collaborates with FOXA1 to establish a targetable pathway in endocrine therapy-resistant breast cancer. Cell Rep. 2019;29(4):889–903.e10. doi: 10.1016/j.celrep.2019.09.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Zhang Y., Chan H.L., Garcia-Martinez L., et al. Estrogen induces dynamic ERα and RING1B recruitment to control gene and enhancer activities in luminal breast cancer. Sci Adv. 2020;6(23) doi: 10.1126/sciadv.aaz7249. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Tsai M.C., Manor O., Wan Y., et al. Long noncoding RNA as modular scaffold of histone modification complexes. Science. 2010;329(5992):689–693. doi: 10.1126/science.1192002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Dutta A., Li H., Abounader R. Cryptic lncRNA-encoded ORFs: a hidden source of regulatory proteins. J Clin Invest. 2023;133(5) doi: 10.1172/JCI167271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Zheng C., Wei Y., Zhang P., et al. CRISPR/Cas9 screen uncovers functional translation of cryptic lncRNA-encoded open reading frames in human cancer. J Clin Invest. 2023;133(5) doi: 10.1172/JCI159940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Li J., He D., Bi Y., Liu S. The emerging roles of exosomal miRNAs in breast cancer progression and potential clinical applications. Breast Cancer. 2023;15:825–840. doi: 10.2147/BCTT.S432750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Xue Q., Lin Z., Yin P., et al. Transcriptional activation of steroidogenic factor-1 by hypomethylation of the 5' CpG island in endometriosis. J Clin Endocrinol Metab. 2007;92(8):3261–3267. doi: 10.1210/jc.2007-0494. [DOI] [PubMed] [Google Scholar]
  • 79.Monteiro J.B., Colón-Díaz M., García M., et al. Endometriosis is characterized by a distinct pattern of histone 3 and histone 4 lysine modifications. Reprod Sci. 2014;21(3):305–318. doi: 10.1177/1933719113497267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Dyson M.T., Roqueiro D., Monsivais D., et al. Genome-wide DNA methylation analysis predicts an epigenetic switch for GATA factor expression in endometriosis. PLoS Genet. 2014;10(3) doi: 10.1371/journal.pgen.1004158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Bernardi L.A., Dyson M.T., Tokunaga H., et al. The essential role of GATA6 in the activation of estrogen synthesis in endometriosis. Reprod Sci. 2019;26(1):60–69. doi: 10.1177/1933719118756751. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Monsivais D., Dyson M.T., Yin P., et al. ERβ- and prostaglandin E2-regulated pathways integrate cell proliferation via Ras-like and estrogen-regulated growth inhibitor in endometriosis. Mol Endocrinol. 2014;28(8):1304–1315. doi: 10.1210/me.2013-1421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Zhou M., Fu J., Xiao L., et al. miR-196a overexpression activates the MEK/ERK signal and represses the progesterone receptor and decidualization in eutopic endometrium from women with endometriosis. Hum Reprod. 2016;31(11):2598–2608. doi: 10.1093/humrep/dew223. [DOI] [PubMed] [Google Scholar]
  • 84.Chooniedass-Kothari S., Vincett D., Yan Y., et al. The protein encoded by the functional steroid receptor RNA activator is a new modulator of ER alpha transcriptional activity. FEBS Lett. 2010;584(6):1174–1180. doi: 10.1016/j.febslet.2010.02.024. [DOI] [PubMed] [Google Scholar]
  • 85.Bombail V., Gibson D.A., Collins F., MacPherson S., Critchley H.O.D., Saunders P.T.K. A Role for the orphan nuclear receptor estrogen-related receptor alpha in endometrial stromal cell decidualization and expression of genes implicated in energy metabolism. J Clin Endocrinol Metab. 2010;95(10):E224–E228. doi: 10.1210/jc.2010-0154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Li J., Xu C., Lee H.J., et al. A genomic and epigenomic atlas of prostate cancer in Asian populations. Nature. 2020;580(7801):93–99. doi: 10.1038/s41586-020-2135-x. [DOI] [PubMed] [Google Scholar]
  • 87.Pomerantz M.M., Qiu X., Zhu Y., et al. Prostate cancer reactivates developmental epigenomic programs during metastatic progression. Nat Genet. 2020;52(8):790–799. doi: 10.1038/s41588-020-0664-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Parolia A., Eyunni S., Verma B.K., et al. NSD2 is a requisite subunit of the AR/FOXA1 neo-enhanceosome in promoting prostate tumorigenesis. Nat Genet. 2024;56(10):2132–2143. doi: 10.1038/s41588-024-01893-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Selth L.A., Townley S.L., Bert A.G., et al. Circulating microRNAs predict biochemical recurrence in prostate cancer patients. Br J Cancer. 2013;109(3):641–650. doi: 10.1038/bjc.2013.369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Fernandes R.C., Toubia J., Townley S., et al. Post-transcriptional gene regulation by microRNA-194 promotes neuroendocrine transdifferentiation in prostate cancer. Cell Rep. 2021;34(1) doi: 10.1016/j.celrep.2020.108585. [DOI] [PubMed] [Google Scholar]
  • 91.Mahfoudhi E., Talhaoui I., Cabagnols X., et al. TET2-mediated 5-hydroxymethylcytosine induces genetic instability and mutagenesis. DNA Repair. 2016;43:78–88. doi: 10.1016/j.dnarep.2016.05.031. [DOI] [PubMed] [Google Scholar]
  • 92.Smolle M.A., Bauernhofer T., Pummer K., Calin G.A., Pichler M. Current insights into long non-coding RNAs (LncRNAs) in prostate cancer. Int J Mol Sci. 2017;18(2):473. doi: 10.3390/ijms18020473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Morgan R., da Silveira W.A., Kelly R.C., Overton I., Allott E.H., Hardiman G. Long non-coding RNAs and their potential impact on diagnosis, prognosis, and therapy in prostate cancer: Racial, ethnic, and geographical considerations. Expert Rev Mol Diagn. 2021;21(12):1257–1271. doi: 10.1080/14737159.2021.1996227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Zhang A., Zhao J.C., Kim J., et al. LncRNA HOTAIR enhances the androgen-receptor-mediated transcriptional program and drives castration-resistant prostate cancer. Cell Rep. 2015;13(1):209–221. doi: 10.1016/j.celrep.2015.08.069. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Cui Z., Ren S., Lu J., et al. The prostate cancer-up-regulated long noncoding RNA PlncRNA-1 modulates apoptosis and proliferation through reciprocal regulation of androgen receptor. Urol Oncol. 2013;31(7):1117–1123. doi: 10.1016/j.urolonc.2011.11.030. [DOI] [PubMed] [Google Scholar]
  • 96.Choi J.B., Sim D.Y., Lee H.J., et al. The microRNA-193a-5p induced ROS production and disturbed colocalization between STAT3 and androgen receptor play critical roles in cornin induced apoptosis. Phytother Res. 2024;38(2):1059–1070. doi: 10.1002/ptr.8097. [DOI] [PubMed] [Google Scholar]
  • 97.Zhang Y., Pitchiaya S., Cieślik M., et al. Analysis of the androgen receptor-regulated lncRNA landscape identifies a role for ARLNC1 in prostate cancer progression. Nat Genet. 2018;50(6):814–824. doi: 10.1038/s41588-018-0120-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Robinson D., Van Allen E.M., Wu Y.M., et al. Integrative clinical genomics of advanced prostate cancer. Cell. 2015;161(5):1215–1228. doi: 10.1016/j.cell.2015.05.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Sarwar S., Alamro A., Huq F., Alghamdi A. Insights into the role of epigenetic factors determining the estrogen response in estrogen-positive ovarian cancer and prospects of combining epi-drugs with endocrine therapy. Front Genet. 2022;13 doi: 10.3389/fgene.2022.812077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Beddows I., Fan H., Heinze K., et al. Cell state of origin impacts development of distinct endometriosis-related ovarian carcinoma histotypes. Cancer Res. 2024;84(1):26–38. doi: 10.1158/0008-5472.CAN-23-1362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Venkata P.P., Jayamohan S., He Y., et al. Pharmacological inhibition of KDM1A/LSD1 enhances estrogen receptor beta-mediated tumor suppression in ovarian cancer. Cancer Lett. 2023;575 doi: 10.1016/j.canlet.2023.216383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Cheng F.H.C., Lin H.Y., Hwang T.W., et al. E2F6 functions as a competing endogenous RNA, and transcriptional repressor, to promote ovarian cancer stemness. Cancer Sci. 2019;110(3):1085–1095. doi: 10.1111/cas.13920. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Kojima M., Sone K., Oda K., et al. The histone methyltransferase SMYD2 is a novel therapeutic target for the induction of apoptosis in ovarian clear cell carcinoma cells. Oncol Lett. 2020;20(5):153. doi: 10.3892/ol.2020.12014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Jiang Y., Trescott L., Holcomb J., et al. Structural insights into estrogen receptor α methylation by histone methyltransferase SMYD2, a cellular event implicated in estrogen signaling regulation. J Mol Biol. 2014;426(20):3413–3425. doi: 10.1016/j.jmb.2014.02.019. [DOI] [PubMed] [Google Scholar]
  • 105.Arakawa Y., Watanabe M., Inoue N., Sarumaru M., Hidaka Y., Iwatani Y. Association of polymorphisms in DNMT1, DNMT3A, DNMT3B, MTHFR and MTRR genes with global DNA methylation levels and prognosis of autoimmune thyroid disease. Clin Exp Immunol. 2012;170(2):194–201. doi: 10.1111/j.1365-2249.2012.04646.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Liu T., Sun J., Wang Z., et al. Changes in the DNA methylation and hydroxymethylation status of the intercellular adhesion molecule 1 gene promoter in thyrocytes from autoimmune thyroiditis patients. Thyroid. 2017;27(6):838–845. doi: 10.1089/thy.2016.0576. [DOI] [PubMed] [Google Scholar]
  • 107.Joseph B., Ji M., Liu D., Hou P., Xing M. Lack of mutations in the thyroid hormone receptor (TR) alpha and beta genes but frequent hypermethylation of the TRbeta gene in differentiated thyroid tumors. J Clin Endocrinol Metab. 2007;92(12):4766–4770. doi: 10.1210/jc.2007-0812. [DOI] [PubMed] [Google Scholar]
  • 108.Chen J.Q., Papp G., Szodoray P., Zeher M. The role of microRNAs in the pathogenesis of autoimmune diseases. Autoimmun Rev. 2016;15(12):1171–1180. doi: 10.1016/j.autrev.2016.09.003. [DOI] [PubMed] [Google Scholar]
  • 109.Stefan M., Wei C., Lombardi A., et al. Genetic-epigenetic dysregulation of thymic TSH receptor gene expression triggers thyroid autoimmunity. Proc Natl Acad Sci USA. 2014;111(34):12562–12567. doi: 10.1073/pnas.1408821111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Ling C., Del Guerra S., Lupi R., et al. Epigenetic regulation of PPARGC1A in human type 2 diabetic islets and effect on insulin secretion. Diabetologia. 2008;51(4):615–622. doi: 10.1007/s00125-007-0916-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Tateishi K., Okada Y., Kallin E.M., Zhang Y. Role of Jhdm2a in regulating metabolic gene expression and obesity resistance. Nature. 2009;458(7239):757–761. doi: 10.1038/nature07777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Pilon C., Rebellato A., Urbanet R., et al. Methylation status of vitamin D receptor gene promoter in benign and malignant adrenal tumors. Internet J Endocrinol. 2015;2015 doi: 10.1155/2015/375349. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Wei Z., Yoshihara E., He N., et al. Vitamin D switches BAF complexes to protect β cells. Cell. 2018;173(5):1135–1149. doi: 10.1016/j.cell.2018.04.013. e15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Zhong H., Xiao J., Chen W., et al. DNA methylation of pituitary growth hormone is involved in male growth superiority of Nile tilapia (Oreochromis niloticus) Comp Biochem Physiol B Biochem Mol Biol. 2014;171:42–48. doi: 10.1016/j.cbpb.2014.03.006. [DOI] [PubMed] [Google Scholar]
  • 115.Bartlett A.A., Lapp H.E., Hunter R.G. Epigenetic mechanisms of the glucocorticoid receptor. Trends Endocrinol Metabol. 2019;30(11):807–818. doi: 10.1016/j.tem.2019.07.003. [DOI] [PubMed] [Google Scholar]
  • 116.van Straten E.M., Bloks V.W., Huijkman N.C., et al. The liver X-receptor gene promoter is hypermethylated in a mouse model of prenatal protein restriction. Am J Physiol Regul Integr Comp Physiol. 2010;298(2):R275–R282. doi: 10.1152/ajpregu.00413.2009. [DOI] [PubMed] [Google Scholar]
  • 117.Tate T., Xiang T., Wobker S.E., et al. Pparg signaling controls bladder cancer subtype and immune exclusion. Nat Commun. 2021;12(1):6160. doi: 10.1038/s41467-021-26421-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Liu S., Shi J., Wang L., et al. Loss of EMP1 promotes the metastasis of human bladder cancer cells by promoting migration and conferring resistance to ferroptosis through activation of PPAR gamma signaling. Free Radic Biol Med. 2022;189:42–57. doi: 10.1016/j.freeradbiomed.2022.06.247. [DOI] [PubMed] [Google Scholar]
  • 119.Caruso J.A., Wang X., Murrow L.M., et al. Loss of PPARγ activity characterizes early protumorigenic stromal reprogramming and dictates the therapeutic window of opportunity. Proc Natl Acad Sci USA. 2023;120(42) doi: 10.1073/pnas.2303774120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Plumber S.A., Tate T., Al-Ahmadie H., et al. Rosiglitazone and trametinib exhibit potent anti-tumor activity in a mouse model of muscle invasive bladder cancer. Nat Commun. 2024;15(1):6538. doi: 10.1038/s41467-024-50678-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Goldstein J.T., Berger A.C., Shih J., et al. Genomic activation of PPARG reveals a candidate therapeutic axis in bladder cancer. Cancer Res. 2017;77(24):6987–6998. doi: 10.1158/0008-5472.CAN-17-1701. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Strillacci A., Sansone P., Rajasekhar V.K., et al. ERα-LBD, an isoform of estrogen receptor alpha, promotes breast cancer proliferation and endocrine resistance. NPJ Breast Cancer. 2022;8(1):96. doi: 10.1038/s41523-022-00470-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Liu H., Lee E.S., Deb Los Reyes A., Zapf J.W., Jordan V.C. Silencing and reactivation of the selective estrogen receptor modulator-estrogen receptor alpha complex. Cancer Res. 2001;61(9):3632–3639. [PubMed] [Google Scholar]
  • 124.Osborne C.K., Bardou V., Hopp T.A., et al. Role of the estrogen receptor coactivator AIB1 (SRC-3) and HER-2/neu in tamoxifen resistance in breast cancer. J Natl Cancer Inst. 2003;95(5):353–361. doi: 10.1093/jnci/95.5.353. [DOI] [PubMed] [Google Scholar]
  • 125.Verret B., Cortes J., Bachelot T., Andre F., Arnedos M. Efficacy of PI3K inhibitors in advanced breast cancer. Ann Oncol. 2019;30(Suppl 10):x12–x20. doi: 10.1093/annonc/mdz381. [DOI] [PubMed] [Google Scholar]
  • 126.Toska E., Castel P., Chhangawala S., et al. PI3K inhibition activates SGK1 via a feedback loop to promote chromatin-based regulation of ER-dependent gene expression. Cell Rep. 2019;27(1):294–306. doi: 10.1016/j.celrep.2019.02.111. e5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Leenhardt F., Alexandre M., Jacot W. Alpelisib for the treatment of PIK3CA-mutated, hormone receptor-positive, HER2-negative metastatic breast cancer. Expet Opin Pharmacother. 2021;22(6):667–675. doi: 10.1080/14656566.2021.1873952. [DOI] [PubMed] [Google Scholar]
  • 128.Blawski R., Vokshi B.H., Guo X., et al. Methylation of the chromatin modifier KMT2D by SMYD2 contributes to therapeutic response in hormone-dependent breast cancer. Cell Rep. 2024;43(5) doi: 10.1016/j.celrep.2024.114174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Oronsky B., Oronsky N., Knox S., Fanger G., Scicinski J. Episensitization: therapeutic tumor resensitization by epigenetic agents: a review and reassessment. Anti Cancer Agents Med Chem. 2014;14(8):1121–1127. doi: 10.2174/1871520614666140418144610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Zucchetti B., Shimada A.K., Katz A., Curigliano G. The role of histone deacetylase inhibitors in metastatic breast cancer. Breast. 2019;43:130–134. doi: 10.1016/j.breast.2018.12.001. [DOI] [PubMed] [Google Scholar]
  • 131.Sukocheva O.A., Lukina E., Friedemann M., Menschikowski M., Hagelgans A., Aliev G. The crucial role of epigenetic regulation in breast cancer anti-estrogen resistance: current findings and future perspectives. Semin Cancer Biol. 2022;82:35–59. doi: 10.1016/j.semcancer.2020.12.004. [DOI] [PubMed] [Google Scholar]
  • 132.Shiino S., Kinoshita T., Yoshida M., et al. Prognostic impact of discordance in hormone receptor status between primary and recurrent sites in patients with recurrent breast cancer. Clin Breast Cancer. 2016;16(4):e133–e140. doi: 10.1016/j.clbc.2016.05.014. [DOI] [PubMed] [Google Scholar]
  • 133.Achinger-Kawecka J., Stirzaker C., Portman N., et al. The potential of epigenetic therapy to target the 3D epigenome in endocrine-resistant breast cancer. Nat Struct Mol Biol. 2024;31(3):498–512. doi: 10.1038/s41594-023-01181-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Munster P.N., Thurn K.T., Thomas S., et al. A phase II study of the histone deacetylase inhibitor vorinostat combined with tamoxifen for the treatment of patients with hormone therapy-resistant breast cancer. Br J Cancer. 2011;104(12):1828–1835. doi: 10.1038/bjc.2011.156. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Liang D., Feng Y., Zandkarimi F., et al. Ferroptosis surveillance independent of GPX4 and differentially regulated by sex hormones. Cell. 2023;186(13):2748–2764. doi: 10.1016/j.cell.2023.05.003. e22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Nardone A., Qiu X., Spisak S., et al. A distinct chromatin state drives therapeutic resistance in invasive lobular breast cancer. Cancer Res. 2022;82(20):3673–3686. doi: 10.1158/0008-5472.CAN-21-3186. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Llinas-Bertran A., Bellet-Ezquerra M., Seoane J.A. Epigenetic control of cancer cell dormancy and awakening in endocrine therapy resistance. Cancer Discov. 2024;14(5):704–706. doi: 10.1158/2159-8290.CD-24-0282. [DOI] [PubMed] [Google Scholar]
  • 138.Stone A., Zotenko E., Locke W.J., et al. DNA methylation of oestrogen-regulated enhancers defines endocrine sensitivity in breast cancer. Nat Commun. 2015;6:7758. doi: 10.1038/ncomms8758. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Ma S., Tang T., Probst G., et al. Transcriptional repression of estrogen receptor alpha by YAP reveals the Hippo pathway as therapeutic target for ER+ breast cancer. Nat Commun. 2022;13(1):1061. doi: 10.1038/s41467-022-28691-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Goetz M.P., Toi M., Huober J., et al. Abemaciclib plus a nonsteroidal aromatase inhibitor as initial therapy for HR+ HER2− advanced breast cancer: Final overall survival results of MONARCH 3. Ann Oncol. 2024;35(8):718–727. doi: 10.1016/j.annonc.2024.04.013. [DOI] [PubMed] [Google Scholar]
  • 141.Turner N.C., Oliveira M., Howell S.J., et al. Capivasertib in hormone receptor-positive advanced breast cancer. N Engl J Med. 2023;388(22):2058–2070. doi: 10.1056/NEJMoa2214131. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Viswanadhapalli S., Ma S., Sareddy G.R., et al. Estrogen receptor coregulator binding modulator (ERX-11) enhances the activity of CDK4/6 inhibitors against estrogen receptor-positive breast cancers. Breast Cancer Res. 2019;21(1):150. doi: 10.1186/s13058-019-1227-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Tang F., Xu D., Wang S., et al. Chromatin profiles classify castration-resistant prostate cancers suggesting therapeutic targets. Science. 2022;376(6596) doi: 10.1126/science.abe1505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Heemers H.V., Tindall D.J. Androgen receptor (AR) coregulators: a diversity of functions converging on and regulating the AR transcriptional complex. Endocr Rev. 2007;28(7):778–808. doi: 10.1210/er.2007-0019. [DOI] [PubMed] [Google Scholar]
  • 145.Rotili D., Tomassi S., Conte M., et al. Pan-histone demethylase inhibitors simultaneously targeting Jumonji C and lysine-specific demethylases display high anticancer activities. J Med Chem. 2014;57(1):42–55. doi: 10.1021/jm4012802. [DOI] [PubMed] [Google Scholar]
  • 146.Zanconato F., Battilana G., Forcato M., et al. Transcriptional addiction in cancer cells is mediated by YAP/TAZ through BRD4. Nat Med. 2018;24(10):1599–1610. doi: 10.1038/s41591-018-0158-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Bennett L., Jaiswal P.K., Harkless R.V., et al. Glucocorticoid receptor (GR) activation is associated with increased cAMP/PKA signaling in castration-resistant prostate cancer. Mol Cancer Therapeut. 2024;23(4):552–563. doi: 10.1158/1535-7163.MCT-22-0479. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Jones K., Zhang Y., Kong Y., et al. Epigenetics in prostate cancer treatment. J Transl Genet Genom. 2021;5:341–356. doi: 10.20517/jtgg.2021.19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Bai Y., Zhang Z., Cheng L., et al. Inhibition of enhancer of zeste homolog 2 (EZH2) overcomes enzalutamide resistance in castration-resistant prostate cancer. J Biol Chem. 2019;294(25):9911–9923. doi: 10.1074/jbc.RA119.008152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Yuan H., Han Y., Wang X., et al. SETD2 restricts prostate cancer metastasis by integrating EZH2 and AMPK signaling pathways. Cancer Cell. 2020;38(3):350–365. doi: 10.1016/j.ccell.2020.05.022. e7. [DOI] [PubMed] [Google Scholar]
  • 151.Dong J., Wu Z., Wang D., et al. Hsp70 binds to the androgen receptor N-terminal domain and modulates the receptor function in prostate cancer cells. Mol Cancer Therapeut. 2019;18(1):39–50. doi: 10.1158/1535-7163.MCT-18-0432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Grasso C.S., Wu Y.M., Robinson D.R., et al. The mutational landscape of lethal castration-resistant prostate cancer. Nature. 2012;487(7406):239–243. doi: 10.1038/nature11125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Xu Y., Yang Y., Wang Z., et al. ZNF397 deficiency triggers TET2-driven lineage plasticity and AR-targeted therapy resistance in prostate cancer. Cancer Discov. 2024;14(8):1496–1521. doi: 10.1158/2159-8290.CD-23-0539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Kukkonen K., Taavitsainen S., Huhtala L., et al. Chromatin and epigenetic dysregulation of prostate cancer development, progression, and therapeutic response. Cancers. 2021;13(13):3325. doi: 10.3390/cancers13133325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Arosh J.A., Lee J., Starzinski-Powitz A., Banu S.K. Selective inhibition of prostaglandin E2 receptors EP2 and EP4 modulates DNA methylation and histone modification machinery proteins in human endometriotic cells. Mol Cell Endocrinol. 2015;409:51–58. doi: 10.1016/j.mce.2015.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Bergman M.D., Schachter B.S., Karelus K., Combatsiaris E.P., Garcia T., Nelson J.F. Up-regulation of the uterine estrogen receptor and its messenger ribonucleic acid during the mouse estrous cycle: the role of estradiol. Endocrinology. 1992;130(4):1923–1930. doi: 10.1210/endo.130.4.1547720. [DOI] [PubMed] [Google Scholar]
  • 157.Wang X., Cao Q., Yu L., Shi H., Xue B., Shi H. Epigenetic regulation of macrophage polarization and inflammation by DNA methylation in obesity. JCI Insight. 2016;1(19) doi: 10.1172/jci.insight.87748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Sheng S., Margarida Bernardo M., Dzinic S.H., Chen K., Heath E.I., Sakr W.A. Tackling tumor heterogeneity and phenotypic plasticity in cancer precision medicine: our experience and a literature review. Cancer Metastasis Rev. 2018;37(4):655–663. doi: 10.1007/s10555-018-9767-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Hess-Stumpp H., Bracker T.U., Henderson D., Politz O. MS-275, a potent orally available inhibitor of histone deacetylases: the development of an anticancer agent. Int J Biochem Cell Biol. 2007;39(7–8):1388–1405. doi: 10.1016/j.biocel.2007.02.009. [DOI] [PubMed] [Google Scholar]
  • 160.Li Y., Wang L., Zhang M., et al. Advanced glycation end products inhibit the osteogenic differentiation potential of adipose-derived stem cells by modulating Wnt/β-catenin signalling pathway via DNA methylation. Cell Prolif. 2020;53(6) doi: 10.1111/cpr.12834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Dudakovic A., Evans J.M., Li Y., et al. Histone deacetylase inhibition promotes osteoblast maturation by altering the histone H4 epigenome and reduces Akt phosphorylation. J Biol Chem. 2013;288(40):28783–28791. doi: 10.1074/jbc.M113.489732. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Peng J., Ramatchandirin B., Wang Y., et al. The P300 acetyltransferase inhibitor C646 promotes membrane translocation of insulin receptor protein substrate and interaction with the insulin receptor. J Biol Chem. 2022;298(3) doi: 10.1016/j.jbc.2022.101621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Zoghbi H.Y., Beaudet A.L. Epigenetics and human disease. Cold Spring Harbor Perspect Biol. 2016;8(2):a019497. doi: 10.1101/cshperspect.a019497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Feng J., Meng X. Histone modification and histone modification-targeted anti-cancer drugs in breast cancer: fundamentals and beyond. Front Pharmacol. 2022;13 doi: 10.3389/fphar.2022.946811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Wu Y.L., Lin Z.J., Li C.C., et al. Epigenetic regulation in metabolic diseases: mechanisms and advances in clinical study. Signal Transduct Targeted Ther. 2023;8(1):98. doi: 10.1038/s41392-023-01333-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Mondal P., Natesh J., Penta D., Meeran S.M. Progress and promises of epigenetic drugs and epigenetic diets in cancer prevention and therapy: a clinical update. Semin Cancer Biol. 2022;83:503–522. doi: 10.1016/j.semcancer.2020.12.006. [DOI] [PubMed] [Google Scholar]
  • 167.Rodriguez-Cortez V.C., Hernando H., de la Rica L., Vento R., Ballestar E. Epigenomic deregulation in the immune system. Epigenomics. 2011;3(6):697–713. doi: 10.2217/epi.11.99. [DOI] [PubMed] [Google Scholar]

Articles from Genes & Diseases are provided here courtesy of Chongqing Medical University

RESOURCES