Next Article in Journal
A Two-Part Mixed Model for Differential Expression Analysis in Single-Cell High-Throughput Gene Expression Data
Next Article in Special Issue
Slow RNAPII Transcription Elongation Rate, Low Levels of RNAPII Pausing, and Elevated Histone H1 Content at Promoters Associate with Higher m6A Deposition on Nascent mRNAs
Previous Article in Journal
An Identification and Characterization of the Axolotl (Ambystoma mexicanum, Amex) Telomerase Reverse Transcriptase (Amex TERT)
Previous Article in Special Issue
Driving Chromatin Organisation through N6-methyladenosine Modification of RNA: What Do We Know and What Lies Ahead?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigenetic and Epitranscriptomic Control in Prostate Cancer

by
Judith López
1,2,†,
Ana M. Añazco-Guenkova
1,2,†,
Óscar Monteagudo-García
1,2 and
Sandra Blanco
1,2,*
1
Centro de Investigación del Cáncer and Instituto de Biología Molecular y Celular del Cáncer, Consejo Superior de Investigaciones Científicas (CSIC)—University of Salamanca, 37007 Salamanca, Spain
2
Instituto de Investigación Biomédica de Salamanca (IBSAL), Hospital Universitario de Salamanca, 37007 Salamanca, Spain
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2022, 13(2), 378; https://doi.org/10.3390/genes13020378
Submission received: 8 January 2022 / Revised: 12 February 2022 / Accepted: 16 February 2022 / Published: 18 February 2022
(This article belongs to the Special Issue Epigenomics and Epitranscriptomics Crosstalk)

Abstract

:
The initiation of prostate cancer has been long associated with DNA copy-number alterations, the loss of specific chromosomal regions and gene fusions, and driver mutations, especially those of the Androgen Receptor. Non-mutational events, particularly DNA and RNA epigenetic dysregulation, are emerging as key players in tumorigenesis. In this review we summarize the molecular changes linked to epigenetic and epitranscriptomic dysregulation in prostate cancer and the role that alterations to DNA and RNA modifications play in the initiation and progression of prostate cancer.

Graphical Abstract

1. Prostate Cancer

Prostate cancer (PCa) is the second-most diagnosed cancer in men worldwide. In 2019 it accounted for nearly one in five new diagnoses. It is the first cancer in terms of prevalence and is also a leading cause of male cancer-associated deaths [1,2]. Early detection through testing for the prostate specific antigen (PSA) and the improvement of procedures for surgical intervention radiation therapy and androgen deprivation therapy (ADT) have significantly reduced the number of deaths [3]. However, in more advanced or aggressive forms of the pathology, PCa can evolve to stages characterised by invasion of the seminal vesicles followed by metastasis especially in the bone, usually resulting in the death of the patient. The progression to metastatic disease is commonly linked to the fact that the cancer becomes androgen-independent, a frequent feature in advanced prostate cancer [1]. In fact, while ADT is initially effective in the majority of men with PCa, in around 20% of cases, patients progress to castration-resistant prostate cancer (CRPC) for which treatment options are very limited, revealing that other genetic or non-mutational factors may account for the initiation and progression of the disease [4]. Until recently, the first-line treatment options for metastatic CRPC were taxane chemotherapeutic agents [5]; unfortunately, one-third of patients fail to respond to initial treatment and, within 24 months, those who initially respond will develop resistance [6], emphasizing the need to find new therapeutic targets.
Over recent decades, a number of other genetic alterations have been identified associated with PCa. The malignancy is generally characterised by frequent androgen-regulated promoters’ fusion with members of the E26 transformation-specific (ETS) family, such as ERG. This fusion was found in 53% of tumours when the complete sequences of seven primary human prostate tumours and their paired normal counterparts were analysed [7]. DNA copy number variations (CNVs) are also frequent [8], especially loss of heterozygosity (LOH), such as a loss of chromosome 8p21 and 10q23 regions, which occurs in up to 85% of high-grade prostatic intraepithelial neoplasias (PIN) and adenocarcinomas [1]. Loss of function alterations in cell cycle genes were found also with high frequency, especially in CDKN1B (p27) and CDKN2A (p16) genes [9]. Comparative analysis in more advanced and metastatic cancers showed that APC alterations were enriched in these types of cancer, while alterations in ATM and amplifications in AR were specifically enriched in CRPC. Furthermore, other genes such as PTEN were commonly found altered in all stages of PCa [10]. Despite all these alterations, no single one was considered as the main driver of the disease.
With the advent of high-throughput next-generation sequencing, recurrent point mutations in several genes, including Androgen Receptor (AR), SPOP, TP53, FOXA1 and PTEN, are found associated with PCa incidence [11]. In recent years, through genomic profiling of prostate tumours with all clinical spectra, high frequencies of somatic and germline alterations have been found in DNA damage repair genes (DDR), phosphatidylinositol 3-kinase (PI3K), and mitogen-activated protein kinase (MAPK) signalling pathways; in genes including BRCA1, BRCA2 and ATM (DDR pathway), PTEN, PIK3CA, AKT1 (PI3K pathway), BRAF, HRAS and KRAS (MAPK pathway) [10,12], revealing possible drivers of disease initiation, metastasis and castration resistance.
This emerging era of high-throughput sequencing is revealing a complex scenario of the molecular drivers of prostate cancer, some of which are not associated with permanent DNA changes, but with epigenetic changes such as differential DNA methylation patterns [13]. These findings could improve patient outcomes through the tailored of personalised medicine, and the selection and identification of patient populations with a high risk of developing more aggressive forms of the disease. In the next section, we will comprehensively summarize the role of the epigenetic machinery in regulating the development and progression of prostate cancer, as well as recent advances in the development of epigenetic inhibitors as a promising therapeutic strategy to treat PCa, along with most relevant clinical trials involving epigenetic drugs which are summarised in Table 1.

2. Epigenetic Alterations in Prostate Cancer

Until now, profiling studies of primary PCa have been focused on the most studied alterations of this tumour type, such as AR alterations, DNA copy number and single point mutations or mRNA expression [29,30,31]. However, with the increase in large-scale genome sequencing and integrated multi-dimensional analyses projects such as The Cancer Genome Atlas (TCGA), the “Encyclopedia of DNA Elements” (ENCODE) or the International Cancer Genome Consortium (ICGC), a different picture started to emerge, where epigenetic changes can lead to chromatin remodelling and aberrant gene expression, which can have severe pathological consequences [32]. In cancer research, recent studies have developed a comprehensive profile of hundreds of primary prostate carcinomas by combining epigenetics, RNA-seq and ChIP-seq [33]. Through multiparametric genomic data integration, it was possible to uncover three subtypes of PCa with differential biological and clinical features, for a tumour type known to be difficult to classify [33]. Other studies have also established PCa subtypes based on distinct epigenetic changes. For instance, in the study by Armenia et al., the authors identified a new class of ETS-fusion-negative PCa defined by epigenetic alterations [34]. Using TCGA methylation and RNA-seq data, Xu et al. performed an epigenetic integrative analysis between normal and PCa tissue, in order to detect the pathways in which DNA methylation-driven genes were significantly enriched [35]. More recently, in the study by Lin et al., using single-cell RNA-seq profiles, the authors identified new signature genes and cell subtypes among CRPC cells [36]. All this evidence brings out a clear role for epigenetic regulation in PCa control.
Mechanistically many studies have shed light on the molecular effects underlying epigenetic dysregulation in PCa. One of the most frequent DNA methylation changes occurs at the GSTP1 promoter, a fact which was already described 20 years ago. GSTP1 modulates several signalling pathways involved in proliferation, differentiation and apoptosis [37]. After this finding, many other recurrent epigenetic alterations have been described, and may be used in the future as a biomarker for the evaluation of PCa diagnosis and prognosis. Others include the promoter CpG island hypermethylation of PTEN, which causes its silencing [38], or the hypermethylation of the tumour suppressor gene CDKN2A (which encodes p16) that leads to increased proliferation, thus contributing to carcinogenesis [39]. Even the loss of AR expression is regulated in 30% of CRPC by hypermethylation of its promoter [40]. More interestingly, recent studies have described that, in metastatic CRPC and tumours that progress to AR-independency, epigenetic principal regulators are clearly altered, as well as key factor players in chromatin biology [40].
Besides DNA methylation, other epigenetic marks regulate chromatin structure and gene expression. Among the plethora of epigenetic master regulators, we can find writers, proteins that introduce chemical modifications to DNA and/or histones; readers that identify and interpret those modifications with its domains; and erasers, whose function is to remove the modifications added by the writers [41]. Uncontrolled activity or expression can lead to tumorigenesis through different molecular events, some of which we highlight here.

2.1. Writers

DNA methylation is mainly carried out by a family of DNA methyltransferases (DNMTs). Five different enzymes exist in humans, DNMT1, DNMT3a, DNMT3b, DNMT3c, and DNMT3L. From those, DNMT3a, DNMT3b and DNMT3c directly carry out the de novo methylation, while DNMT3L performs it indirectly [42]. DNMT1 participates in the maintenance of the hemi-methylated DNA strand during replication [43]. DNA methyltransferases are frequently affected epigenetic regulators in PCa, leading to hypermethylation and the subsequent silencing of key tumour suppressor genes or oncogenes including PTEN, CDKN2A, or AR [44,45,46]. Hypermethylation of oncogenes such as YAP1 has also been associated with neuroendocrine prostate cancer (NEPC) [47]. Expression alterations of DNMT family members have been associated with altered immune infiltration patterns and biochemical recurrence [48]. In addition, the targeted activity of specific methyltransferases can promote tumorigenesis. For example, DNMT1- and DNMT3b-mediated methylation regulates RAD9 transcription, which induces tumorigenicity [49]. DNMT3a promotes epithelial-to-mesenchymal transition (EMT) by regulating the transcription of key mRNAs [50]. In sum, DNA methylation studies have revealed that DNA methylation levels are a promising approach to classify prostate cancer patients and improve diagnostic tools to predict clinical outcomes more accurately. In addition, drug development efforts to target aberrant DNA hypermethylation led to the development of the DNMT inhibitor 5’-azacitidine. The use of these DNMT inhibitors in pre-clinical studies has been shown to suppress tumour growth [51], and currently, DNMT inhibitors in combination with chemotherapy, AR inhibitors and immunotherapy are in clinical development for prostate cancer (Table 1).
Histone modifications are also common epigenetic marks. The most critical marks are catalysed by histone methyltransferases (HMT) and histone acetyltransferases (HATs). Lysine methyltransferases (KMT) are an important family involved in gene transcriptional regulation [52]. Specifically, KMT1A and KMT1E, also known as SUV39H1 and SETDB1 respectively, are upregulated in PCa cells, increasing their migration and invasion; while KMT1B (or SUV39H2) has been shown to enhance androgen-dependent activity by interacting with AR (Figure 1) [53,54]. Similarly, other epigenetic writers have been proposed as biomarkers for PCa. For instance, the lysine methyltransferase SET and MYND domain-containing protein 3 (SMYD3) is found upregulated in PCa, and its increased expression has been linked to increased cell migration and proliferation (Figure 1) [55]. Upregulated SETDB1 is associated with prognosis and is suggested to promote PCa bone metastases through the WNT pathway [56]. Members of the protein arginine methyltransferase family are also aberrantly expressed in prostate cancer cells, e.g., PRMT5, which catalyses histone arginine methylation at histone H4R3, causing epigenetic inactivation of several tumour suppressors and thus promoting prostate cancer cell growth [57].
Alterations in some epigenetic writers such as upregulation of PRMT1 and Coactivator-Associated Arginine Methyltransferase 1 (CARM1) are found in early tumours, suggesting that histone modifications and chromatin remodelling may act as epigenetic drivers at the initial stages of the disease [58]. In addition, within the nuclear receptor binding SET domain (NSD) family, probably the most relevant protein in cancer is NSD2, which was first associated with oncogenesis in tumours such as multiple myeloma [59]; in recent years, it has been found overexpressed in several solid tumours including PCa, where it is especially upregulated in metastatic stages and is correlated with recurrence (Figure 1) [60]. NSD2 catalyses the di-methylation of histone H3 at lysine 36 (H3K36me2), and thus regulating chromatin accessibility and permissive gene transcription [61]. In addition, in vitro studies have indicated that NSD2 modulates TWIST1 to promote EMT and invasiveness in prostate cancer cell lines [62]. More interestingly, NF-kB pathway genes including IL-6, IL-8 or VEGFA are transcriptionally regulated by NSD2, and thus upregulation of NSD2 in CRPC results in enhanced cell proliferation, survival and increased expression of inflammatory cytokines, which in turn induce NSD2 expression through a positive feedback loop [63]. NSD2 also interacts with the AR DNA-binding domain enhancing AR transcriptional activity [64], suggesting that NSD2 may be implicated in promoting ADT tolerance [40]. EZH2, a member of the Polycomb Repressive Complex 2 (PRC2) that regulates transcriptional silencing via histone H3 methylation at lysine 27, is elevated in PCa too [65]. Mechanistically, increased EZH2 expression favours lineage plasticity and neuroendocrine differentiation in androgen-independent tumours [66,67]. EZH2, like other histone modifiers, can also modulate AR recruitment to target sites [68], and EHZ2 inhibition, alone or in combination with AR inhibitors, has resulted in synergistic AR inhibition, resulting in the complete suppression of AR signalling [69,70]. More recently, EZH2 inhibition, by activating the double-stranded RNA-STING stress response, has been shown to increase interferon pathway activity and PD-L1 expression by tumour cells, suggesting that the combination of HMTs inhibitors with immune checkpoint inhibitors could improve the therapeutic outcome [71]. Altogether, these findings have resulted in the initiation of several clinical trials with EZH2 inhibitors, alone or in combination therapy with other targeting agents or immunotherapy (Table 1) (Figure 1).
Histone acetyltransferases (HATs) increased activity can also promote PCa by either acetylating histones or transcription factors, thus inducing transcription [72]. For example, p300 and CREB-binding protein (CBP), by acetylating AR, increase its transcriptional activity [73], and its inhibition in pre-clinical and early clinical trials has shown to downregulate the AR-dependent transcriptional program, modulate the expression levels of several biomarker in CRPC biopsies and tumour growth in both castration-sensitive and castration-resistant prostate tumours [74,75]. More recently, p300/CBP inhibition has been shown to decrease secretion of exosomal PD-L1 by tumour cells, suggesting that the combination of HAT inhibitors with immune checkpoint inhibitors could play a synergic role and improve therapeutic efficacy [76].
All this evidence indicates epigenetic writers as promising therapeutic targets, and several molecules targeting p300/CBP, PRMT5, EZH2 are being tested under clinical trials, alone or in combination with other therapies such as AR blockade or immunotherapy in advanced prostate tumours (Figure 1), which we summarize in Table 1 [16,77,78,79,80].

2.2. Readers

The bromodomain (BRD)-containing proteins, whose function is to induce transcription initiation through chromatin remodelling, are some of the most studied in prostate cancer cells among other epigenetic readers. More than 70% of NEPC and more than 50% of primary and metastatic PCas show some dysregulation in any of the bromodomain containing proteins [81]. Within this large family, the bromodomain and extraterminal (BET) proteins have been associated with prostate cancer progression. For instance, BRD4, which is the most critical, recognises acetylated lysines at enhancer regions, thus stimulating RNA polymerase II-dependent transcription by recruiting the elongation factor P-TEFb [82]. In addition, in PCa BRD4 also interacts with AR, recruiting AR to sequence-specific DNA-binding motifs to drive AR-mediated gene transcription, making AR-dependent cell lines selectively sensitive to BRD4 inhibitors [83,84]. BET bromodomain proteins have numerous AR-independent functions by activating c-Myc-dependent and c-Myc–independent transcriptional regulators upregulated in CRPC [85].
Apart from the BET family, several BRD-containing proteins have been associated with mCRPC. TRIM24 is a transcription co-regulator overexpressed in CRPC, and is essential for cancer cell proliferation [86]. BAZ2A binds to H3K14ac, repressing transcription and promoting a more aggressive form of PCa [87]. H3K4me2,3 epigenetic reader CHD1 deficiency can lead to a dramatic decrease in cell proliferation, survival, and tumorigenic potential [88]. ATAD2, BRD8, CREBBP, or KTM2A, by recognizing acetylated histones, act as superenhancers and transcriptional regulators that initiate chromatin restructuring to promote tumorigenesis. This evidence show that dysregulated BRD-containing proteins drive aberrant transcriptional programs promoting oncogenesis. This evidence has brought out the potential of targeting these chromatin remodellers, revealing that their inhibition may disrupt the dysregulated transcriptional networks with oncogenic functions. In fact, several small molecule inhibitors, blockers and proteolysis-targeting chimeras (PROTACs), which induce degradation of BET proteins through ubiquitination followed by proteolysis have been developed, and some are currently tested in clinical trials, which we summarize in Table 1 (Figure 1) [16,89,90,91,92].

2.3. Erasers

The critical erasers in PCa are DNA demethylases, histone demethylases (HDMT) and histone deacetylases (HDACs), which cause hypomethylation or deacetylation and consequently the upregulation of gene expression, leading to disease progression and higher cell invasion and metastasis [93].
To date, the only known DNA demethylases are TET enzymes, comprised of three members, namely TET1-3. DNA demethylation is a three-step process, where TETs participate in the first one, oxidizing m5C to 5-hydorxymethylation (hm5C) [94]. Decreased expression of TET enzymes has been associated with PCa. For instance, downregulation of TET2 has been implicated in the regulation of AR signalling and prostate cancer development [95,96]. TET1 is downregulated in PCa, and its depletion promotes tumour growth and metastasis in prostate xenograft models and correlates with poor survival rates [97].
Within HDMTs, lysine demethylases (KDM) play important roles in cancer. For example, lysine-specific demethylase 1 (LSD1/KDM1A) overexpression has been highly studied for its oncogenic potential. Its overexpression is associated with increased tumour progression and promotion of carcinogenesis in several cancer types, including PCa [98,99]. In PCa, higher expression of LSD1 is correlated with recurrence and poor survival in metastatic patients, and its function has been shown to be distinctive in androgen-dependent and refractory PCa (Figure 1) [98,100]. In fact, LSD1 is among the best-known modulators of AR transcriptional activity, which can either stimulate or suppress AR transcriptional activity, unveiling a dual role in PCa progression, which is common in other chromatin remodellers [98,101]. In addition, LSD1 can demethylate other key transcription factors in PCa such as FOXA1, leading to its activation and recruitment to AR-dependent enhancers [102]. Moreover, LSD1-mediated epigenetic changes can activate other key genes such as Centrosome-associated protein E (CENPE) [103], as well as cooperate with ZNF217, promoting the expression of genes involved in tumorigenesis, thus leading to CRPC promotion [104].
Many other histone demethylases have been associated to PCa, and similarly to LSD1, their gain or loss of activity seem to have a dual role (Figure 1). H3K4 demethylase KDM5D (JARID1D) is downregulated in metastatic PCa [105], and its loss is associated with resistance to docetaxel in PCa [106]. By contrast, other HMTs are upregulated and their increased activity leads to cancer progression. For instance, KDM5C (JARID1DC) is upregulated in PCa and has recently emerged as a predictive biomarker for therapy failure after prostatectomy [107]. KDM6A deletion inhibits tumour progression [108]. KDM6B (JMJD3) regulates c-Myc expression, thus promoting cell proliferation [109]. KDM5B (JARID1DB) plays a role as an AR coactivator, and is upregulated in PCa [110]. Similarly, the KDM4 (JMJD2) family of H3K9 demethylates, either alone or cooperating with LSD1, also play a significant role in modulating AR transcriptional activity, [52,111,112,113], or promoting the expression of AR spliced variants inducing resistance to ADT therapies [114]. KDM4B promotes AR-independent survival by cooperating with BMyb, and its inhibition reduces the growth of AR-independent PCa cells [115]. KDM3B has been implicated in androgen-independent CRPC, and its genetic or pharmacological suppression has shown to reduce the survival of CRPC cells versus castration-sensitive cells [116]. The histone demethylase PHD-finger protein 8 (PHF8) is upregulated in PCa, acts as a transcriptional coactivator of AR via H4K20 demethylation and promotes cancer cell proliferation, migration, invasion, and neuroendocrine differentiation [117,118]. KDM8 (JMJD5) promotes cellular proliferation and controls the activity of AR, HIF-1a, and EZH2 [119]. All this evidence shows that histone methylation dysregulation promotes oncogenesis by promoting chromatin structural changes, and suggests that targeting histone lysine demethylation is a good anti-cancer therapeutic option. In fact, recently developed LSD1 inhibitors have resulted in attenuated tumour growth, and are currently under clinical testing [104,120] (Table 1) (Figure 1).
Decreased histone deacetylase (HDACs) activity is found associated with PCa too. Contrary to HATs resulting in the increased acetylation of histones or AR, the activity of deacetylases such as SIRT1 or SIRT2 can regulate cellular growth and sensitivity to ADT therapy through AR deacetylation [121], and the use of small molecule inhibitors has been shown to effectively reduce tumour growth, EMT and metastasis, and sensitize mCRPC to ADT therapy (Figure 1) (Table 1) [23,27,122]. These studies have shown that single-agent HDAC inhibitor clinical trials have not shown significant activity. However, their combination with AR inhibitors has resulted in improved therapeutic efficacy [23]. While those results need to be further evaluated with newer AR inhibitors, they suggest that combining HADCs and AR inhibitors may be an effective strategy to re-sensitize tumours to AR inhibition.
In conclusion, all this evidence highlights that epigenetic alterations are commonly associated with PCa and, most importantly, can be used to stratify patient risk, predict clinical outcomes, and to determine best therapeutic options or to reveal molecular pathway vulnerability to distinct chemotherapeutic agents [123,124]. In addition, the oncogenic potential of altered epigenomes in cancer cells suggests that rewiring the epigenome may be an effective strategy to sensitize or kill cancer cells. In fact, several HMT/HDMTs, BET bromodomain, HDACs [40], and DNMT inhibitors are currently under clinical or pre-clinical stages in PCa [125,126], some of which have been approved by the FDA for other malignancies such as myelodysplastic syndrome (MDS), acute myeloid leukaemia (AML) or acute lymphocytic leukaemia (ALL) [127] (Table 1). Yet, further research will be needed to completely understand the extensive transcriptional networks that these epigenetic changes regulate, and how cancer cell fates may be affected. In addition, more studies are necessary to assess their clinical efficacy, to reduce secondary and adverse effects, and to optimise their use in combination with conventional chemotherapies and with predictive biomarkers to select patients that would benefit from these therapies.

3. Epitranscriptomics Alterations in Prostate Cancer

Similarly to DNA, RNA can also be modified. Despite being known for over 50 years, the study of RNA modifications has suffered a delay regarding epigenetics, probably due to the lack of suitable tools for their study [128]. Thus, the emergence of this field, known as epitranscriptomics, is closely linked to the recent refinement of tools such as mass spectrometry, next-generation sequencing [128] or cryo-electron microscopy [129], which have enabled the discovery of over 170 RNA modifications [130,131].
These modifications are found in all types of RNA, from messenger RNA (mRNA) to non-coding RNAs such as ribosomal RNA (rRNA), transfer RNA (tRNA), microRNAs (miRNAs) and long noncoding RNAs (lncRNAs) among many others [130]. tRNAs are the more extensively modified, with an average of 15% modified nucleotides per molecule and involving a large number of enzymes and a high diversity of modifications (reviewed in [132,133]). In rRNA, around 130 individual modifications can be found, with 2’-O-methylation of the ribose and pseudouridine (Ψ) being the most frequent modification (reviewed in [134]). In the case of mRNA, the most abundant internal modification is N6-Methyladenosine (m6A), with around 0.1-0.4% adenines of all mRNAs being modified [135].
In contrast with DNA modifications, which are known to mainly regulate gene expression [136], RNA modifications control many functions apart from transcription such as RNA stability, location, splicing, degradation or translation efficiency [134,137,138]. For instance, 5-methylcytosine (m5C) methylation of tRNAs stabilises their structure and protects them from nuclease-mediated cleavage [137]. However, the role or importance of most of these modifications are still unknown and, for others, it is only starting to emerge.
Despite the great diversity of modified nucleotides in RNA and the huge expansion of the field in the past years, little is known about the role of RNA modifications in PCa. In this review, we will summarize the most relevant RNA modifications found to date to have a regulatory role in prostate cancer.

3.1. m6A Deposition in RNA and Its Role in PCa

Methylation of position N6 of adenosine is the most studied modification. This modification mainly occurs in mRNA, but also in non-coding RNAs such as tRNAs, rRNA, miRNA, lncRNAs and circular RNAs (circRNAs) [139]. Despite the widespread occurrence, m6A deposition in mRNA is, by far, the best characterised. Recent improvements in high-throughput methods have demonstrated that m6A is not randomly distributed, but is specifically enriched near stop codons in 3’-untranslated regions (UTRs), within long exons, in intergenic regions and introns and at 5’-UTRs [135,139].
m6A deposition is modulated by the dynamic crosstalk between methyltransferases or “writers” and demethylases or “erasers”. In nascent pre-mRNA, m6A is deposited by a multimeric methyltransferase complex. The catalytic core of this complex is constituted by methyltransferase-like 3 (METTL3) and methyltransferase-like 14 (METTL14), which catalyses the formation of m6A and recognises RNA substrates, respectively [140]. This complex also requires several cofactors such as RNA-binding motif protein 15 (RBM15), Wilms’ tumour-associated protein (WTAP), Cbl proto-oncogene-like 1 (CBLL1, also known as HAKAI), zinc finger CCCH-type containing 13 (ZC3H13) and Vir-like m6A methyltransferase-associated (VIRMA, also known as KIAA1429) [135,140]. Recent studies have identified another m6A methyltransferase, METTL16, which acts as an independent RNA methyltransferase and mainly modifies mRNA and small nuclear RNAs (snRNAs) [141]. In addition, METTL5 has also been identified as a 18S rRNA m6A methyltransferase acting as a heterodimer together with TRMT112 [142] (Figure 2A).
Ten years ago, the discovery of two demethylases, fat mass and obesity-associated protein (FTO) and AlkB homolog 5 (ALKBH5), brought to light the reversible nature of this modification [143] and recently, another member of AlkB homolog family, ALKBH3, has been reported as a novel m6A demethylase [140]. However, the role of the m6A demethylases is still controversial. In 2017, Mauer et al. found that FTO is able to demethylate m6Am at the 5’ cap, rather than m6A, thereby decreasing mRNA stability. The disparity of this discovery initiates a discussion about m6A detection techniques and concerning the nature of its reversibility [144]. Moreover, m6A can be specifically recognised by a group of RNA binding enzymes or “readers” that recruit downstream complexes mediating the function of this modification [135]. Several readers have been reported, including: YTH domain-containing proteins (YTHDF1/2/3 and YTHDC1/2), heterogeneous nuclear ribonucleoproteins (including hnRNPC, hnRNPG and hnRNPA2B1), insulin-like growth factor 2 mRNA-binding proteins (IGF2BP1/2/3), proline-rich and coiled-coil-containing protein 2A (PRRC2A) and the fragile X mental retardation 1 (FMR1) [135] (Figure 2A).
Recent computational studies have found a complex scenario in the expression of several writers, readers and erasers in PCa [145,146,147,148,149], suggesting the use of combined risk scores of several markers for prognosis prediction [145,147,148]. However, other studies have established a clear oncogenic role for increased m6A deposition (Figure 2B). For example, METTL3 is overexpressed in PCa patients, leading to increased m6A RNA methylation [146,150,151,152,153]. In addition, several studies confirmed that METTL3 knocked-down (METTL3-KD) represses the proliferation, tumorigenic invasion, and migration capacity of PCa cells in a catalytic-dependent manner in vitro [149,150,151,152,153], and reduces tumour growth in vivo [150]. The downregulation of METTL3-mediated m6A methylation leads to the alteration of several downstream pathways. For instance, Cai et al. showed that METTL3 alters proliferation by regulating apoptosis through the Sonic Hedgehog-GLI pathway in a catalytic-dependent manner [150]. MYC proto-oncogene, which plays a key role in PCa, has also been shown to be a target of METTL3 [150,152] and their expression levels correlate in PCa patient tissues [152]. In addition, METTL3 overexpression leads to an increased expression of Integrin B1 (ITGB1) mRNA in a m6A-dependent manner, thus altering PCa cell adhesion and motility [151]. Similarly, the lncRNA NEAT-1 is methylated by METTL3 and its expression is higher in bone metastasis [154]. NEAT-1 is recruited to CYCLIN-L1 and CDK19 gene promoters through RNA–DNA interactions in a m6A-dependent manner, leading to increased proliferation and migration both in vitro and in xenografts [154]. Moreover, METTL3 downregulation affects ubiquitin-specific protease USP4 mRNA levels, leading to decreased migration and invasiveness in PCa [153]. In addition, one of the pathways by which an increase of METTL3 promotes bone metastasis is through the upregulation of the lncRNA PCAT6 via the m6A reader IGFBP2 [155]. These results, together with the increased expression of METTL3 in PCa with bone metastasis compared to primary tumours [151], suggest that METTL3-mediated methylation might play an important role in PCa progression and metastasis. More recently, studies have shown an interplay between m6A regulator expression and immune infiltration. In Zhao et al., low METTL3 and HNRNPA2B1 expression correlates with increased immune cell infiltration [156]. In Liu et al. tumours with worse prognosis show a significant decreased expression of METTL14 and ZC3H13, markedly high expression of KIAA1429 and HNRNPA2B1, and are characterised by high intratumor heterogeneity, Th2 cell infiltration, and low Th17 cell infiltration [157], highlighting the functional role of the epitranscriptome in regulating a wide range of oncogenic processes.
Other members of the methyltransferase complex have been found dysregulated in PCa. VIRMA has been found upregulated in PCa [158], and its knockdown, by regulating the abundance of the oncogenic lncRNAs CCAT1 and CCAT2, can reduce the proliferation, migration and invasion of PCa cells through MYC regulation [158]. The m6A reader YTHDF2 has been found upregulated in PCa, its expression predicts worse overall survival, and its knockdown inhibits proliferation and migration of PCa in vivo and in vitro, by binding and degrading m6A-methylated LHPP and NKX3-1 mRNAs, resulting in increased AKT activity [159]. Interestingly, increased expression of YTHDF2 has been found to be epigenetically regulated by the H3K4me3 demethylase KDM5A. By binding to the miR-495 promoter, KDM5A leads to miR-495 transcriptional inhibition and decreased expression, which results in decreased silencing of YTHDF2 mRNA [160]. By contrast, FTO has been found downregulated in PCa tissues and cell lines [161]. Similarly to the effects of the increased expression of m6A writers, downregulation of FTO leads to increased cell invasion and migration capacity by increasing m6A levels [161].
In sum, all this evidence suggests that increased m6A deposition has an oncogenic effect in prostate cancer cells and targeting METTL3 could have clinical benefits for PCa patients. With the recent emergence of METTL3 as a promising oncogenic target in several haematological malignancies and solid tumours, a great effort is being made to develop small molecule inhibitors to target m6A deposition. Two recent inhibitors have been developed that target and inhibit METTL3, STM2457 and UZH2. These two inhibitors have shown to decrease m6A methylation activity in vitro and in vivo. In addition, they have shown potent inhibitory potential in several cancer cell lines including PCa, and a promising anti-cancer effect in cell lines and pre-clinical models of acute myeloid leukaemia (AML) [162,163]. These promising results open the door to explore new therapeutic possibilities in cancer research.

3.2. m5C and hm5C Deposition in RNA and Its Role in PCa

m5C deposition in RNA is conserved in all life domains [131] and has been found most prevalently in tRNA and rRNA, but also, less frequently, in mRNA, lncRNA, vault RNA (vtRNA) and small nucleolar RNA (snoRNA) [164,165]. In mammals, there are two families of enzymes known to specifically methylate cytosine at position 5: the NOL1/NOP2/sun (NSUN) family and the DNA methyltransferase family member 2 (DNMT2) [131,165]. To date, NOP2 (also known as nucleolar antigen p120) is the only m5C methyltransferase that has been demonstrated to be associated with PCa. Increased expression of NOP2 has been considered a marker of bad prognosis for years, correlating with Gleason score, PSA serum levels and recurrence after radical prostatectomy [166,167]. Mechanistically, NOP2 catalyses the methylation of cytoplasmic 28S rRNA [168]. This methyltransferase is expressed in the late G1 and S phases of the cell cycle and is known to regulate nuclear activation associated with proliferation [166]. In addition, a recent genome-wide association study combining enhancer and eQTL mapping has reported that lower expression of NSUN4 is associated with an increased prostate cancer risk [169]. NSUN4 targets 12S mitochondrial rRNA, found on the small subunit [170], yet its potential role in PCa remains to be validated. In addition, whether other m5C RNA methyltransferases are associated with PCa is also still uncovered.
Similarly to DNA, hm5C and f5C have been found in mRNA and tRNAs in vivo [171]. ALKBH1 has been found to catalyse the conversion of m5C into f5C in tRNAs in vivo [172]. In vitro studies have shown that TET enzymes are able to oxidise m5C to hm5C in RNA [173]. Despite this finding, the enzyme that catalyses hm5C conversion in mRNA is still unknown. TET2 is found downregulated in PCa [95,174], yet whether decreased hm5C RNA levels are linked to PCa needs to be elucidated.
From a therapeutic point of view, m5C as well as other RNA modifications have been linked to the survival, proliferation and differentiation of tumour cells in several cancers [175,176,177]. Interestingly, m5C as well as other RNA modifications have been associated with drug resistance regulation [175,177,178]. For instance, combined knockdown of NSUN2 and METTL1, which catalyses tRNA 7-methylguanosine, has shown to sensitize tumour cells to the chemotherapeutic agent 5’-fluorouracil (5-FU) [178], confirming the link between RNA methylation, cell survival and chemotherapy. In addition, in skin cancer, a lack of NSUN2 specifically sensitizes tumour-initiating cells to 5-FU [177], further showing that the combination of classic chemotherapeutic agents and the inhibition of RNA modifications could become a more effective therapeutic strategy. This role of RNA modifications has not been evaluated in PCa, but aforementioned investigations linking several RNA-modifying enzymes with key PCa tumorigenic pathways will pave the way to finding novel therapeutic targets.

3.3. Pseudouridine in RNA and Its Role in PCa

Pseudouridine (Ψ), the C5-glycoside isomer of uridine, is among the most well-known RNA modifications. It comprises about 5% of the total of all RNA nucleotides and is found in almost all types of non-coding and coding RNAs [179]. Ψ biosynthesis can be carried out both by stand-alone enzymes and by RNA-guided ribonucleoprotein (RNP) complexes. To date, eleven pseudouridine synthases have been identified in humans; PUS1, TRUB2, PUS3, PUS4, RPUSD1, RPUSD2, RPUSD3, RPUSD4, PUS7, PUS7L, and PUS10 [180]. RNP complexes are constituted by the enzymatic core components dyskerin (DKC1), which carries the enzymatic activity, and three additional core proteins: nucleolar protein 10 (NOP10), glycine-arginine-rich protein 1 (GAR1) and non-histone protein 2 (NHP2); in addition, non-coding RNAs called H/ACA box snoRNAs that guide the protein complex to the specific uridines to be modified and NAF1 and SHQ1, which are H/ACA-specific chaperones [181]. This modification is irreversible and has shown to play an important role in gene expression regulation and maintenance of structural stability [135]. Similarly, the fate of Ψ-targeted RNAs is dependent on their readers. Recent studies have identified MetRS as a potential Ψ reader [182].
Aberrant deposition of Ψ, dysregulated expression of pseudouridylases and somatic mutations have been linked to PCa. For instance, in an integrative genomic profiling study of prostate cancer, somatic mutations in the chromosomal 3p were identified as tumour suppressors; most importantly, SHQ1 was the only gene in the 3p region that harboured tumour associated mutations [7,30], and loss-of-function studies confirmed a tumour growth-suppressive role of SHQ1 [183]. DKC1 is overexpressed in PCa and its knockdown decreases cell proliferation in prostate adenocarcinoma cell lines [184]. While, mechanistically, it is still unknown whether increased DKC1 expression correlates with Ψ increased levels, some studies have shown that targeting its binding site to the RNA component of Telomerase (hTR) can modulate telomerase activity [185]. However, Ψ levels have been found to be increased in the rRNA of PCa cell lines compared to prostate epithelial cells, as well as in PCa tissue compared to normal adjacent tissues [186], suggesting an oncogenic role for elevated Ψ levels. Computational analysis has also shown the upregulation of several H/ACA snoRNAs and amplification of DKC1 in metastatic tumours compared to primary tumours [186,187,188,189]. Furthermore, the studies showed that the silencing of those snoRNA in PCa cell lines reduced their proliferation and metastatic potential [189,190].
Other functional studies have highlighted an oncogenic role for increased Ψ deposition. For example, the stand-alone pseudouridine synthase PUS10 has shown to play a key role in TRAIL-induced apoptosis, participating in the release of cytochrome C and SMAC from the mitochondria in PCa cell lines [191]. Moreover, increased abundance of PUS1 was recently found to be associated with an increased risk of relapse [192]. Furthermore, metabolomic analysis of urine samples has found increased levels of Ψ in PCa patients, suggesting Ψ levels in liquid non-invasive biopsies could be a novel predictive biomarker to use in combination with others, such as PSA levels, in order to improve diagnosis [193]. Taken together, the evidence so far shows that Ψ increased deposition has an oncogenic potential, yet the molecular mechanisms underlying this effect still need to be fully investigated to establish whether targeting Ψ writers may have clinical benefits for PCa patients. Nonetheless, Ψ arises as a promising minimally invasive biomarker for PCa detection, but further studies and standardization are needed in order to obtain a fast, cost-effective detection method for the clinical practice.

3.4. RNA Editing in PCa

RNA editing is also a very prevalent post-transcriptional event in human transcriptomes. The most common type of RNA editing results from the conversion of adenosine (A) to inosine (I) in double-stranded RNA, a process catalysed by the adenosine deaminase acting on the RNA (ADAR) family of enzymes. A-to-I editing is highly prevalent within Alu elements, as well as introns, untranslated regions (UTRs) and coding transcripts [194]. In a comprehensive study to identify edited RNAs in prostate cancer patients, 16 paired DNA–RNA sequence libraries from prostate tumour specimens were analysed. Over a hundred thousand putative RNA editing events were found on introns and UTRs, and coding regions predicted to result in deleterious amino acid alterations [195]. In a later transcriptome-wide study, elevated RNA-editing and expression of ADAR enzymes were found across multiple cancer tissues, including the prostate [196]. Moreover, rare germline heterozygous variants in ADAR predispose to prostate cancer [197]. Despite these findings, little is known on the mechanistic link between elevated RNA editing levels and PCa, and few have explored the molecular causes. For instance, RNA editing of AR mRNA has been found in PCa cell lines, which leads to mutated AR with a gain-of-function activity, suggesting a contribution to hormone-refractory phenotypes [198]. Another study implicates an inflammation-driven malignant transformation due to increased type I interferon expression [197].
Taken together, these studies disclose an important regulatory mechanism in cancer for RNA editing; however, a systematic effort to define the putative roles of edited RNAs in PCa tumorigenesis and progression remains to be established.

3.5. 2′-O-methylation in PCa

One of the most abundant ribonucleotides in rRNA is 2′-O-methylation (Nm, ribomethylation), with up to 112 different positions in human ribosomes. Similarly to Ψ, its deposition is carried out by RNA-guided ribonucleoprotein (RNP) complexes formed by the core proteins NOP56, NOP58, SNU13 and fibrillarin (FBL) and the guiding box C/D snoRNAs [199]. Altered expression or mutations in the components of the 2′-O-methylation machinery have been found to be associated with several cancer types, yet in PCa, only expression changes in snoRNA U50 have been associated with a tumour suppressor function in PCa [200]. More recently, a study unveiled a methyltransferase-independent function of the oncogene EZH2 that relies on a direct interaction with fibrillarin, leading to enhanced rRNA 2′-O methylation and protein translation [201]. This study highlights an important alliance between epitranscriptomic and epigenetic pathways in tumorigenesis, and suggests that increased 2′-O methylation may have important consequences for cancer development.

4. Concluding Remarks

Despite the initial response to hormone-deprivation treatment, one of the main problems in PCa management is the relapse and progression rate to metastatic tumour, which has limited therapeutic options, none of them completely curative [40,202]. This intensifies the urgent need for the investigation of new therapeutic approaches.
Recent evidence has highlighted that epigenetic alterations are emerging as potential biomarkers to stratify PCa patients and predict clinical outcomes [40]. Epigenetic alterations are most common in advanced PCa, being especially dysregulated in metastatic CRPC [13]. These findings suggest an important role of epigenetic regulation in advanced phases of the disease and indicate that epigenetic mechanisms may regulate tumour selective pressures. The use of epigenetic modulators has been growing in recent years, and currently, six epigenetic drugs are approved by FDA for cancer treatment, mainly for haematological malignancies [127]. Regarding PCa, despite the huge number of studies pointing to epigenetic modulators as prognostic markers, none of them are used nowadays in clinical practice. However, clinical trials have shown only mild results in PCa patients, probably because most of them have been undertaken in late-stage, heavily pre-treated patients and without considering tumour subtypes [127,203,204]. A deep understanding of the molecular mechanism underlying the epigenetic mechanism and tumour biology will allow the development of successful clinical trials and the eventual approval of epigenetics-based therapies for PCa.
As in DNA, RNA modifications are also known to regulate responses to environmental signals [137,177] suggesting that they too may regulate cancer cells’ survival of challenges occurred during tumour expansion or therapies, making them attractive therapeutic targets. However, unlike epigenetics, epitranscriptomics has not reached the clinic yet. Changes in several RNA modifications have been linked to different tumours including PCa [150,151,152,154,158,160,180,186,204], revealing their potential role as tumour biomarkers. However, their use is still limited by the lack of easy, sensitive, cost-effective and reliable high-throughput detection methods. In addition, the aberrant expression of RNA-modifying enzymes has also been reported in PCa [150,151,152,158,166], but their specific roles in regulating tumorigenesis remain to be further characterised.
Similarly to epigenetics, RNA modifications are emerging as promising therapeutic targets, and great efforts are now being made to develop small molecule inhibitors to rewire the aberrant cancer epitranscriptomes. However, targeting RNA modifications could be fairly complicated since they are linked to most aspects of RNA biology, and their alteration could involve undesirable toxic effects. Moreover, the role of RNA modifications is context-dependent and could differ between cancers or even between different cell populations [177,205]. Thus, there is still a long road ahead that will require great research efforts in order to fully understand the biology of RNA modifications and the means to effectively target them, so that ground-breaking epitranscriptomics can finally reach the clinic.

Author Contributions

J.L., A.M.A.-G. and S.B. designed and wrote the manuscript. Ó.M.-G., wrote the manuscript. J.L., A.M.A.-G., Ó.M.-G. and S.B. designed the figures. J.L. and A.M.A.-G. equally contributed to this manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the Spanish Ministry of Science and Innovation MCIN/AEI/ 10.13039/501100011033 grant PID2019-111692RB-I00. In addition, we acknowledge funding from The Scientific Foundation AECC (LABAE19040BLAN), the University of Salamanca and Fundación Memoria de Samuel Solorzano Barruso (FS/36-2020), and Consejería de Educación de la Junta de Castilla y León (CSI264P20) to S.B., A.M.A-G. was supported by USAL-Santander PhD fellowship. J.L. was supported by The Scientific Foundation AECC predoctoral fellowship (PRDSA19002LÓPE) and the Spanish Ministry of Universities predoctoral fellowship (FPU19/01190). The Instituto de Biología Molecular y Celular del Cancer is supported by the Programa de Apoyo a Planes Estratégicos de Investigación de Estructuras de Investigación de Excelencia, co-financed by the Castilla–León autonomous government and the European Regional Development Fund (CLC–2017–01).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shen, M.M.; Abate-Shen, C. Molecular genetics of prostate cancer: New prospects for old challenges. Genes Dev. 2010, 24, 1967–2000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2019. CA Cancer J. Clin. 2019, 69, 7–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Pernar, C.H.; Ebot, E.M.; Wilson, K.M.; Mucci, L.A. The Epidemiology of Prostate Cancer. Cold Spring Harb. Perspect. Med. 2018, 8, a030361. [Google Scholar] [CrossRef] [PubMed]
  4. Schatten, H. Brief Overview of Prostate Cancer Statistics, Grading, Diagnosis and Treatment Strategies. Adv. Exp. Med. Biol. 2018, 1095, 1–14. [Google Scholar] [CrossRef] [PubMed]
  5. Nunzio, C.D.E.; Presicce, F.; Giacinti, S.; Bassanelli, M.; Tubaro, A. Castration-resistance prostate cancer: What is in the pipeline? Minerva Urol. Nefrol. 2018, 70, 22–41. [Google Scholar] [CrossRef]
  6. Armstrong, C.M.; Gao, A.C. Adaptive pathways and emerging strategies overcoming treatment resistance in castration resistant prostate cancer. Asian J. Urol. 2016, 3, 185–194. [Google Scholar] [CrossRef] [Green Version]
  7. Berger, M.F.; Lawrence, M.S.; Demichelis, F.; Drier, Y.; Cibulskis, K.; Sivachenko, A.Y.; Sboner, A.; Esgueva, R.; Pflueger, D.; Sougnez, C.; et al. The genomic complexity of primary human prostate cancer. Nature 2011, 470, 214–220. [Google Scholar] [CrossRef] [Green Version]
  8. Abeshouse, A.; Ahn, J.; Akbani, R.; Ally, A.; Amin, S.; Andry, C.D.; Annala, M.; Aprikian, A.; Armenia, J.; Arora, A.; et al. The Molecular Taxonomy of Primary Prostate Cancer. Cell 2015, 163, 1011–1025. [Google Scholar] [CrossRef] [Green Version]
  9. Abate-Shen, C.; Shen, M.M. Molecular genetics of prostate cancer. Genes Dev. 2000, 14, 2410–2434. [Google Scholar] [CrossRef] [Green Version]
  10. Abida, W.; Armenia, J.; Gopalan, A.; Brennan, R.; Walsh, M.; Barron, D.; Danila, D.; Rathkopf, D.; Morris, M.; Slovin, S.; et al. Prospective Genomic Profiling of Prostate Cancer Across Disease States Reveals Germline and Somatic Alterations That May Affect Clinical Decision Making. JCO Precis. Oncol. 2017, 2017, PO.17.00029. [Google Scholar] [CrossRef]
  11. Barbieri, C.E.; Baca, S.C.; Lawrence, M.S.; Demichelis, F.; Blattner, M.; Theurillat, J.P.; White, T.A.; Stojanov, P.; Van Allen, E.; Stransky, N.; et al. Exome sequencing identifies recurrent SPOP, FOXA1 and MED12 mutations in prostate cancer. Nat. Genet. 2012, 44, 685–689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Pritchard, C.C.; Mateo, J.; Walsh, M.F.; De Sarkar, N.; Abida, W.; Beltran, H.; Garofalo, A.; Gulati, R.; Carreira, S.; Eeles, R.; et al. Inherited DNA-Repair Gene Mutations in Men with Metastatic Prostate Cancer. N. Engl. J. Med. 2016, 375, 443–453. [Google Scholar] [CrossRef] [PubMed]
  13. Friedlander, T.W.; Roy, R.; Tomlins, S.A.; Ngo, V.T.; Kobayashi, Y.; Azameera, A.; Rubin, M.A.; Pienta, K.J.; Chinnaiyan, A.; Ittmann, M.M.; et al. Common structural and epigenetic changes in the genome of castration-resistant prostate cancer. Cancer Res. 2012, 72, 616–625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Sonpavde, G.; Aparicio, A.M.; Zhan, F.; North, B.; Delaune, R.; Garbo, L.E.; Rousey, S.R.; Weinstein, R.E.; Xiao, L.; Boehm, K.A.; et al. Azacitidine favorably modulates PSA kinetics correlating with plasma DNA LINE-1 hypomethylation in men with chemonaïve castration-resistant prostate cancer. Urol. Oncol. 2011, 29, 682–689. [Google Scholar] [CrossRef] [PubMed]
  15. Singal, R.; Ramachandran, K.; Gordian, E.; Quintero, C.; Zhao, W.; Reis, I.M. Phase I/II study of azacitidine, docetaxel, and prednisone in patients with metastatic castration-resistant prostate cancer previously treated with docetaxel-based therapy. Clin. Genitourin. Cancer 2015, 13, 22–31. [Google Scholar] [CrossRef] [PubMed]
  16. Aggarwal, R.R.; Schweizer, M.T.; Nanus, D.M.; Pantuck, A.J.; Heath, E.I.; Campeau, E.; Attwell, S.; Norek, K.; Snyder, M.; Bauman, L.; et al. A Phase Ib/IIa Study of the Pan-BET Inhibitor ZEN-3694 in Combination with Enzalutamide in Patients with Metastatic Castration-resistant Prostate Cancer. Clin. Cancer Res. 2020, 26, 5338–5347. [Google Scholar] [CrossRef]
  17. Piha-Paul, S.A.; Sachdev, J.C.; Barve, M.; LoRusso, P.; Szmulewitz, R.; Patel, S.P.; Lara, P.N., Jr.; Chen, X.; Hu, B.; Freise, K.J.; et al. First-in-Human Study of Mivebresib (ABBV-075), an Oral Pan-Inhibitor of Bromodomain and Extra Terminal Proteins, in Patients with Relapsed/Refractory Solid Tumors. Clin. Cancer Res. 2019, 25, 6309–6319. [Google Scholar] [CrossRef] [Green Version]
  18. Lewin, J.; Soria, J.C.; Stathis, A.; Delord, J.P.; Peters, S.; Awada, A.; Aftimos, P.G.; Bekradda, M.; Rezai, K.; Zeng, Z.; et al. Phase Ib Trial With Birabresib, a Small-Molecule Inhibitor of Bromodomain and Extraterminal Proteins, in Patients With Selected Advanced Solid Tumors. J. Clin. Oncol. 2018, 36, 3007–3014. [Google Scholar] [CrossRef]
  19. Bradley, D.; Rathkopf, D.; Dunn, R.; Stadler, W.M.; Liu, G.; Smith, D.C.; Pili, R.; Zwiebel, J.; Scher, H.; Hussain, M. Vorinostat in advanced prostate cancer patients progressing on prior chemotherapy (National Cancer Institute Trial 6862): Trial results and interleukin-6 analysis: A study by the Department of Defense Prostate Cancer Clinical Trial Consortium and University of Chicago Phase 2 Consortium. Cancer 2009, 115, 5541–5549. [Google Scholar] [CrossRef] [Green Version]
  20. Schneider, B.J.; Kalemkerian, G.P.; Bradley, D.; Smith, D.C.; Egorin, M.J.; Daignault, S.; Dunn, R.; Hussain, M. Phase I study of vorinostat (suberoylanilide hydroxamic acid, NSC 701852) in combination with docetaxel in patients with advanced and relapsed solid malignancies. Investig. New Drugs 2012, 30, 249–257. [Google Scholar] [CrossRef]
  21. Wheler, J.J.; Janku, F.; Falchook, G.S.; Jackson, T.L.; Fu, S.; Naing, A.; Tsimberidou, A.M.; Moulder, S.L.; Hong, D.S.; Yang, H.; et al. Phase I study of anti-VEGF monoclonal antibody bevacizumab and histone deacetylase inhibitor valproic acid in patients with advanced cancers. Cancer Chemother. Pharmacol. 2014, 73, 495–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Rathkopf, D.E.; Picus, J.; Hussain, A.; Ellard, S.; Chi, K.N.; Nydam, T.; Allen-Freda, E.; Mishra, K.K.; Porro, M.G.; Scher, H.I.; et al. A phase 2 study of intravenous panobinostat in patients with castration-resistant prostate cancer. Cancer Chemother. Pharmacol. 2013, 72, 537–544. [Google Scholar] [CrossRef] [Green Version]
  23. Ferrari, A.C.; Alumkal, J.J.; Stein, M.N.; Taplin, M.E.; Babb, J.; Barnett, E.S.; Gomez-Pinillos, A.; Liu, X.; Moore, D.; DiPaola, R.; et al. Epigenetic Therapy with Panobinostat Combined with Bicalutamide Rechallenge in Castration-Resistant Prostate Cancer. Clin. Cancer Res. 2019, 25, 52–63. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Rathkopf, D.; Wong, B.Y.; Ross, R.W.; Anand, A.; Tanaka, E.; Woo, M.M.; Hu, J.; Dzik-Jurasz, A.; Yang, W.; Scher, H.I. A phase I study of oral panobinostat alone and in combination with docetaxel in patients with castration-resistant prostate cancer. Cancer Chemother. Pharmacol. 2010, 66, 181–189. [Google Scholar] [CrossRef] [PubMed]
  25. Ferrari, A.C.; Stein, M.N.; Alumkal, J.J.; Gomez-Pinillos, A.; Catamero, D.D.; Mayer, T.M.; Collins, F.; Beer, T.M.; DiPaola, R.S. A phase I/II randomized study of panobinostat and bicalutamide in castration-resistant prostate cancer (CRPC) patients progressing on second-line hormone therapy. J. Clin. Oncol. 2011, 29, 156. [Google Scholar] [CrossRef]
  26. Molife, L.R.; Attard, G.; Fong, P.C.; Karavasilis, V.; Reid, A.H.; Patterson, S.; Riggs, C.E., Jr.; Higano, C.; Stadler, W.M.; McCulloch, W.; et al. Phase II, two-stage, single-arm trial of the histone deacetylase inhibitor (HDACi) romidepsin in metastatic castration-resistant prostate cancer (CRPC). Ann. Oncol. 2010, 21, 109–113. [Google Scholar] [CrossRef] [PubMed]
  27. Eigl, B.J.; North, S.; Winquist, E.; Finch, D.; Wood, L.; Sridhar, S.S.; Powers, J.; Good, J.; Sharma, M.; Squire, J.A.; et al. A phase II study of the HDAC inhibitor SB939 in patients with castration resistant prostate cancer: NCIC clinical trials group study IND195. Investig. New Drugs 2015, 33, 969–976. [Google Scholar] [CrossRef]
  28. Ryan, Q.C.; Headlee, D.; Acharya, M.; Sparreboom, A.; Trepel, J.B.; Ye, J.; Figg, W.D.; Hwang, K.; Chung, E.J.; Murgo, A.; et al. Phase I and pharmacokinetic study of MS-275, a histone deacetylase inhibitor, in patients with advanced and refractory solid tumors or lymphoma. J. Clin. Oncol. 2005, 23, 3912–3922. [Google Scholar] [CrossRef]
  29. Varambally, S.; Yu, J.; Laxman, B.; Rhodes, D.R.; Mehra, R.; Tomlins, S.A.; Shah, R.B.; Chandran, U.; Monzon, F.A.; Becich, M.J.; et al. Integrative genomic and proteomic analysis of prostate cancer reveals signatures of metastatic progression. Cancer Cell 2005, 8, 393–406. [Google Scholar] [CrossRef] [Green Version]
  30. Taylor, B.S.; Schultz, N.; Hieronymus, H.; Gopalan, A.; Xiao, Y.; Carver, B.S.; Arora, V.K.; Kaushik, P.; Cerami, E.; Reva, B.; et al. Integrative genomic profiling of human prostate cancer. Cancer Cell 2010, 18, 11–22. [Google Scholar] [CrossRef] [Green Version]
  31. Hieronymus, H.; Schultz, N.; Gopalan, A.; Carver, B.S.; Chang, M.T.; Xiao, Y.; Heguy, A.; Huberman, K.; Bernstein, M.; Assel, M.; et al. Copy number alteration burden predicts prostate cancer relapse. Proc. Natl. Acad. Sci. USA 2014, 111, 11139–11144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Hanahan, D.; Weinberg, R.A. The hallmarks of cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef] [Green Version]
  33. Stelloo, S.; Nevedomskaya, E.; Kim, Y.; Schuurman, K.; Valle-Encinas, E.; Lobo, J.; Krijgsman, O.; Peeper, D.S.; Chang, S.L.; Feng, F.Y.; et al. Integrative epigenetic taxonomy of primary prostate cancer. Nat. Commun. 2018, 9, 4900. [Google Scholar] [CrossRef] [PubMed]
  34. Armenia, J.; Wankowicz, S.A.M.; Liu, D.; Gao, J.; Kundra, R.; Reznik, E.; Chatila, W.K.; Chakravarty, D.; Han, G.C.; Coleman, I.; et al. Publisher Correction: The long tail of oncogenic drivers in prostate cancer. Nat. Genet. 2019, 51, 1194. [Google Scholar] [CrossRef] [PubMed]
  35. Xu, N.; Wu, Y.P.; Ke, Z.B.; Liang, Y.C.; Cai, H.; Su, W.T.; Tao, X.; Chen, S.H.; Zheng, Q.S.; Wei, Y.; et al. Identification of key DNA methylation-driven genes in prostate adenocarcinoma: An integrative analysis of TCGA methylation data. J. Transl. Med. 2019, 17, 311. [Google Scholar] [CrossRef] [PubMed]
  36. Lin, X.D.; Lin, N.; Lin, T.T.; Wu, Y.P.; Huang, P.; Ke, Z.B.; Lin, Y.Z.; Chen, S.H.; Zheng, Q.S.; Wei, Y.; et al. Identification of marker genes and cell subtypes in castration-resistant prostate cancer cells. J. Cancer 2021, 12, 1249–1257. [Google Scholar] [CrossRef] [PubMed]
  37. Martignano, F.; Gurioli, G.; Salvi, S.; Calistri, D.; Costantini, M.; Gunelli, R.; De Giorgi, U.; Foca, F.; Casadio, V. GSTP1 Methylation and Protein Expression in Prostate Cancer: Diagnostic Implications. Dis. Markers 2016, 2016, 4358292. [Google Scholar] [CrossRef] [Green Version]
  38. Suzuki, H.; Freije, D.; Nusskern, D.R.; Okami, K.; Cairns, P.; Sidransky, D.; Isaacs, W.B.; Bova, G.S. Interfocal heterogeneity of PTEN/MMAC1 gene alterations in multiple metastatic prostate cancer tissues. Cancer Res. 1998, 58, 204–209. [Google Scholar]
  39. Jarrard, D.F.; Bova, G.S.; Ewing, C.M.; Pin, S.S.; Nguyen, S.H.; Baylin, S.B.; Cairns, P.; Sidransky, D.; Herman, J.G.; Isaacs, W.B. Deletional, mutational, and methylation analyses of CDKN2 (p16/MTS1) in primary and metastatic prostate cancer. Genes Chromosomes Cancer 1997, 19, 90–96. [Google Scholar] [CrossRef]
  40. Ruggero, K.; Farran-Matas, S.; Martinez-Tebar, A.; Aytes, A. Epigenetic Regulation in Prostate Cancer Progression. Curr. Mol. Biol. Rep. 2018, 4, 101–115. [Google Scholar] [CrossRef]
  41. Biswas, S.; Rao, C.M. Epigenetic tools (The Writers, The Readers and The Erasers) and their implications in cancer therapy. Eur. J. Pharmacol. 2018, 837, 8–24. [Google Scholar] [CrossRef] [PubMed]
  42. Lyko, F. The DNA methyltransferase family: A versatile toolkit for epigenetic regulation. Nat Rev Genet 2018, 19, 81–92. [Google Scholar] [CrossRef]
  43. Uysal, F.; Cinar, O.; Can, A. Knockdown of Dnmt1 and Dnmt3a gene expression disrupts preimplantation embryo development through global DNA methylation. J. Assist. Reprod. Genet. 2021, 38, 3135–3144. [Google Scholar] [CrossRef]
  44. Turnham, D.J.; Bullock, N.; Dass, M.S.; Staffurth, J.N.; Pearson, H.B. The PTEN Conundrum: How to Target PTEN-Deficient Prostate Cancer. Cells 2020, 9, 2342. [Google Scholar] [CrossRef]
  45. Zhao, R.; Choi, B.Y.; Lee, M.H.; Bode, A.M.; Dong, Z. Implications of Genetic and Epigenetic Alterations of CDKN2A (p16(INK4a)) in Cancer. EBioMedicine 2016, 8, 30–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Nakayama, T.; Watanabe, M.; Suzuki, H.; Toyota, M.; Sekita, N.; Hirokawa, Y.; Mizokami, A.; Ito, H.; Yatani, R.; Shiraishi, T. Epigenetic regulation of androgen receptor gene expression in human prostate cancers. Lab. Investig. 2000, 80, 1789–1796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Asrani, K.; Torres, A.F.C.; Woo, J.; Vidotto, T.; Tsai, H.K.; Luo, J.; Corey, E.; Hanratty, B.; Coleman, I.; Yegnasubramanian, S.; et al. Reciprocal YAP1 loss and INSM1 expression in neuroendocrine prostate cancer. J. Pathol. 2021, 255, 425–437. [Google Scholar] [CrossRef]
  48. Wang, L.; Ren, G.; Lin, B. Expression of 5-methylcytosine regulators is highly associated with the clinical phenotypes of prostate cancer and DNMTs expression predicts biochemical recurrence. Cancer Med. 2021, 10, 5681–5695. [Google Scholar] [CrossRef]
  49. Zhu, A.; Hopkins, K.M.; Friedman, R.A.; Bernstock, J.D.; Broustas, C.G.; Lieberman, H.B. DNMT1 and DNMT3B regulate tumorigenicity of human prostate cancer cells by controlling RAD9 expression through targeted methylation. Carcinogenesis 2021, 42, 220–231. [Google Scholar] [CrossRef]
  50. Mancini, M.; Grasso, M.; Muccillo, L.; Babbio, F.; Precazzini, F.; Castiglioni, I.; Zanetti, V.; Rizzo, F.; Pistore, C.; De Marino, M.G.; et al. DNMT3A epigenetically regulates key microRNAs involved in epithelial-to-mesenchymal transition in prostate cancer. Carcinogenesis 2021, 42, 1449–1460. [Google Scholar] [CrossRef]
  51. Lin, J.; Haffner, M.C.; Zhang, Y.; Lee, B.H.; Brennen, W.N.; Britton, J.; Kachhap, S.K.; Shim, J.S.; Liu, J.O.; Nelson, W.G.; et al. Disulfiram is a DNA demethylating agent and inhibits prostate cancer cell growth. Prostate 2011, 71, 333–343. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Coffey, K.; Rogerson, L.; Ryan-Munden, C.; Alkharaif, D.; Stockley, J.; Heer, R.; Sahadevan, K.; O’Neill, D.; Jones, D.; Darby, S.; et al. The lysine demethylase, KDM4B, is a key molecule in androgen receptor signalling and turnover. Nucleic Acids Res. 2013, 41, 4433–4446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Yu, T.; Wang, C.; Yang, J.; Guo, Y.; Wu, Y.; Li, X. Metformin inhibits SUV39H1-mediated migration of prostate cancer cells. Oncogenesis 2017, 6, e324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Askew, E.B.; Bai, S.; Parris, A.B.; Minges, J.T.; Wilson, E.M. Androgen receptor regulation by histone methyltransferase Suppressor of variegation 3-9 homolog 2 and Melanoma antigen-A11. Mol. Cell Endocrinol. 2017, 443, 42–51. [Google Scholar] [CrossRef] [Green Version]
  55. Huang, L.; Xu, A.M. SET and MYND domain containing protein 3 in cancer. Am. J. Transl. Res. 2017, 9, 1–14. [Google Scholar]
  56. Li, Q.; Li, Y.; Wang, Y.; Cui, Z.; Gong, L.; Qu, Z.; Zhong, Y.; Zhou, J.; Zhou, Y.; Gao, Y.; et al. Quantitative proteomic study of human prostate cancer cells with different metastatic potentials. Int. J. Oncol. 2016, 48, 1437–1446. [Google Scholar] [CrossRef] [Green Version]
  57. Stopa, N.; Krebs, J.E.; Shechter, D. The PRMT5 arginine methyltransferase: Many roles in development, cancer and beyond. Cell Mol. Life Sci. 2015, 72, 2041–2059. [Google Scholar] [CrossRef]
  58. Grypari, I.M.; Logotheti, S.; Zolota, V.; Troncoso, P.; Efstathiou, E.; Bravou, V.; Melachrinou, M.; Logothetis, C.; Tzelepi, V. The protein arginine methyltransferases (PRMTs) PRMT1 and CARM1 as candidate epigenetic drivers in prostate cancer progression. Medicine 2021, 100, e27094. [Google Scholar] [CrossRef]
  59. Keats, J.J.; Maxwell, C.A.; Taylor, B.J.; Hendzel, M.J.; Chesi, M.; Bergsagel, P.L.; Larratt, L.M.; Mant, M.J.; Reiman, T.; Belch, A.R.; et al. Overexpression of transcripts originating from the MMSET locus characterizes all t(4;14)(p16;q32)-positive multiple myeloma patients. Blood 2005, 105, 4060–4069. [Google Scholar] [CrossRef] [Green Version]
  60. Li, N.; Xue, W.; Yuan, H.; Dong, B.; Ding, Y.; Liu, Y.; Jiang, M.; Kan, S.; Sun, T.; Ren, J.; et al. AKT-mediated stabilization of histone methyltransferase WHSC1 promotes prostate cancer metastasis. J. Clin. Investig. 2017, 127, 1284–1302. [Google Scholar] [CrossRef] [Green Version]
  61. Li, Y.; Trojer, P.; Xu, C.F.; Cheung, P.; Kuo, A.; Drury, W.J., 3rd; Qiao, Q.; Neubert, T.A.; Xu, R.M.; Gozani, O.; et al. The target of the NSD family of histone lysine methyltransferases depends on the nature of the substrate. J. Biol. Chem. 2009, 284, 34283–34295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Ezponda, T.; Popovic, R.; Shah, M.Y.; Martinez-Garcia, E.; Zheng, Y.; Min, D.J.; Will, C.; Neri, A.; Kelleher, N.L.; Yu, J.; et al. The histone methyltransferase MMSET/WHSC1 activates TWIST1 to promote an epithelial-mesenchymal transition and invasive properties of prostate cancer. Oncogene 2013, 32, 2882–2890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Yang, P.; Guo, L.; Duan, Z.J.; Tepper, C.G.; Xue, L.; Chen, X.; Kung, H.J.; Gao, A.C.; Zou, J.X.; Chen, H.W. Histone methyltransferase NSD2/MMSET mediates constitutive NF-κB signaling for cancer cell proliferation, survival, and tumor growth via a feed-forward loop. Mol. Cell Biol. 2012, 32, 3121–3131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Kang, H.B.; Choi, Y.; Lee, J.M.; Choi, K.C.; Kim, H.C.; Yoo, J.Y.; Lee, Y.H.; Yoon, H.G. The histone methyltransferase, NSD2, enhances androgen receptor-mediated transcription. FEBS Lett. 2009, 583, 1880–1886. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Melling, N.; Thomsen, E.; Tsourlakis, M.C.; Kluth, M.; Hube-Magg, C.; Minner, S.; Koop, C.; Graefen, M.; Heinzer, H.; Wittmer, C.; et al. Overexpression of enhancer of zeste homolog 2 (EZH2) characterizes an aggressive subset of prostate cancers and predicts patient prognosis independently from pre- and postoperatively assessed clinicopathological parameters. Carcinogenesis 2015, 36, 1333–1340. [Google Scholar] [CrossRef] [Green Version]
  66. Ku, S.Y.; Rosario, S.; Wang, Y.; Mu, P.; Seshadri, M.; Goodrich, Z.W.; Goodrich, M.M.; Labbe, D.P.; Gomez, E.C.; Wang, J.; et al. Rb1 and Trp53 cooperate to suppress prostate cancer lineage plasticity, metastasis, and antiandrogen resistance. Science 2017, 355, 78–83. [Google Scholar] [CrossRef] [Green Version]
  67. Mu, P.; Zhang, Z.; Benelli, M.; Karthaus, W.R.; Hoover, E.; Chen, C.C.; Wongvipat, J.; Ku, S.Y.; Gao, D.; Cao, Z.; et al. SOX2 promotes lineage plasticity and antiandrogen resistance in TP53- and RB1-deficient prostate cancer. Science 2017, 355, 84–88. [Google Scholar] [CrossRef] [Green Version]
  68. Xu, K.; Wu, Z.J.; Groner, A.C.; He, H.H.; Cai, C.; Lis, R.T.; Wu, X.; Stack, E.C.; Loda, M.; Liu, T.; et al. EZH2 oncogenic activity in castration-resistant prostate cancer cells is Polycomb-independent. Science 2012, 338, 1465–1469. [Google Scholar] [CrossRef] [Green Version]
  69. Kim, J.; Lee, Y.; Lu, X.; Song, B.; Fong, K.W.; Cao, Q.; Licht, J.D.; Zhao, J.C.; Yu, J. Polycomb- and Methylation-Independent Roles of EZH2 as a Transcription Activator. Cell Rep. 2018, 25, 2808–2820.e4. [Google Scholar] [CrossRef] [Green Version]
  70. Xiao, L.; Tien, J.C.; Vo, J.; Tan, M.; Parolia, A.; Zhang, Y.; Wang, L.; Qiao, Y.; Shukla, S.; Wang, X.; et al. Epigenetic Reprogramming with Antisense Oligonucleotides Enhances the Effectiveness of Androgen Receptor Inhibition in Castration-Resistant Prostate Cancer. Cancer Res. 2018, 78, 5731–5740. [Google Scholar] [CrossRef] [Green Version]
  71. Morel, K.L.; Sheahan, A.V.; Burkhart, D.L.; Baca, S.C.; Boufaied, N.; Liu, Y.; Qiu, X.; Canadas, I.; Roehle, K.; Heckler, M.; et al. EZH2 inhibition activates a dsRNA-STING-interferon stress axis that potentiates response to PD-1 checkpoint blockade in prostate cancer. Nat. Cancer 2021, 2, 444–456. [Google Scholar] [CrossRef] [PubMed]
  72. Valdes-Mora, F.; Gould, C.M.; Colino-Sanguino, Y.; Qu, W.; Song, J.Z.; Taylor, K.M.; Buske, F.A.; Statham, A.L.; Nair, S.S.; Armstrong, N.J.; et al. Acetylated histone variant H2A.Z is involved in the activation of neo-enhancers in prostate cancer. Nat. Commun. 2017, 8, 1346. [Google Scholar] [CrossRef] [PubMed]
  73. Fu, M.; Wang, C.; Reutens, A.T.; Wang, J.; Angeletti, R.H.; Siconolfi-Baez, L.; Ogryzko, V.; Avantaggiati, M.L.; Pestell, R.G. p300 and p300/cAMP-response element-binding protein-associated factor acetylate the androgen receptor at sites governing hormone-dependent transactivation. J. Biol. Chem. 2000, 275, 20853–20860. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Jin, L.; Garcia, J.; Chan, E.; de la Cruz, C.; Segal, E.; Merchant, M.; Kharbanda, S.; Raisner, R.; Haverty, P.M.; Modrusan, Z.; et al. Therapeutic Targeting of the CBP/p300 Bromodomain Blocks the Growth of Castration-Resistant Prostate Cancer. Cancer Res. 2017, 77, 5564–5575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Lasko, L.M.; Jakob, C.G.; Edalji, R.P.; Qiu, W.; Montgomery, D.; Digiammarino, E.L.; Hansen, T.M.; Risi, R.M.; Frey, R.; Manaves, V.; et al. Discovery of a selective catalytic p300/CBP inhibitor that targets lineage-specific tumours. Nature 2017, 550, 128–132. [Google Scholar] [CrossRef] [PubMed]
  76. Liu, J.; He, D.; Cheng, L.; Huang, C.; Zhang, Y.; Rao, X.; Kong, Y.; Li, C.; Zhang, Z.; Liu, J.; et al. p300/CBP inhibition enhances the efficacy of programmed death-ligand 1 blockade treatment in prostate cancer. Oncogene 2020, 39, 3939–3951. [Google Scholar] [CrossRef]
  77. Wu, C.; Jin, X.; Yang, J.; Yang, Y.; He, Y.; Ding, L.; Pan, Y.; Chen, S.; Jiang, J.; Huang, H. Inhibition of EZH2 by chemo- and radiotherapy agents and small molecule inhibitors induces cell death in castration-resistant prostate cancer. Oncotarget 2016, 7, 3440–3452. [Google Scholar] [CrossRef]
  78. Ge, R.; Wang, Z.; Montironi, R.; Jiang, Z.; Cheng, M.; Santoni, M.; Huang, K.; Massari, F.; Lu, X.; Cimadamore, A.; et al. Epigenetic modulations and lineage plasticity in advanced prostate cancer. Ann. Oncol. 2020, 31, 470–479. [Google Scholar] [CrossRef] [Green Version]
  79. Davies, A.; Zoubeidi, A.; Selth, L.A. The epigenetic and transcriptional landscape of neuroendocrine prostate cancer. Endocr. Relat. Cancer 2020, 27, R35–R50. [Google Scholar] [CrossRef]
  80. Welti, J.; Sharp, A.; Brooks, N.; Yuan, W.; McNair, C.; Chand, S.N.; Pal, A.; Figueiredo, I.; Riisnaes, R.; Gurel, B.; et al. Targeting the p300/CBP Axis in Lethal Prostate Cancer. Cancer Discov. 2021, 11, 1118–1137. [Google Scholar] [CrossRef]
  81. Urbanucci, A.; Mills, I.G. Bromodomain-containing proteins in prostate cancer. Mol. Cell Endocrinol. 2018, 462, 31–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Devaiah, B.N.; Case-Borden, C.; Gegonne, A.; Hsu, C.H.; Chen, Q.; Meerzaman, D.; Dey, A.; Ozato, K.; Singer, D.S. BRD4 is a histone acetyltransferase that evicts nucleosomes from chromatin. Nat. Struct. Mol. Biol. 2016, 23, 540–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Asangani, I.A.; Dommeti, V.L.; Wang, X.; Malik, R.; Cieslik, M.; Yang, R.; Escara-Wilke, J.; Wilder-Romans, K.; Dhanireddy, S.; Engelke, C.; et al. Therapeutic targeting of BET bromodomain proteins in castration-resistant prostate cancer. Nature 2014, 510, 278–282. [Google Scholar] [CrossRef] [PubMed]
  84. Faivre, E.J.; Wilcox, D.; Lin, X.; Hessler, P.; Torrent, M.; He, W.; Uziel, T.; Albert, D.H.; McDaniel, K.; Kati, W.; et al. Exploitation of Castration-Resistant Prostate Cancer Transcription Factor Dependencies by the Novel BET Inhibitor ABBV-075. Mol. Cancer Res. 2017, 15, 35–44. [Google Scholar] [CrossRef] [Green Version]
  85. Coleman, D.J.; Gao, L.; King, C.J.; Schwartzman, J.; Urrutia, J.; Sehrawat, A.; Tayou, J.; Balter, A.; Burchard, J.; Chiotti, K.E.; et al. BET bromodomain inhibition blocks the function of a critical AR-independent master regulator network in lethal prostate cancer. Oncogene 2019, 38, 5658–5669. [Google Scholar] [CrossRef] [PubMed]
  86. Groner, A.C.; Cato, L.; de Tribolet-Hardy, J.; Bernasocchi, T.; Janouskova, H.; Melchers, D.; Houtman, R.; Cato, A.C.B.; Tschopp, P.; Gu, L.; et al. TRIM24 Is an Oncogenic Transcriptional Activator in Prostate Cancer. Cancer Cell 2016, 29, 846–858. [Google Scholar] [CrossRef] [Green Version]
  87. Peña-Hernández, R.; Aprigliano, R.; Carina Frommel, S.; Pietrzak, K.; Steiger, S.; Roganowicz, M.; Lerra, L.; Bizzarro, J.; Santoro, R. BAZ2A-mediated repression via H3K14ac-marked enhancers promotes prostate cancer stem cells. EMBO Rep. 2021, 22, e53014. [Google Scholar] [CrossRef]
  88. Zhao, D.; Lu, X.; Wang, G.; Lan, Z.; Liao, W.; Li, J.; Liang, X.; Chen, J.R.; Shah, S.; Shang, X.; et al. Synthetic essentiality of chromatin remodelling factor CHD1 in PTEN-deficient cancer. Nature 2017, 542, 484–488. [Google Scholar] [CrossRef] [Green Version]
  89. Jang, M.K.; Mochizuki, K.; Zhou, M.; Jeong, H.S.; Brady, J.N.; Ozato, K. The bromodomain protein Brd4 is a positive regulatory component of P-TEFb and stimulates RNA polymerase II-dependent transcription. Mol. Cell 2005, 19, 523–534. [Google Scholar] [CrossRef]
  90. Lovén, J.; Hoke, H.A.; Lin, C.Y.; Lau, A.; Orlando, D.A.; Vakoc, C.R.; Bradner, J.E.; Lee, T.I.; Young, R.A. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell 2013, 153, 320–334. [Google Scholar] [CrossRef] [Green Version]
  91. Pérez-Salvia, M.; Esteller, M. Bromodomain inhibitors and cancer therapy: From structures to applications. Epigenetics 2017, 12, 323–339. [Google Scholar] [CrossRef]
  92. Raina, K.; Lu, J.; Qian, Y.; Altieri, M.; Gordon, D.; Rossi, A.M.; Wang, J.; Chen, X.; Dong, H.; Siu, K.; et al. PROTAC-induced BET protein degradation as a therapy for castration-resistant prostate cancer. Proc. Natl. Acad. Sci. USA 2016, 113, 7124–7129. [Google Scholar] [CrossRef] [Green Version]
  93. Pakneshan, P.; Xing, R.H.; Rabbani, S.A. Methylation status of uPA promoter as a molecular mechanism regulating prostate cancer invasion and growth in vitro and in vivo. FASEB J. 2003, 17, 1081–1088. [Google Scholar] [CrossRef] [PubMed]
  94. Zhang, M.; Wang, J.; Zhang, K.; Lu, G.; Liu, Y.; Ren, K.; Wang, W.; Xin, D.; Xu, L.; Mao, H.; et al. Ten-eleven translocation 1 mediated-DNA hydroxymethylation is required for myelination and remyelination in the mouse brain. Nat. Commun. 2021, 12, 5091. [Google Scholar] [CrossRef] [PubMed]
  95. Nickerson, M.L.; Das, S.; Im, K.M.; Turan, S.; Berndt, S.I.; Li, H.; Lou, H.; Brodie, S.A.; Billaud, J.N.; Zhang, T.; et al. TET2 binds the androgen receptor and loss is associated with prostate cancer. Oncogene 2017, 36, 2172–2183. [Google Scholar] [CrossRef] [Green Version]
  96. Takayama, K.; Misawa, A.; Suzuki, T.; Takagi, K.; Hayashizaki, Y.; Fujimura, T.; Homma, Y.; Takahashi, S.; Urano, T.; Inoue, S. TET2 repression by androgen hormone regulates global hydroxymethylation status and prostate cancer progression. Nat. Commun. 2015, 6, 8219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Hsu, C.H.; Peng, K.L.; Kang, M.L.; Chen, Y.R.; Yang, Y.C.; Tsai, C.H.; Chu, C.S.; Jeng, Y.M.; Chen, Y.T.; Lin, F.M.; et al. TET1 suppresses cancer invasion by activating the tissue inhibitors of metalloproteinases. Cell Rep. 2012, 2, 568–579. [Google Scholar] [CrossRef] [Green Version]
  98. Metzger, E.; Wissmann, M.; Yin, N.; Muller, J.M.; Schneider, R.; Peters, A.H.; Gunther, T.; Buettner, R.; Schule, R. LSD1 demethylates repressive histone marks to promote androgen-receptor-dependent transcription. Nature 2005, 437, 436–439. [Google Scholar] [CrossRef]
  99. Hayami, S.; Kelly, J.D.; Cho, H.S.; Yoshimatsu, M.; Unoki, M.; Tsunoda, T.; Field, H.I.; Neal, D.E.; Yamaue, H.; Ponder, B.A.; et al. Overexpression of LSD1 contributes to human carcinogenesis through chromatin regulation in various cancers. Int. J. Cancer 2011, 128, 574–586. [Google Scholar] [CrossRef]
  100. Kahl, P.; Gullotti, L.; Heukamp, L.C.; Wolf, S.; Friedrichs, N.; Vorreuther, R.; Solleder, G.; Bastian, P.J.; Ellinger, J.; Metzger, E.; et al. Androgen receptor coactivators lysine-specific histone demethylase 1 and four and a half LIM domain protein 2 predict risk of prostate cancer recurrence. Cancer Res. 2006, 66, 11341–11347. [Google Scholar] [CrossRef] [Green Version]
  101. Cai, C.; He, H.H.; Chen, S.; Coleman, I.; Wang, H.; Fang, Z.; Chen, S.; Nelson, P.S.; Liu, X.S.; Brown, M.; et al. Androgen receptor gene expression in prostate cancer is directly suppressed by the androgen receptor through recruitment of lysine-specific demethylase 1. Cancer Cell 2011, 20, 457–471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Gao, S.; Chen, S.; Han, D.; Wang, Z.; Li, M.; Han, W.; Besschetnova, A.; Liu, M.; Zhou, F.; Barrett, D.; et al. Chromatin binding of FOXA1 is promoted by LSD1-mediated demethylation in prostate cancer. Nat. Genet. 2020, 52, 1011–1017. [Google Scholar] [CrossRef] [PubMed]
  103. Liang, Y.; Ahmed, M.; Guo, H.; Soares, F.; Hua, J.T.; Gao, S.; Lu, C.; Poon, C.; Han, W.; Langstein, J.; et al. LSD1-Mediated Epigenetic Reprogramming Drives CENPE Expression and Prostate Cancer Progression. Cancer Res. 2017, 77, 5479–5490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Sehrawat, A.; Gao, L.; Wang, Y.; Bankhead, A., 3rd; McWeeney, S.K.; King, C.J.; Schwartzman, J.; Urrutia, J.; Bisson, W.H.; Coleman, D.J.; et al. LSD1 activates a lethal prostate cancer gene network independently of its demethylase function. Proc. Natl. Acad. Sci. USA 2018, 115, E4179–E4188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Li, N.; Dhar, S.S.; Chen, T.Y.; Kan, P.Y.; Wei, Y.; Kim, J.H.; Chan, C.H.; Lin, H.K.; Hung, M.C.; Lee, M.G. JARID1D Is a Suppressor and Prognostic Marker of Prostate Cancer Invasion and Metastasis. Cancer Res. 2016, 76, 831–843. [Google Scholar] [CrossRef] [Green Version]
  106. Komura, K.; Jeong, S.H.; Hinohara, K.; Qu, F.; Wang, X.; Hiraki, M.; Azuma, H.; Lee, G.S.; Kantoff, P.W.; Sweeney, C.J. Resistance to docetaxel in prostate cancer is associated with androgen receptor activation and loss of KDM5D expression. Proc. Natl. Acad. Sci. USA 2016, 113, 6259–6264. [Google Scholar] [CrossRef] [Green Version]
  107. Stein, J.; Majores, M.; Rohde, M.; Lim, S.; Schneider, S.; Krappe, E.; Ellinger, J.; Dietel, M.; Stephan, C.; Jung, K.; et al. KDM5C is overexpressed in prostate cancer and is a prognostic marker for prostate-specific antigen-relapse following radical prostatectomy. Am. J. Pathol. 2014, 184, 2430–2437. [Google Scholar] [CrossRef]
  108. Tang, D.; He, J.; Dai, Y.; Zhou, H.; Zhang, C.; Leng, Q.; Geng, X.; Fu, D.; Jiang, H.; Sun, R.; et al. Targeting KDM6A Suppresses SREBP1c-Dependent Lipid Metabolism and Prostate Tumorigenesis. Cancer Res 2021. [Google Scholar] [CrossRef]
  109. Yildirim-Buharalioglu, G. KDM6B Regulates Prostate Cancer Cell Proliferation by Controlling c-MYC Expression. Mol. Pharmacol. 2021, 101, 106–119. [Google Scholar] [CrossRef]
  110. Xiang, Y.; Zhu, Z.; Han, G.; Ye, X.; Xu, B.; Peng, Z.; Ma, Y.; Yu, Y.; Lin, H.; Chen, A.P.; et al. JARID1B is a histone H3 lysine 4 demethylase up-regulated in prostate cancer. Proc. Natl. Acad. Sci. USA 2007, 104, 19226–19231. [Google Scholar] [CrossRef] [Green Version]
  111. Shin, S.; Janknecht, R. Activation of androgen receptor by histone demethylases JMJD2A and JMJD2D. Biochem. Biophys. Res. Commun. 2007, 359, 742–746. [Google Scholar] [CrossRef] [PubMed]
  112. Wissmann, M.; Yin, N.; Muller, J.M.; Greschik, H.; Fodor, B.D.; Jenuwein, T.; Vogler, C.; Schneider, R.; Gunther, T.; Buettner, R.; et al. Cooperative demethylation by JMJD2C and LSD1 promotes androgen receptor-dependent gene expression. Nat. Cell Biol. 2007, 9, 347–353. [Google Scholar] [CrossRef] [PubMed]
  113. Yamane, K.; Toumazou, C.; Tsukada, Y.; Erdjument-Bromage, H.; Tempst, P.; Wong, J.; Zhang, Y. JHDM2A, a JmjC-containing H3K9 demethylase, facilitates transcription activation by androgen receptor. Cell 2006, 125, 483–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Duan, L.; Chen, Z.; Lu, J.; Liang, Y.; Wang, M.; Roggero, C.M.; Zhang, Q.J.; Gao, J.; Fang, Y.; Cao, J.; et al. Histone lysine demethylase KDM4B regulates the alternative splicing of the androgen receptor in response to androgen deprivation. Nucleic Acids Res. 2019, 47, 11623–11636. [Google Scholar] [CrossRef] [PubMed]
  115. Duan, L.; Rai, G.; Roggero, C.; Zhang, Q.J.; Wei, Q.; Ma, S.H.; Zhou, Y.; Santoyo, J.; Martinez, E.D.; Xiao, G.; et al. KDM4/JMJD2 Histone Demethylase Inhibitors Block Prostate Tumor Growth by Suppressing the Expression of AR and BMYB-Regulated Genes. Chem. Biol. 2015, 22, 1185–1196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Sarac, H.; Morova, T.; Pires, E.; McCullagh, J.; Kaplan, A.; Cingoz, A.; Bagci-Onder, T.; Onder, T.; Kawamura, A.; Lack, N.A. Systematic characterization of chromatin modifying enzymes identifies KDM3B as a critical regulator in castration resistant prostate cancer. Oncogene 2020, 39, 2187–2201. [Google Scholar] [CrossRef] [Green Version]
  117. Maina, P.K.; Shao, P.; Liu, Q.; Fazli, L.; Tyler, S.; Nasir, M.; Dong, X.; Qi, H.H. c-MYC drives histone demethylase PHF8 during neuroendocrine differentiation and in castration-resistant prostate cancer. Oncotarget 2016, 7, 75585–75602. [Google Scholar] [CrossRef] [Green Version]
  118. Tong, D.; Liu, Q.; Liu, G.; Yuan, W.; Wang, L.; Guo, Y.; Lan, W.; Zhang, D.; Dong, S.; Wang, Y.; et al. The HIF/PHF8/AR axis promotes prostate cancer progression. Oncogenesis 2016, 5, e283. [Google Scholar] [CrossRef] [Green Version]
  119. Wang, H.J.; Pochampalli, M.; Wang, L.Y.; Zou, J.X.; Li, P.S.; Hsu, S.C.; Wang, B.J.; Huang, S.H.; Yang, P.; Yang, J.C.; et al. KDM8/JMJD5 as a dual coactivator of AR and PKM2 integrates AR/EZH2 network and tumor metabolism in CRPC. Oncogene 2019, 38, 17–32. [Google Scholar] [CrossRef] [Green Version]
  120. Civenni, G.; Zoppi, G.; Vazquez, R.; Shinde, D.; Paganoni, A.; Kokanovic, A.; Lee, S.H.; Ruggeri, B.; Carbone, G.M.; Catapano, C.V. Abstract 1379: INCB059872, a novel FAD-directed LSD1 Inhibitor, is active in prostate cancer models and impacts prostate cancer stem-like cells. Cancer Res. 2018, 78, 1379. [Google Scholar] [CrossRef]
  121. Lavery, D.N.; Bevan, C.L. Androgen receptor signalling in prostate cancer: The functional consequences of acetylation. J. Biomed. Biotechnol. 2011, 2011, 862125. [Google Scholar] [CrossRef] [Green Version]
  122. Ruscetti, M.; Dadashian, E.L.; Guo, W.; Quach, B.; Mulholland, D.J.; Park, J.W.; Tran, L.M.; Kobayashi, N.; Bianchi-Frias, D.; Xing, Y.; et al. HDAC inhibition impedes epithelial-mesenchymal plasticity and suppresses metastatic, castration-resistant prostate cancer. Oncogene 2016, 35, 3781–3795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Lakshminarasimhan, R.; Liang, G. The Role of DNA Methylation in Cancer. Adv. Exp. Med. Biol. 2016, 945, 151–172. [Google Scholar] [CrossRef] [PubMed]
  124. Kelly, T.K.; De Carvalho, D.D.; Jones, P.A. Epigenetic modifications as therapeutic targets. Nat. Biotechnol. 2010, 28, 1069–1078. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Juo, Y.Y.; Gong, X.J.; Mishra, A.; Cui, X.; Baylin, S.B.; Azad, N.S.; Ahuja, N. Epigenetic therapy for solid tumors: From bench science to clinical trials. Epigenomics 2015, 7, 215–235. [Google Scholar] [CrossRef] [PubMed]
  126. Yamazaki, J.; Issa, J.P. Epigenetic aspects of MDS and its molecular targeted therapy. Int. J. Hematol. 2013, 97, 175–182. [Google Scholar] [CrossRef] [Green Version]
  127. Graça, I.; Pereira-Silva, E.; Henrique, R.; Packham, G.; Crabb, S.J.; Jerónimo, C. Epigenetic modulators as therapeutic targets in prostate cancer. Clin. Epigenet. 2016, 8, 98. [Google Scholar] [CrossRef] [Green Version]
  128. Helm, M.; Motorin, Y. Detecting RNA modifications in the epitranscriptome: Predict and validate. Nat. Rev. Genet. 2017, 18, 275–291. [Google Scholar] [CrossRef]
  129. Natchiar, S.K.; Myasnikov, A.G.; Kratzat, H.; Hazemann, I.; Klaholz, B.P. Visualization of chemical modifications in the human 80S ribosome structure. Nature 2017, 551, 472–477. [Google Scholar] [CrossRef]
  130. Boccaletto, P.; Machnicka, M.A.; Purta, E.; Pitkowski, P.; Bagiski, B.; Wirecki, T.K.; de Crcy-Lagard, V.; Ross, R.; Limbach, P.A.; Kotter, A.; et al. MODOMICS: A database of RNA modification pathways. 2017 update. Nucleic Acids Res. 2017, 46, D303–D307. [Google Scholar] [CrossRef]
  131. Garcia-Vilchez, R.; Sevilla, A.; Blanco, S. Post-transcriptional regulation by cytosine-5 methylation of RNA. Biochim. Biophys. Acta Gene Regul. Mech. 2018, 1862, 240–252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Krutyhołowa, R.; Zakrzewski, K.; Glatt, S. Charging the code—tRNA modification complexes. Curr. Opin. Struct. Biol. 2019, 55, 138–146. [Google Scholar] [CrossRef] [PubMed]
  133. Pan, T. Modifications and functional genomics of human transfer RNA. Cell Res. 2018, 28, 395–404. [Google Scholar] [CrossRef] [PubMed]
  134. Sloan, K.E.; Warda, A.S.; Sharma, S.; Entian, K.-D.; Lafontaine, D.; Bohnsack, M.T. Tuning the ribosome: The influence of rRNA modification on eukaryotic ribosome biogenesis and function. RNA Biol. 2017, 14, 1–16. [Google Scholar] [CrossRef] [PubMed]
  135. Nombela, P.; Miguel-López, B.; Blanco, S. The role of m(6)A, m(5)C and Ψ RNA modifications in cancer: Novel therapeutic opportunities. Mol. Cancer 2021, 20, 18. [Google Scholar] [CrossRef] [PubMed]
  136. Bird, A.P. CpG-rich islands and the function of DNA methylation. Nature 1986, 321, 209–213. [Google Scholar] [CrossRef] [PubMed]
  137. Blanco, S.; Dietmann, S.; Flores, J.V.; Hussain, S.; Kutter, C.; Humphreys, P.; Lukk, M.; Lombard, P.; Treps, L.; Popis, M.; et al. Aberrant methylation of tRNAs links cellular stress to neuro-developmental disorders. EMBO J. 2014, 33, 2020–2039. [Google Scholar] [CrossRef]
  138. Li, Y.; Ge, Y.-Z.; Xu, L.; Xu, Z.; Dou, Q.; Jia, R. The Potential Roles of RNA N6-Methyladenosine in Urological Tumors. Front. Cell Dev. Biol. 2020, 8, 579919. [Google Scholar] [CrossRef]
  139. Dominissini, D.; Moshitch-Moshkovitz, S.; Schwartz, S.; Salmon-Divon, M.; Ungar, L.; Osenberg, S.; Cesarkas, K.; Jacob-Hirsch, J.; Amariglio, N.; Kupiec, M.; et al. Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature 2012, 485, 201–206. [Google Scholar] [CrossRef]
  140. Zhou, Z.; Lv, J.; Yu, H.; Han, J.; Yang, X.; Feng, D.; Wu, Q.; Yuan, B.; Lu, Q.; Yang, H. Mechanism of RNA modification N6-methyladenosine in human cancer. Mol. Cancer 2020, 19, 104. [Google Scholar] [CrossRef]
  141. Warda, A.S.; Kretschmer, J.; Hackert, P.; Lenz, C.; Urlaub, H.; Höbartner, C.; Sloan, K.E.; Bohnsack, M.T. Human METTL16 is a N(6)-methyladenosine (m(6)A) methyltransferase that targets pre-mRNAs and various non-coding RNAs. EMBO Rep. 2017, 18, 2004–2014. [Google Scholar] [CrossRef] [PubMed]
  142. van Tran, N.; Ernst, F.G.M.; Hawley, B.R.; Zorbas, C.; Ulryck, N.; Hackert, P.; Bohnsack, K.E.; Bohnsack, M.T.; Jaffrey, S.R.; Graille, M.; et al. The human 18S rRNA m6A methyltransferase METTL5 is stabilized by TRMT112. Nucleic Acids Res. 2019, 47, 7719–7733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Li, Z.; Weng, H.; Su, R.; Weng, X.; Zuo, Z.; Li, C.; Huang, H.; Nachtergaele, S.; Dong, L.; Hu, C.; et al. FTO Plays an Oncogenic Role in Acute Myeloid Leukemia as a N(6)-Methyladenosine RNA Demethylase. Cancer Cell. 2017, 31, 127–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Mauer, J.; Luo, X.; Blanjoie, A.; Jiao, X.; Grozhik, A.V.; Patil, D.P.; Linder, B.; Pickering, B.F.; Vasseur, J.J.; Chen, Q.; et al. Reversible methylation of m(6)Am in the 5’ cap controls mRNA stability. Nature 2017, 541, 371–375. [Google Scholar] [CrossRef] [Green Version]
  145. Xu, J.; Liu, Y.; Liu, J.; Xu, T.; Cheng, G.; Shou, Y.; Tong, J.; Liu, L.; Zhou, L.; Xiao, W.; et al. The Identification of Critical m6A RNA Methylation Regulators as Malignant Prognosis Factors in Prostate Adenocarcinoma. Front. Genet. 2020, 11, 602485. [Google Scholar] [CrossRef] [PubMed]
  146. Wu, Q.; Xie, X.; Huang, Y.; Meng, S.; Li, Y.; Wang, H.; Hu, Y. N6-methyladenosine RNA methylation regulators contribute to the progression of prostate cancer. J. Cancer 2021, 12, 682–692. [Google Scholar] [CrossRef]
  147. Ji, G.; Huang, C.; He, S.; Gong, Y.; Song, G.; Li, X.; Zhou, L. Comprehensive analysis of m6A regulators prognostic value in prostate cancer. Aging 2020, 12, 14863–14884. [Google Scholar] [CrossRef]
  148. Zhang, Q.; Luan, J.; Song, L.; Wei, X.; Xia, J.; Song, N. Malignant Evaluation and Clinical Prognostic Values of M6A RNA Methylation Regulators in Prostate Cancer. J. Cancer 2021, 12, 3575–3586. [Google Scholar] [CrossRef]
  149. Su, H.; Wang, Y.; Li, H. RNA m6A Methylation Regulators Multi-Omics Analysis in Prostate Cancer. Front. Genet. 2021, 12, 768041. [Google Scholar] [CrossRef]
  150. Cai, J.; Yang, F.; Zhan, H.; Situ, J.; Li, W.; Mao, Y.; Luo, Y. RNA m(6)A Methyltransferase METTL3 Promotes the Growth of Prostate Cancer by Regulating Hedgehog Pathway. Onco. Targets Ther. 2019, 12, 9143–9152. [Google Scholar] [CrossRef] [Green Version]
  151. Li, E.; Wei, B.; Wang, X.; Kang, R. METTL3 enhances cell adhesion through stabilizing integrin β1 mRNA via an m6A-HuR-dependent mechanism in prostatic carcinoma. Am. J. Cancer. Res. 2020, 10, 1012–1025. [Google Scholar] [PubMed]
  152. Yuan, Y.; Du, Y.; Wang, L.; Liu, X. The M6A methyltransferase METTL3 promotes the development and progression of prostate carcinoma via mediating MYC methylation. J. Cancer. 2020, 11, 3588–3595. [Google Scholar] [CrossRef] [PubMed]
  153. Chen, Y.; Pan, C.; Wang, X.; Xu, D.; Ma, Y.; Hu, J.; Chen, P.; Xiang, Z.; Rao, Q.; Han, X. Silencing of METTL3 effectively hinders invasion and metastasis of prostate cancer cells. Theranostics 2021, 11, 7640–7657. [Google Scholar] [CrossRef] [PubMed]
  154. Wen, S.; Wei, Y.; Zen, C.; Xiong, W.; Niu, Y.; Zhao, Y. Long non-coding RNA NEAT1 promotes bone metastasis of prostate cancer through N6-methyladenosine. Mol. Cancer 2020, 19, 171. [Google Scholar] [CrossRef] [PubMed]
  155. Lang, C.; Yin, C.; Lin, K.; Li, Y.; Yang, Q.; Wu, Z.; Du, H.; Ren, D.; Dai, Y.; Peng, X. m 6 A modification of lncRNA PCAT6 promotes bone metastasis in prostate cancer through IGF2BP2-mediated IGF1R mRNA stabilization. Clin. Transl. Med. 2021, 11, e426. [Google Scholar] [CrossRef]
  156. Zhao, Y.; Sun, H.; Zheng, J.; Shao, C. Analysis of RNA m 6 A methylation regulators and tumour immune cell infiltration characterization in prostate cancer. Artif. Cells Nanomed. Biotechnol. 2021, 49, 407–435. [Google Scholar] [CrossRef]
  157. Liu, Z.; Zhong, J.; Zeng, J.; Duan, X.; Lu, J.; Sun, X.; Liu, Q.; Liang, Y.; Lin, Z.; Zhong, W.; et al. Characterization of the m6A-Associated Tumor Immune Microenvironment in Prostate Cancer to Aid Immunotherapy. Front. Immunol. 2021, 12, 735170. [Google Scholar] [CrossRef]
  158. Barros-Silva, D.; Lobo, J.; Guimarães-Teixeira, C.; Carneiro, I.; Oliveira, J.; Martens-Uzunova, E.S.; Henrique, R.; Jerónimo, C. VIRMA-Dependent N6-Methyladenosine Modifications Regulate the Expression of Long Non-Coding RNAs CCAT1 and CCAT2 in Prostate Cancer. Cancers 2020, 12, 771. [Google Scholar] [CrossRef] [Green Version]
  159. Li, J.; Xie, H.; Ying, Y.; Chen, H.; Yan, H.; He, L.; Xu, M.; Xu, X.; Liang, Z.; Liu, B.; et al. YTHDF2 mediates the mRNA degradation of the tumor suppressors to induce AKT phosphorylation in N6-methyladenosine-dependent way in prostate cancer. Mol. Cancer 2020, 19, 152. [Google Scholar] [CrossRef]
  160. Du, C.; Lv, C.; Feng, Y.; Yu, S. Activation of the KDM5A/miRNA-495/YTHDF2/m6A-MOB3B axis facilitates prostate cancer progression. J. Exp. Clin. Cancer Res. 2020, 39, 223. [Google Scholar] [CrossRef]
  161. Zhu, K.; Li, Y.; Xu, Y. The FTO m6A demethylase inhibits the invasion and migration of prostate cancer cells by regulating total m6A levels. Life Sci. 2021, 271, 119180. [Google Scholar] [CrossRef] [PubMed]
  162. Dolbois, A.; Bedi, R.K.; Bochenkova, E.; Müller, A.; Moroz-Omori, E.V.; Huang, D.; Caflisch, A. 1,4,9-Triazaspiro[5.5]undecan-2-one Derivatives as Potent and Selective METTL3 Inhibitors. J. Med. Chem. 2021, 64, 12738–12760. [Google Scholar] [CrossRef] [PubMed]
  163. Yankova, E.; Blackaby, W.; Albertella, M.; Rak, J.; De Braekeleer, E.; Tsagkogeorga, G.; Pilka, E.S.; Aspris, D.; Leggate, D.; Hendrick, A.G.; et al. Small-molecule inhibition of METTL3 as a strategy against myeloid leukaemia. Nature 2021, 593, 597–601. [Google Scholar] [CrossRef] [PubMed]
  164. Squires, J.E.; Patel, H.R.; Nousch, M.; Sibbritt, T.; Humphreys, D.T.; Parker, B.J.; Suter, C.M.; Preiss, T. Widespread occurrence of 5-methylcytosine in human coding and non-coding RNA. Nucleic Acids Res. 2012, 40, 5023–5033. [Google Scholar] [CrossRef] [PubMed]
  165. Motorin, Y.; Lyko, F.; Helm, M. 5-methylcytosine in RNA: Detection, enzymatic formation and biological functions. Nucleic Acids Res. 2010, 38, 1415–1430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Bantis, A.; Giannopoulos, A.; Gonidi, M.; Liossi, A.; Aggelonidou, E.; Petrakakou, E.; Athanassiades, P.; Athanassiadou, P. Expression of p120, Ki-67 and PCNA as proliferation biomarkers in imprint smears of prostate carcinoma and their prognostic value. Cytopathology 2004, 15, 25–31. [Google Scholar] [CrossRef]
  167. Kallakury, B.V.; Sheehan, C.E.; Rhee, S.J.; Fisher, H.A.; Kaufman, R.P., Jr.; Rifkin, M.D.; Ross, J.S. The prognostic significance of proliferation-associated nucleolar protein p120 expression in prostate adenocarcinoma: A comparison with cyclins A and B1, Ki-67, proliferating cell nuclear antigen, and p34cdc2. Cancer 1999, 85, 1569–1576. [Google Scholar] [CrossRef]
  168. Sharma, S.; Yang, J.; Watzinger, P.; Kotter, P.; Entian, K.D. Yeast Nop2 and Rcm1 methylate C2870 and C2278 of the 25S rRNA, respectively. Nucleic Acids Res. 2013, 41, 9062–9076. [Google Scholar] [CrossRef] [Green Version]
  169. Kar, S.P.; Beesley, J.; Amin Al Olama, A.; Michailidou, K.; Tyrer, J.; Kote-Jarai, Z.; Lawrenson, K.; Lindstrom, S.; Ramus, S.J.; Thompson, D.J.; et al. Genome-Wide Meta-Analyses of Breast, Ovarian, and Prostate Cancer Association Studies Identify Multiple New Susceptibility Loci Shared by at Least Two Cancer Types. Cancer Discov. 2016, 6, 1052–1067. [Google Scholar] [CrossRef] [Green Version]
  170. Metodiev, M.D.; Spåhr, H.; Loguercio Polosa, P.; Meharg, C.; Becker, C.; Altmueller, J.; Habermann, B.; Larsson, N.-G.; Ruzzenente, B. NSUN4 is a dual function mitochondrial protein required for both methylation of 12S rRNA and coordination of mitoribosomal assembly. PLoS Genet. 2014, 10, e1004110. [Google Scholar] [CrossRef] [Green Version]
  171. Delatte, B.; Wang, F.; Ngoc, L.V.; Collignon, E.; Bonvin, E.; Deplus, R.; Calonne, E.; Hassabi, B.; Putmans, P.; Awe, S.; et al. RNA biochemistry. Transcriptome-wide distribution and function of RNA hydroxymethylcytosine. Science 2016, 351, 282–285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Haag, S.; Sloan, K.E.; Ranjan, N.; Warda, A.S.; Kretschmer, J.; Blessing, C.; Hübner, B.; Seikowski, J.; Dennerlein, S.; Rehling, P.; et al. NSUN3 and ABH1 modify the wobble position of mt-tRNAMet to expand codon recognition in mitochondrial translation. EMBO J. 2016, 35, 2104–2119. [Google Scholar] [CrossRef] [PubMed]
  173. Fu, L.; Guerrero, C.R.; Zhong, N.; Amato, N.J.; Liu, Y.; Liu, S.; Cai, Q.; Ji, D.; Jin, S.G.; Niedernhofer, L.J.; et al. Tet-mediated formation of 5-hydroxymethylcytosine in RNA. J. Am. Chem. Soc. 2014, 136, 11582–11585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Kamdar, S.; Isserlin, R.; Van der Kwast, T.; Zlotta, A.R.; Bader, G.D.; Fleshner, N.E.; Bapat, B. Exploring targets of TET2-mediated methylation reprogramming as potential discriminators of prostate cancer progression. Clin. Epigenet. 2019, 11, 54. [Google Scholar] [CrossRef] [Green Version]
  175. Delaunay, S.; Frye, M. RNA modifications regulating cell fate in cancer. Nat. Cell Biol. 2019, 21, 552–559. [Google Scholar] [CrossRef] [PubMed]
  176. Barbieri, I.; Tzelepis, K.; Pandolfini, L.; Shi, J.; Millán-Zambrano, G.; Robson, S.C.; Aspris, D.; Migliori, V.; Bannister, A.J.; Han, N.; et al. Promoter-bound METTL3 maintains myeloid leukaemia by m6A-dependent translation control. Nature 2017, 552, 126–131. [Google Scholar] [CrossRef]
  177. Blanco, S.; Bandiera, R.; Popis, M.; Hussain, S.; Lombard, P.; Aleksic, J.; Sajini, A.; Tanna, H.; Cortes-Garrido, R.; Gkatza, N.; et al. Stem cell function and stress response are controlled by protein synthesis. Nature 2016, 534, 335–340. [Google Scholar] [CrossRef] [Green Version]
  178. Okamoto, M.; Fujiwara, M.; Hori, M.; Okada, K.; Yazama, F.; Konishi, H.; Xiao, Y.; Qi, G.; Shimamoto, F.; Ota, T.; et al. tRNA modifying enzymes, NSUN2 and METTL1, determine sensitivity to 5-fluorouracil in HeLa cells. PLoS Genet. 2014, 10, e1004639. [Google Scholar] [CrossRef]
  179. Rintala-Dempsey, A.C.; Kothe, U. Eukaryotic stand-alone pseudouridine synthases—RNA modifying enzymes and emerging regulators of gene expression? RNA Biol. 2017, 14, 1185–1196. [Google Scholar] [CrossRef] [Green Version]
  180. Stockert, J.A.; Weil, R.; Yadav, K.K.; Kyprianou, N.; Tewari, A.K. Pseudouridine as a novel biomarker in prostate cancer. Urol. Oncol. 2021, 39, 63–71. [Google Scholar] [CrossRef]
  181. Yu, Y.T.; Meier, U.T. RNA-guided isomerization of uridine to pseudouridine–Pseudouridylation. RNA Biol. 2014, 11, 1483–1494. [Google Scholar] [CrossRef] [Green Version]
  182. Levi, O.; Arava, Y.S. Pseudouridine-mediated translation control of mRNA by methionine aminoacyl tRNA synthetase. Nucleic Acids Res. 2021, 49, 432–443. [Google Scholar] [CrossRef] [PubMed]
  183. Nallar, S.C.; Lin, L.; Srivastava, V.; Gade, P.; Hofmann, E.R.; Ahmed, H.; Reddy, S.P.; Kalvakolanu, D.V. GRIM-1, a novel growth suppressor, inhibits rRNA maturation by suppressing small nucleolar RNAs. PLoS ONE 2011, 6, e24082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Sieron, P.; Hader, C.; Hatina, J.; Engers, R.; Wlazlinski, A.; Muller, M.; Schulz, W.A. DKC1 overexpression associated with prostate cancer progression. Br. J. Cancer 2009, 101, 1410–1416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Katunaric, M.; Zamolo, G. Modulating telomerase activity in tumor patients by targeting dyskerin binding site for hTR. Med. Hypotheses 2012, 79, 319–320. [Google Scholar] [CrossRef] [PubMed]
  186. Stockert, J.A.; Gupta, A.; Herzog, B.; Yadav, S.S.; Tewari, A.K.; Yadav, K.K. Predictive value of pseudouridine in prostate cancer. Am. J. Clin. Exp. Urol. 2019, 7, 262–272. [Google Scholar]
  187. McMahon, M.; Contreras, A.; Ruggero, D. Small RNAs with big implications: New insights into H/ACA snoRNA function and their role in human disease. Wiley Interdiscip. Rev. RNA 2015, 6, 173–189. [Google Scholar] [CrossRef] [Green Version]
  188. Martens-Uzunova, E.S.; Jalava, S.E.; Dits, N.F.; van Leenders, G.J.; Moller, S.; Trapman, J.; Bangma, C.H.; Litman, T.; Visakorpi, T.; Jenster, G. Diagnostic and prognostic signatures from the small non-coding RNA transcriptome in prostate cancer. Oncogene 2012, 31, 978–991. [Google Scholar] [CrossRef] [Green Version]
  189. Crea, F.; Quagliata, L.; Michael, A.; Liu, H.H.; Frumento, P.; Azad, A.A.; Xue, H.; Pikor, L.; Watahiki, A.; Morant, R.; et al. Integrated analysis of the prostate cancer small-nucleolar transcriptome reveals SNORA55 as a driver of prostate cancer progression. Mol. Oncol. 2016, 10, 693–703. [Google Scholar] [CrossRef]
  190. Gong, J.; Li, Y.; Liu, C.J.; Xiang, Y.; Li, C.; Ye, Y.; Zhang, Z.; Hawke, D.H.; Park, P.K.; Diao, L.; et al. A Pan-cancer Analysis of the Expression and Clinical Relevance of Small Nucleolar RNAs in Human Cancer. Cell Rep. 2017, 21, 1968–1981. [Google Scholar] [CrossRef] [Green Version]
  191. Jana, S.; Hsieh, A.C.; Gupta, R. Reciprocal amplification of caspase-3 activity by nuclear export of a putative human RNA-modifying protein, PUS10 during TRAIL-induced apoptosis. Cell Death Dis. 2017, 8, e3093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Sinha, A.; Huang, V.; Livingstone, J.; Wang, J.; Fox, N.S.; Kurganovs, N.; Ignatchenko, V.; Fritsch, K.; Donmez, N.; Heisler, L.E.; et al. The Proteogenomic Landscape of Curable Prostate Cancer. Cancer Cell 2019, 35, 414–427.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Pérez-Rambla, C.; Puchades-Carrasco, L.; García-Flores, M.; Rubio-Briones, J.; López-Guerrero, J.A.; Pineda-Lucena, A. Non-invasive urinary metabolomic profiling discriminates prostate cancer from benign prostatic hyperplasia. Metabolomics 2017, 13, 52. [Google Scholar] [CrossRef] [PubMed]
  194. Christofi, T.; Zaravinos, A. RNA editing in the forefront of epitranscriptomics and human health. J. Transl. Med. 2019, 17, 319. [Google Scholar] [CrossRef]
  195. Mo, F.; Wyatt, A.W.; Sun, Y.; Brahmbhatt, S.; McConeghy, B.J.; Wu, C.; Wang, Y.; Gleave, M.E.; Volik, S.V.; Collins, C.C. Systematic identification and characterization of RNA editing in prostate tumors. PLoS ONE 2014, 9, e101431. [Google Scholar] [CrossRef] [Green Version]
  196. Paz-Yaacov, N.; Bazak, L.; Buchumenski, I.; Porath, H.T.; Danan-Gotthold, M.; Knisbacher, B.A.; Eisenberg, E.; Levanon, E.Y. Elevated RNA Editing Activity Is a Major Contributor to Transcriptomic Diversity in Tumors. Cell Rep. 2015, 13, 267–276. [Google Scholar] [CrossRef] [Green Version]
  197. Beyer, U.; Brand, F.; Martens, H.; Weder, J.; Christians, A.; Elyan, N.; Hentschel, B.; Westphal, M.; Schackert, G.; Pietsch, T.; et al. Rare ADAR and RNASEH2B variants and a type I interferon signature in glioma and prostate carcinoma risk and tumorigenesis. Acta Neuropathol. 2017, 134, 905–922. [Google Scholar] [CrossRef] [Green Version]
  198. Martinez, H.D.; Jasavala, R.J.; Hinkson, I.; Fitzgerald, L.D.; Trimmer, J.S.; Kung, H.J.; Wright, M.E. RNA editing of androgen receptor gene transcripts in prostate cancer cells. J. Biol. Chem. 2008, 283, 29938–29949. [Google Scholar] [CrossRef] [Green Version]
  199. Barros-Silva, D.; Klavert, J.; Jenster, G.; Jeronimo, C.; Lafontaine, D.L.J.; Martens-Uzunova, E.S. The role of OncoSnoRNAs and Ribosomal RNA 2’-O-methylation in Cancer. RNA Biol. 2021, 18, 61–74. [Google Scholar] [CrossRef]
  200. Pacilli, A.; Ceccarelli, C.; Trere, D.; Montanaro, L. SnoRNA U50 levels are regulated by cell proliferation and rRNA transcription. Int. J. Mol. Sci. 2013, 14, 14923–14935. [Google Scholar] [CrossRef] [Green Version]
  201. Yi, Y.; Li, Y.; Meng, Q.; Li, Q.; Li, F.; Lu, B.; Shen, J.; Fazli, L.; Zhao, D.; Li, C.; et al. A PRC2-independent function for EZH2 in regulating rRNA 2’-O methylation and IRES-dependent translation. Nat. Cell Biol. 2021, 23, 341–354. [Google Scholar] [CrossRef] [PubMed]
  202. Seruga, B.; Ocana, A.; Tannock, I.F. Drug resistance in metastatic castration-resistant prostate cancer. Nat. Rev. Clin. Oncol. 2011, 8, 12–23. [Google Scholar] [CrossRef] [PubMed]
  203. Munster, P.N.; Marchion, D.; Thomas, S.; Egorin, M.; Minton, S.; Springett, G.; Lee, J.H.; Simon, G.; Chiappori, A.; Sullivan, D.; et al. Phase I trial of vorinostat and doxorubicin in solid tumours: Histone deacetylase 2 expression as a predictive marker. Br. J. Cancer 2009, 101, 1044–1050. [Google Scholar] [CrossRef] [Green Version]
  204. Li, J.; Meng, S.; Xu, M.; Wang, S.; He, L.; Xu, X.; Wang, X.; Xie, L. Downregulation of N(6)-methyladenosine binding YTHDF2 protein mediated by miR-493-3p suppresses prostate cancer by elevating N(6)-methyladenosine levels. Oncotarget 2017, 9, 3752–3764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Ladang, A.; Rapino, F.; Heukamp, L.C.; Tharun, L.; Shostak, K.; Hermand, D.; Delaunay, S.; Klevernic, I.; Jiang, Z.; Jacques, N.; et al. Elp3 drives Wnt-dependent tumor initiation and regeneration in the intestine. J. Exp. Med. 2015, 212, 2057–2075. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Writers, readers and erasers involved in PCa. Thick arrows indicate over- (red) or under-expression (blue) of the indicated proteins in PCa. Red arrows with flat heads indicate inhibitor compounds under clinical trial against the designated proteins. Writers: KMT1A, KMT1B and KMT1E tri-methylate H3K9; SMYD3 di- and methylates H3K4; PRMT5 methylates H4R3; NSD2 di-methylates H3K36. Erasers: LSD1 demethylates H3K9 and H3K4; KDM5B mono, di- and tri-demethylates H3K4; KDM5B and KDM5C di- and tri-demethylate H3K4. Readers: BRD-containing proteins.
Figure 1. Writers, readers and erasers involved in PCa. Thick arrows indicate over- (red) or under-expression (blue) of the indicated proteins in PCa. Red arrows with flat heads indicate inhibitor compounds under clinical trial against the designated proteins. Writers: KMT1A, KMT1B and KMT1E tri-methylate H3K9; SMYD3 di- and methylates H3K4; PRMT5 methylates H4R3; NSD2 di-methylates H3K36. Erasers: LSD1 demethylates H3K9 and H3K4; KDM5B mono, di- and tri-demethylates H3K4; KDM5B and KDM5C di- and tri-demethylate H3K4. Readers: BRD-containing proteins.
Genes 13 00378 g001
Figure 2. (A) Schematic overview of writers, readers and erasers of m6A methylation. Green ellipses represent writers. Blue ellipses represent regulators of the writer complex. Red ellipses represent erasers. Yellow ellipses represent readers. Orange circles represent m6A methylation mark. (B) Relevant m6A modifiers in PCa control. Thick arrows indicate over- (red) or under-expression (blue) of the indicated proteins in PCa. Increased or decreased expression of writers, erasers or readers leads to increased m6A deposition, which in turn regulates the fate of the indicated mRNAs and lncRNAs, modulating cellular processes: proliferation, apoptosis, adhesion, migration or invasion. Red arrows with flat heads indicate METTL3 inhibitor compounds.
Figure 2. (A) Schematic overview of writers, readers and erasers of m6A methylation. Green ellipses represent writers. Blue ellipses represent regulators of the writer complex. Red ellipses represent erasers. Yellow ellipses represent readers. Orange circles represent m6A methylation mark. (B) Relevant m6A modifiers in PCa control. Thick arrows indicate over- (red) or under-expression (blue) of the indicated proteins in PCa. Increased or decreased expression of writers, erasers or readers leads to increased m6A deposition, which in turn regulates the fate of the indicated mRNAs and lncRNAs, modulating cellular processes: proliferation, apoptosis, adhesion, migration or invasion. Red arrows with flat heads indicate METTL3 inhibitor compounds.
Genes 13 00378 g002
Table 1. Epigenetic inhibitors in clinical trials for PCa.
Table 1. Epigenetic inhibitors in clinical trials for PCa.
DRUGTRAIL IDPHASEPROTOCOLSTATUS
DNMT INHIBITORS
5-AZACYTIDINENCT00384839Phase IIPatients with CRPC received 75 mg/m2 of 5-azacytidine for five consecutive days of a 28-day cycle. Patients were treated until clinical progression up to a maximum of 12 cycles.Completed. 5-Azacytidine modulates PSA (doubling time > 3 months) in 56% of patients. Clinical progression-free survival of 12.4 weeks [14]
5-AZACYTIDINENCT00503984Phase I/IImCRPC (+docetaxel, prednisone)Completed [15].
5-AZACYTIDINENCT00006019Phase IImCRPC (+ Sodium phenylbutyrate)Completed.
DISULFIRAMNCT01118741 CRPCCompleted.
DECITABINENCT03709550Phase I/IImCRPC (+Enzalutamide)Not yet recruiting.
5-AZACYTIDINENCT02959437Phase I/IIAdvanced Solid tumours (+ PD-1 + IDO-1)Terminated by Sponsor
HMT INHIBITORS
PRMT5 INHIBITOR MAK683NCT02900651Phase I/IIDiffuse large B cell lymphoma,
advanced solid tumours
Recruiting
EZH2 INHIBITOR TAZEMETOSTATNCT03213665Phase IIAdvanced solid tumoursActive. Not Recruiting
EZH2 INHIBITOR
CPI-1205
NCT03480646Phase I/IImCRPC (+Abiraterone/prednisone or enzalutamide)Active. Not Recruiting
EZH2 INHIBITOR
PF-06821497
NCT03460977Phase ImCRPCRecruiting
EZH2 INHIBITOR
EPZ-6438
NCT04179864Phase IbmCRPC (+Abiraterone/prednisone or enzalutamide)Recruiting
EZH2 INHIBITOR
SHR2554
NCT03741712Phase I/IImCRPC (+SHR3680)Terminated
EZH1/2 INHIBITOR
DS3201
NCT04388852Phase I/IImCRPC (+Ipilimumab)Recruiting
HAT INHIBITORS
P300/CBP INHIBITOR CCS1477NCT03568656Phase I/IImCRPC (+Abiraterone/prednisone or enzalutamide)Recruiting
P300/CBP INHIBITOR:
FT-7051
NCT04575766Phase ImCRPCRecruiting
BRD-CONTAINING PROTEIN INHIBITORS
BMS-986158NCT02419417Phase I/IIAdvanced solid tumoursActive. Not Recruiting
INCB054329NCT02431260Phase I/IIAdvanced solid tumoursTerminated
INCB057643NCT02711137Phase I/IIAdvanced solid tumours (+abiraterone)Terminated
GS-5829NCT02607228Phase I/IImCRPC (+enzalutamide)Terminated
ZEN003694NCT02711956Phase I/IImCRPC (+enzalutamide)Completed. Longer PFS in a subset of patients [16].
ZEN003694NCT02705469Phase ImCRPCCompleted.
ZEN003694NCT04471974Phase IImCRPC (+Enzalutamide + pembrolizumab)Recruiting
GSK525762NCT03150056Phase ImCRPC (+Abiraterone/prednisone or enzalutamide)Completed
ABBV-075NCT02391480Phase ImCRPCCompleted. Not significant antitumour activity [17].
GSK2820151NCT02630251Phase IAdvanced or recurrent solid tumoursTerminated (In 2017, GSK2820151 was terminated due to development of another BET Inhibitor (GSK525762) with a better understanding of the risk benefit profile.)
OTX015/MK-8628NCT02259114Phase IbmCPRCCompleted. Not significant antitumour activity [18].
PLX2853NCT04556617Phase I/IImCPRC (+Abiraterone/prednisone or olaparib)Recruiting
HDMT INHIBITORS
LSD1 INHIBITOR: INCB059872NCT02712905Phase I/IISolid tumours and hematologic malignancyActive. Not Recruiting
LSD1 INHIBITOR: INCB059872NCT02959437Phase I/IIAdvanced Solid tumours (+pembrolizumab + epacadostat)Terminated by Sponsor
LSD1 INHIBITOR: INCB057643NCT02959437Phase I/IIAdvanced Solid tumours (+pembrolizumab + epacadostat)Terminated by Sponsor
HDAC INHIBITORS
VORINOSTAT/SAHANCT00005634Phase ImCRPCCompleted. Determine the tolerability, pharmacokinetic profile, and biological effects of the drug. Not available [19].
VORINOSTAT/SAHANCT00330161Phase IImCRPC with disease
progression on prior chemotherapy
received 400 mg vorinostat/SAHA
orally each day. Disease progression
measured at 6 months. n = 27
Completed. Toxicity: significant toxicities including
fatigue, nausea. IL-6 was higher in patients with toxicity. Seven percent of patients achieved a stable disease state. No PSA decline >50% observed. Median
time to progression and overall survival were 2.8 and 11.7 months, respectively. Significant toxicities reported [19].
VORINOSTAT/SAHA
AND DOCETAXEL
NCT00565227Phase IPatients with advanced and relapsed
tumours received oral vorinostat/SAHA
for the first 14 days of a 21-day cycle,
with docetaxel I.v. on day 4 of each
cycle. n = 12
Completed. Toxicity: neutropenia, peripheral neuropathy,
and gastrointestinal bleeding. The combination
of vorinostat/SAHA and docetaxel was poorly
tolerated. No responses were identified [20].
VORINOSTAT/SAHANCT00589472Phase IILocalised PCa (+Bicalutamide, goserelin acetate, or leuprolide acetate)Completed.
VALPROIC ACIDNCT00530907Phase ICRPC (+Bevacizumab)Completed [21].
PANOBINOSTAT
(LBH589)
NCT00667862Phase III.v. panobinostat (20 mg/m2) was administered to CRPC patients on days 1 and 8 of a 21-day cycle. Disease progression measured at 24 weeks. n = 35Completed. Toxicity: fatigue, thrombocytopenia, nausea; 14% of patients demonstrated a decrease in PSA
but none >50%. No clinical activity [22].
PANOBINOSTATNCT00878436Phase I/IICRPC (+bicalutamide)Completed [23].
PANOBINOSTAT,
DOCETAXEL,
AND PREDNISONE
NCT00663832Phase ICRPC patients received oral panobinostat (20 mg/m2) the first, third and fifth day of the week for 2 consecutive weeks. In addition, patients received oral Panobinostat (15 mg/m2) with docetaxel I.v. (75 mg/m2) every 21 days and oral prednisone (5 mg) twice every day of a 21-day cycle. n = 16Completed. Toxicity: dyspnoea and neutropenia
The combination in patients with CRPC resulted
in 63% of patients with >50% decline in PSA levels. No relevant anti-tumour activity [24].
PANOBINOSTAT DOCETAXEL
AND PREDNISONE
NCT00493766Phase IOn the one hand, oral panobinostat alone is given to patients with progressing hormone refractory prostate cancer. On the other hand, oral Panobinostat along with I.v. docetaxel and oral prednisone is administered. n = 16Completed. Toxicity: dyspnoea, neutropenia, fatigue. Exposure to oral panobinostat was similar with and without docetaxel [25].
PANOBINOSTAT
NCT00670553Phase ILocalised prostate cancer (+External beam radiotherapy)Completed.
PANOBINOSTATNCT00667862Phase ImCRPCCompleted.
ROMIDEPSINNCT00106418Phase IImCRPC patients received romidepsin
(13 mg/m2) intravenously on days 1, 8, and 15 every 21-day cycle. Disease progression measures at 6 months. n = 35
Completed. Toxicity: nausea, fatigue. Two patients reached a confirmed radiological partial response of over 6 months, in addition to >50% PSA decline. Eleven patients had to discontinue the therapy due to toxicity. Romidepsin demonstrated minimal anti-tumour activity in chemonaive patients with CRPC [26]
PRACINOSTAT (SB939)NCT01075308Phase IImCRPCCompleted [27].
MOCETINOSTAT (MGCD0103)NCT00511576Phase IAdvanced cancer tumoursCelgene terminated its collaboration agreement with MethylGene for the development of MGCD0103.
ENTINOSTAT (MS-275)NCT03829930Phase ICRPC (+Enzalutamide)Terminated (Sponsor discontinued the drug).
ENTINOSTAT (MS-275)NCT00020579Phase IAdvanced solid tumours or lymphoma.Completed [28].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

López, J.; Añazco-Guenkova, A.M.; Monteagudo-García, Ó.; Blanco, S. Epigenetic and Epitranscriptomic Control in Prostate Cancer. Genes 2022, 13, 378. https://doi.org/10.3390/genes13020378

AMA Style

López J, Añazco-Guenkova AM, Monteagudo-García Ó, Blanco S. Epigenetic and Epitranscriptomic Control in Prostate Cancer. Genes. 2022; 13(2):378. https://doi.org/10.3390/genes13020378

Chicago/Turabian Style

López, Judith, Ana M. Añazco-Guenkova, Óscar Monteagudo-García, and Sandra Blanco. 2022. "Epigenetic and Epitranscriptomic Control in Prostate Cancer" Genes 13, no. 2: 378. https://doi.org/10.3390/genes13020378

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

Article Metrics

Back to TopTop