Next Article in Journal
The Roles of ceRNAs-Mediated Autophagy in Cancer Chemoresistance and Metastasis
Next Article in Special Issue
Epigenetic Mechanisms of LncRNAs Binding to Protein in Carcinogenesis
Previous Article in Journal
Time-Resolved Profiling Reveals ATF3 as a Novel Mediator of Endocrine Resistance in Breast Cancer
Previous Article in Special Issue
Non-Coding and Regulatory RNAs as Epigenetic Remodelers of Fatty Acid Homeostasis in Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulatory Mechanisms of Epigenetic miRNA Relationships in Human Cancer and Potential as Therapeutic Targets

by
K. M. Taufiqul Arif
,
Esther K. Elliott
,
Larisa M. Haupt
and
Lyn R. Griffiths
*,‡
Centre for Genomics and Personalised Health, Genomics Research Centre, School of Biomedical Sciences, Institute of Health and Biomedical Innovation, Queensland University of Technology (QUT), 60 Musk Ave., Kelvin Grove, Queensland 4059, Australia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this paper.
These authors contributed equally to this paper.
Cancers 2020, 12(10), 2922; https://doi.org/10.3390/cancers12102922
Submission received: 12 September 2020 / Revised: 3 October 2020 / Accepted: 7 October 2020 / Published: 11 October 2020
(This article belongs to the Collection Regulatory and Non-Coding RNAs in Cancer Epigenetic Mechanisms)

Abstract

:

Simple Summary

By the virtue of targeting multiple genes, a microRNA (miRNA) can infer variable consequences on tumorigenesis by appearing as both a tumour suppressor and oncogene. miRNAs can regulate gene expression by modulating genome-wide epigenetic status of genes that are involved in various cancers. These miRNAs perform direct inhibition of key mediators of the epigenetic machinery, such as DNA methyltransferases (DNMTs) and histone deacetylases (HDACs) genes. Along with miRNAs gene expression, similar to other protein-coding genes, miRNAs are also controlled by epigenetic mechanisms. Overall, this reciprocal interaction between the miRNAs and the epigenetic architecture is significantly implicated in the aberrant expression of miRNAs detected in various human cancers. Comprehensive knowledge of the miRNA-epigenetic dynamics in cancer is essential for the discovery of novel anticancer therapeutics.

Abstract

Initiation and progression of cancer are under both genetic and epigenetic regulation. Epigenetic modifications including alterations in DNA methylation, RNA and histone modifications can lead to microRNA (miRNA) gene dysregulation and malignant cellular transformation and are hereditary and reversible. miRNAs are small non-coding RNAs which regulate the expression of specific target genes through degradation or inhibition of translation of the target mRNA. miRNAs can target epigenetic modifier enzymes involved in epigenetic modulation, establishing a trilateral regulatory “epi–miR–epi” feedback circuit. The intricate association between miRNAs and the epigenetic architecture is an important feature through which to monitor gene expression profiles in cancer. This review summarises the involvement of epigenetically regulated miRNAs and miRNA-mediated epigenetic modulations in various cancers. In addition, the application of bioinformatics tools to study these networks and the use of therapeutic miRNAs for the treatment of cancer are also reviewed. A comprehensive interpretation of these mechanisms and the interwoven bond between miRNAs and epigenetics is crucial for understanding how the human epigenome is maintained, how aberrant miRNA expression can contribute to tumorigenesis and how knowledge of these factors can be translated into diagnostic and therapeutic tool development.

Graphical Abstract

1. Introduction

microRNAs (miRNAs) consist of short sequences of non-coding RNA which regulate translation and expression of specific target genes. Currently, there are over 2500 miRNAs referenced on the global micro RNA database, miRbase [1]. The role of miRNAs is to negatively regulate gene expression through binding to the target mRNA to cause transcriptional repression and/or mRNA degradation without modifying the gene sequence. This occurs through a miRNA recognising their target mRNA using either 2–7 nucleotides (partially complementary) or 6–8 nucleotides at the 5′ end of the miRNA annotated as the “seed region”. The exact manner in which protein translation is downregulated by miRNAs is still not clear and may be due to either mRNA degradation and translation inhibition or a combination of these events [2]. miRNA expression is regulated during hematopoietic cell development and differentiation, with these miRNAs playing a direct regulatory role in processes such as cellular proliferation, differentiation, migration and apoptosis [3,4].
Biogenesis and expression of miRNAs are known to be regulated by epigenetic modifications such as DNA methylation, RNA alterations, and histone modifications, with dysregulation of miRNAs being a hallmark of cancer initiation and metastasis [5]. It has also been established that miRNAs control the expression of epigenetic regulators, including DNA methyltransferases and histone deacetylases [6]. miRNAs are involved in complex double-negative feedback loops where miRNA inhibition of an epigenetic regulator is then controlled at the epigenetic level by the same regulator. This miRNA–epigenetic feedback loop has a significant influence on gene expression levels, and dysregulation of the feedback loop can disrupt normal physiological processes, resulting in disease [7]. It is clear that miRNA–gene associations are not linear, hence, functional heterogeneity of a single miRNA across cell types, tissues and disease stages, increases the degree of difficulty in discerning the direct functional pathways regulated by any miRNA [8]. Aberrant miRNA profiling showing altered regulating factors such as cellular proliferation, and migration has been described in many cancers with the majority showing decreased miRNA expression levels in tumour cells in comparison to normal tissue [9].
The first miRNAs to be correlated with cancer were miR-15 and miR-16 in B-cell leukaemia [10]. Since then, the biogenesis of miRNA and target genes such as tumour suppressor and oncogenes has been well established [3,4,9,11,12] with the aid of both experimental and computational analyses; however, further validation using experimental analyses is required if we are to accurately understand the role of miRNA in the multiple functional pathways mediating cancers [3,5,7,9,13,14,15]. Identifying master regulatory miRNAs which regulate both the target mRNA as well as other miRNAs in either the same or a different pathway could provide early prognostic disease biomarkers for both solid tumours and haematological malignancies [9,16,17]. The degree of protein repression by one miRNA can range from mild (2-fold) to significant repression, highlighting the need for better experimental validation. RNase III Drosha is the core RNA-specific nuclease that executes the initiation step of miRNA processing in the nucleus. The resulting miRNAs then regulate the degree of gene repression by the association of enhanced processing of ribosomal RNA precursors by a nuclear dsRNA ribonuclease called Drosha to produce efficient RNA-induced silencing complex (RISC)-induced cleavage of complementary mRNA, resulting in enhanced mRNA repression [18]. Recent studies have shown that a large number of mature miRNAs are methylated at cytosines and adenines in cancer cell lines, as well as human tissues and serum, potentially providing a more accurate cancer diagnostic tool than miRNA expression [19].
The epigenome (the collection of chemical compounds that interact with and regulate the genome) consists of chemical compounds and proteins which regulate gene expression and protein production within cells. It can also be subjected to miRNA-mediated and other posttranscriptional alterations which result in dysregulated miRNA signatures relating to enhanced oncogene expression and downregulation of tumour suppressor genes, leading to tumorigenesis and cancer progression [20]. The epigenome is normally reversible; therefore, miRNA dysregulation might be predictive of cancer development when miRNA alterations become irreversible [21]. Epigenetic markers can also relate to clinical prognosis and may provide a tool for early detection and cancer treatment. By directly targeting miRNAs or epigenetic machinery, malignancy could potentially be treated using therapeutic agents [11,21,22,23] such as those discussed in this review. A considerable number of studies have demonstrated the orchestrated role of epigenetics and miRNA in diverse cellular processes and complex diseases like cancer. Unfortunately, such evidence is dispersed in the huge body of literature, making it difficult for researchers to investigate their reciprocal regulations. One way to achieve this is thorough computational data processing applications of the experimentally proven interactions between epigenetic modifications and miRNAs that can be stored as a searchable database [24,25]. Such information can provide invaluable information to better understand the molecular mechanisms of “epi–miR–epi” in cancer and encourage targeted research toward epigenetics- or miRNA-related drug development.

2. Epigenetic Regulators of Cancer

2.1. DNA Methylation and miRNA Regulation

Aberrant expression of miRNAs significantly affects the gene regulatory mechanisms implicated in cancer development with miRNAs having the intricate ability to act as both oncogenes (oncomiRs) and suppressors [26]. It is an important fact to consider that one miRNA can regulate numerous genes, and can be targeted by several miRNAs [27]. Increasing evidence has indicated that miRNAs expression is under the control of epigenetic regulation, similar to DNA methylation (hyper and hypo), RNA modification, and post-translational modification of histones [28,29,30,31,32,33]. Interestingly, miRNAs contribute to epigenetic modulation by targeting epigenetic modifiers [34,35]. We have summarised the epigenetic regulation of miRNAs in haem-malignancies and solid tumours (Tables S1 and S2), as well as the miRNAs that affect epigenetic regulation and have been reported in various cancers (Table 1).
Disrupted DNA methylation in miRNA loci often leads to downregulation of the miRNA and a greater likelihood of displaying a malignant phenotype. This is seen when miRNAs are more highly expressed and have greater sequence conservation when flanking regions surrounding the miRNA coding sequence are highly methylated [4,9,11]. miRNA target gene promoters are often negatively correlated with DNA hypomethylation, and DNA methylation is regulated by three catalytically active DNA methyltransferases (DNMTs): DNMT1, DNMT3a, and DNMT3b [71]. The majority of studies on DNMT3a/b and miRNAs have primarily focused to date on the regulation of DNMT3a/b targets. However, more recently direct miRNA targeting of DNMT3a/b has shown the potential for both oncogenic [72] and tumour suppressor activities in the progression of cancer such as breast cancer [73].
A well-known miRNA which is often epigenetically regulated is miR-9. Its expression has been shown to be associated with hypermethylation of a cytosine and guanine separated by a phosphate (CpG) island in the miR-9 loci [74,75,76]. miR-9 hypermethylation is seen in many cancers including solid tumours such as breast, colon, and pancreas along with haematological malignancies like acute lymphoblastic leukaemia [75,77,78,79]. It has been suggested that epigenetic silencing of the miR-9 loci as a result of hypermethylation is often an early disease-associated event in breast carcinogenesis [75]. Other commonly hypermethylated miRNAs include miR-92 and miR-29b which target the TET gene family and act as oncogenic miRNAs (oncomiRs) causing reactivation of silenced oncogenes [80,81,82,83].
A number of miRNAs aberrantly silenced are involved in DNA methylation, histone acetylation, H3K4me3 and H3ac modifications and have been identified in haem-malignancies (Table S1) and bladder, cervical, colorectal, gastric, hepatocellular, lung, melanoma, pancreatic and prostate cancers (Table S2). Some of these miRNAs, including let-7a, miR-9, miR-34b-c, miR-124a, miR-127, miR-129, miR-137, miR-148a, miR-152, miR-203, miR-205, the miR-200 family and miR-375, have been frequently reported in several cancers (Table S2) [40,42,46,47,68,69,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99]. Among these, miR-9, miR-34b-c and miR-148a are frequently hypermethylated in aggressive tumours, with this feature proposed as a possible DNA methylation signature for metastasis [66,100,101,102,103]. In breast cancer, epigenetic regulators like DNMTs (1/3), HDAC, JARID1B (histone H3 lysine 4 demethylase) and sp1 (Sp1 Transcription Factor) cause both higher and lower expression of several miRNAs including miR-124.3, miR148a, miR-375, miR-152, members of miR-200, let-7 and miR-34 families [5,36,85,87,104,105,106,107,108,109]. This may be important clinically, with lower expression levels of miR-29c, miR-148a, miR-148b, miR-26a, miR-26b, and miR-203 demonstrated to contribute to DNMT3b overexpression in hypermethylated breast cancer cell lines (Hs578T, HCC1937 and SUM185). In contrast, knockdown of miR-148b, miR-26b, or miR-29c in non-hypermethylated breast cancer cell lines (MDA-MB-468, MDA-MB-415, and BT20) showed increased DNMT3b mRNA levels [40].
These “oncomiRs” may be switched on by abnormal CpG hypomethylation, an indicator of oncogene activation [110]. As discussed earlier, several studies using a range of cell lines have indicated overexpression of numerous miRNAs following treatment with epigenetic drugs [111,112,113]. The let-7a-3 locus is hypermethylated in normal human lung tissue, but in some lung adenocarcinomas, it becomes hypomethylated and overexpressed [114]. miRNA-mediated deregulation of DNMTs has been observed in several cancers. In lung cancer, the miR-29 family that targets DNMT3a-b was shown to be downregulated [30]. In addition, in hepatocellular carcinoma, miR-29a regulates both the DNA DNMT1 and DNMT3b [40]. miR-148 has also been identified to regulate a DNMT3b splice variant through binding to the coding region in gastric cancers [47]. miR-449a was found to be involved in cell growth and viability regulation through repressing the expression of HDAC-1 in prostate cancer cells [67]. Another miRNA frequently reported in a number of cancers is miR-101 [48,49,68]. The histone methyltransferase EZH2 has been shown to be a direct post-transcriptional target of miR-101, with miR-101 downregulation by VEGF resulting in overexpression of EZH2 in angiogenic endothelial cells [49].

2.2. Histone Acetylation and Deacetylation

The chromatin structure is comprised of DNA and histones made up of a chain of 147 nucleotides wrapped around a histone octamer consisting of two copies of each of four histones: H2A, H2B, H3, and H4. Chromatin features are involved in both activation and repression of transcription [115,116,117]. A relationship between miRNA biogenesis and chromatin features around pre-miRNA genomic regions has also been previously reported in which genes located within active chromatin regions have a higher probability of being targeted by miRNAs [118]. The promoters of miRNA target genes have also been shown to be preferentially located in chromatin domains [119]. Additional to the influence of chromatin on miRNA target gene regulation are the impacts of posttranslational modifications of histones through events such as phosphorylation of serine or threonine residues, acetylation and deacetylation of lysine and methylation of lysine or arginine. Histone modifications that regulate epigenetics include methylation (HDM) and acetylation (HDA) which affect DNA accessibility typically modulated by histone acetyltransferase (HAT) and histone deacetylase (HDAC). HDAC causes condensation of the chromatin structure, which prevents binding of transcription factors or proteins to the DNA strand, resulting in gene silencing [120,121]. HDAC inhibits differentiation pathways critical to cellular development and differentiation, as well as the maturation of antibody response [122]. In cancer, demethylation of genes by HDMs often results in upregulation of genes involved in cellular proliferation, migration, and invasion. This provides an opportunity for miRNA-based agents which could regulate proteins in these pathways, thereby inhibiting malignant cellular proliferation [123].

3. Epigenetic Regulation by miRNAs in Cancer

miRNAs can also directly alter epigenetic regulation by post-transcriptionally suppressing the mRNAs of genes involved in the deposition of epigenetic marks (Table 1). Fabbri et al. demonstrated that the miR-29 family targets DNMT3a and DNMT3b, which indirectly controls genome-wide de novo DNA methylation [62]. Soon after, it was validated that miR-29b induces DNA hypomethylation universally in acute myeloid leukaemia (AML) by direct downregulation of DNMT3a and DNMT3b and indirect repression of DNMT1 [37]. DNMT3a is also a direct target of miR-143 [43]. The epigenetic-cally modulated miR-148a suppresses particular isoforms of DNMT3b by targeting the coding sequence [41,47]. Furthermore, DNMT1 has been confirmed as a target of miR-148a and miR-152 [41]. Such a direct effect on the DNA methylation machinery by epigenetically regulated miRNAs could be an indication of an important feedback regulatory mechanism. Several reports have demonstrated that DNMT3a and DNMT3b are fre-quently overexpressed in cancers with poor prognosis, and targets of the miR-29 family, miR-101, miR-143, mir-148a and miR-152 have been reported in various forms of cancer [37,41,43,45,46,47,51,59,61,62,65,66]. Therefore, it appears that all DNMTs are a direct or indirect target of a subset of miRNAs and perform specific functions in the refinement of the expression levels of these crucial epigenetic regulators. As such, dysregulated miRNA expression may contribute to the widespread and often inconsistent changes in DNA methylation patterns detected in cancers, causing both hypo- and hyper-methylation of certain genes and/or regions of the genome [124].
miRNAs can also directly alter epigenetic regulation by post-transcriptionally suppressing the mRNAs of genes involved in the deposition of epigenetic marks (Table 1). Fabbri et al. demonstrated that the miR-29 family targets DNMT3a and DNMT3b, which indirectly controls genome-wide de novo DNA methylation [62]. Soon after, it was validated that miR-29b induces DNA hypomethylation universally in acute myeloid leukaemia (AML) by direct downregulation of DNMT3a and DNMT3b and indirect repression of DNMT1 [37]. DNMT3a is also a direct target of miR-143 [43]. The epigenetic-cally modulated miR-148a suppresses particular isoforms of DNMT3b by targeting the coding sequence [41,47]. Furthermore, DNMT1 has been confirmed as a target of miR-148a and miR-152 [41]. Such a direct effect on the DNA methylation machinery by epigenetically regulated miR NAs could be an indication of an important feedback regulatory mechanism. Several reports have demonstrated that DNMT3a and DNMT3b are fre-quently overexpressed in cancers with poor prognosis, and targets of the miR-29 family, miR-101, miR-143, mir-148a and miR-152 have been reported in various forms of cancer [37,41,43,45,46,47,51,59,61,62,65,66]. Therefore, it appears that all DNMTs are a direct or indirect target of a subset of miRNAs and perform specific functions in the refinement of the expression levels of these crucial epigenetic regulators. As such, dysregulated miRNA expression may contribute to the widespread and often inconsistent changes in DNA methylation patterns detected in cancers, causing both hypo- and hyper-methylation of certain genes and/or regions of the genome [124].
As identified earlier, miR-101 has been reported in a number of cancers, exhibiting direct effects on the epigenetic machinery. Several studies have shown that miR-101 targets EZH2, the catalytic subunit of PRC2, which implies the repressive H3K27me3 signature [38,48,49,59,68]. There is also evidence that miR-128 (downregulated in T-cell leukaemia) [70], miR-138-5p (overexpressed in squamous cell carcinoma) [69], miR-31 (overexpressed in melanoma) [60] also target EZH2. Other important direct epigenetic targets of miRNAs involve HDAC1 and HDAC4 [125,126]. In prostate cancer, HDAC1 is targeted by miR-34b (overexpressed, rendering H3Krme3 modification, also targets HDAC4) [66] and miR-449a (downregulated) [67]. Higher expression of miR-19a, miR-25, miR-32, miR-92b and miR-96 were found to target protein arginine methyltransferase 5 (PRMT5) in leukaemia and lymphoma cells [58]. Nevertheless, it is obvious that an increasing subset of miRNAs is implicated in the regulation of DNA and histone-modifying enzymes (Figure 1), thus highlighting the reputation of these miRNAs in the establishment and maintenance of genomic sustainability and epigenetic architecture.

4. In Silico Analysis of Epi-miRNA Associations

To date, only a small subclass of miRNA–mRNA pairs predicted in silico has been experimentally confirmed [127]. The miRNA–mRNA interfaces are based on sequence complementarity (seed match) and have provided opportunities for in silico prediction of target genes for miRNAs of interest [128]. Although there are several diverse factors affecting the ability of miRNAs to recognise and bind their target, when considering general prediction strategies, these features can commonly be categorised into five groups: (i) attributes of the “seed” pairing; (ii) evolutionary conservation; (iii) abundance of the target site; (iv) accessibility of the target sites; and (v) thermodynamic stability of the miRNA–mRNA duplex [127,129,130]. Typically, a combination of these strategies is utilised by most of the currently available prediction algorithms, with some depending heavily on a particular combination, with others serving to balance the prediction mechanism [131]. Future studies in epigenetic regulation of miRNA expression and miRNA mediated epigenetic regulation linked to downstream signalling pathways are likely to lead to the development of novel drug targets in cancer therapy.
A handful of in silico tools are available and provide information of the regulatory relationship of epi-miRNA with cancer. Functions and link information to the web-application of five selective tools (EpimiR, HMDD, miRCancer, MethyCancer and miRNet) are presented in Table 2. Using these tools, we extracted epi-miRNAs reported in different cancers from EpimiR and HMDD, validated their status using miRCancer and MethyCancer, and finally conducted a network analysis in the miRNet suite (555 miRNAs against 48 epigenetic regulators). The network analysis (Figure 2) identified a list of 101 miRNAs (Table 3; hypergeometric test) to be differentially associated with at least with 1 network (28 miRNAs) and a maximum of 58 networks (miR-515, adj. p = 2.55 × 1011).

5. Epigenetic Strategies for Cancer Therapy

Epigenetic biomarkers may provide a tool for early disease detection, prognostic indicators and/or cancer prevention through the detection of different or aberrant methylation, histone and expression profiles. Previous studies have identified miRNA signatures of well-established epigenetic miRNAs (epi-miRs) which correlate with overall cancer risk, disease staging and survival [17]. For example, expression signatures of miRNAs have been shown to differentiate between acute lymphoblastic leukaemia (ALL) from acute myeloid leukaemia (AML), subgroups [135,136,137], highlighting their utility as a potential clinical biomarker. Through identification of aberrant hypermethylation of miRNAs, particularly in ALL, promising biomarkers for the prediction of clinical significance have been identified [138].
Synthetic oligonucleotides were developed for use as therapeutic agents via in vivo delivery due to their more robust nature against degradation when compared to RNA [139,140]. The five current most common applications used in the regulation of miRNA target gene expression in cancer are reviewed below and include miRNA mimic, anti-miRNA oligonucleotides (anti-miRs), miRNA sponges, miRNA masking and epigenetic inhibitory molecules (Figure 3). These incorporate either the use of oligonucleotide- or virus-based constructs to inhibit oncogenic miRNA or to reactivate a repressed miRNA or tumour suppressor miRNA [22,139,140].

5.1. miRNA Mimics and Inhibitors

The use of miRNAs as a potential epigenetic treatment for specific malignancies is an evolving area of research. Several miRNA-based therapeutics (Figure 3) have been studied as potential cancer therapies for both solid tumours and haematological malignancies in specific tumour environments which display different miRNA expression profiles [11,141,142,143]. As epigenetic silencing of miRNAs is involved in the regulation of key pathways such as leukemogenesis, it may provide a target for epigenetic drugs and provide an avenue for inducing re-expression of key regulatory miRNAs. Two main applications used to inhibit tumour development are miRNA mimics and anti-miRs, also known as anti-miRNA oligonucleotides (AMO) [139,140,143,144]. miRNA mimics are made up of synthetic double-stranded RNA which mimic endogenous miRNAs to bind to target gene mRNAs and result in posttranscriptional gene repression of oncogenes as well as to re-express silenced tumour-suppressive miRNAs (tsmiR) or tumour suppressor genes, thereby inhibiting cancer cell proliferation and cell cycle progression [140,145]. As an example, successful application of a miR-218 mimic was used against acute promyelocytic leukaemia cells and showed reduced cell viability and promoted apoptosis [146]. Conversely, tsmiR-497 was found to inhibit breast cancer cell proliferation and disease progression [145]. miRNA mimics have also provided great insight into the functional impact of specific miRNAs in signalling pathways such as JAK/STAT and demethylation, which are commonly implicated in malignant phenotypes [142].
Anti-miRs can also alter miRNA-related pathways by binding and blocking oncogenic miRNA access to the mRNA transcript, thereby resulting in either slowed or repressed tumour development, as in the case of anti-miR-126 which resulted in successful inhibition of leukemic cells [147]. miRNA mimics and anti-miRs have the potential to offer personalised miRNA expression therapeutics [141]. The main advantage of epigenetic miRNA (epi-miR) therapeutics would be the ability to regulate multiple pathways through the modification of a single miRNA [5] to provide an adjunct therapeutic agent in the management of specific cancers. Current limitations of this epi-miR technology include the difficulties in ascertaining master regulatory miRNA to prevent unwanted impacts on off-target genes and pathways, along with the low efficiency of these applications due to the variability in tissue and staging-specific miRNA expression [148].

5.2. miRNA Sponges

An alternative miRNA inhibitor which can be expressed in cells is termed an miRNA sponge. These inhibitors are transcripts derived from promoter regions which contain multiple common binding sites to the target oncomiR of interest [22]. Vectors encoding these sponges are transiently transfected into cells and allow the sponge to bind to the target oncomiR to inhibit mRNA binding [22]. Sponges are 2–7 nucleotides long and inhibit miRNAs through a complementary heptameric sequence. This allows a single sponge to inhibit an entire miRNA family [22] and has been successfully applied to the repression of oncogenes (Figure 3) in various types of cancers such as leukaemia, sarcoma, breast cancer, renal cancer, lung cancer and melanoma [74,149]. In addition to this, some researchers have successfully used a multi-potent miRNA sponge to simultaneously target and repress multiple oncogenic miRNAs such as miR-155, miR-21 and miR-221/222 in breast and pancreatic cancer cell lines: MDA-MB-436, MCF-7, MIA-Paca-2, Panc-1 and BxPC3 [150].

5.3. miRNA Masking

miRNA-masking (miR-mask) applications contain AMO which have been commonly modified with single-chain 2′-O-methyl to increase binding and nuclease activity. miR masks form a 22 nucleotide antisense to an mRNA target of an endogenous miRNA of interest [23], and rather than interacting with the miRNA, they complement and bind to the 3′-UTR site of the target mRNA. This action allows the miR mask to block the endogenous miRNA binding site on the target mRNA disrupting the miRNA inhibitory function. A recent study established that targeting miR-522 led to reduced proliferation of non-small cell lung cancer [151]. Rather than being gene-specific, the effect of miR masks is sequence-specific with potential for adverse side effects and toxicity (Figure 3) [152]. To date, miR-masking efficiency, accuracy, and toxicity still remain inconsistent, making them less suitable for therapeutic use in the treatment of cancers.

5.4. Epigenetic Inhibitory Molecules

Epigenetic drug therapies contain inhibitory molecules which target epigenetic machinery such as DNA methyltransferase inhibitors (DNMTi) and histone deacetylase inhibitors (HDACi) [153,154,155]. DNMTi drugs are used to irreversibly inhibit DNMT enzymatic activity and trigger its degradation This application is well established as an epigenetic regulatory agent for the inhibition of epigenetic mutations and oncogene expression [153,154] (Figure 3). A therapeutic application for DNMTis, for example, is to inhibit abnormal tumour gene expression by disrupting key tumour initiation and progression pathways [156]. HDACi alters the acetylation and deacetylation of histone lysines, in which deacetylation is a known contributor to abnormal gene expression in malignancy. Therefore, HDACis can be used to block cell proliferation, promote differentiation and induce apoptosis to reverse cancer initiation and progression [155]. Both DNMTi and HDACi applications have been successfully used in repression of various types of haematological malignancies, however, this application showed variable and limited success in solid tumours [157,158].

6. Epigenetic Therapeutics in Cancer Clinical Trials

As we have seen by the successful clinical introduction of epigenetic inhibitors like DNMTi and HDACi [153] in the treatment of haematological malignancies [157], epigenetic-based applications are powerful therapeutic agents used in cancer care. To date, several clinical studies have been completed, while others are currently underway investigating the use of miRNA biomarkers in various types of cancer. A phase 1 trial using miR-16 mimic incorporated with epidermal growth factor receptor (EGFR) targeting antibody called TargomiR, was used for patients with either recurrent malignant pleural mesothelioma or non-small cell lung cancer [159,160]. The findings were promising and showed that the miR-16 mimic is beneficial for patients with terminal mesothelioma patients with less than a 10% chance of 5-year survival [161]. A less successful 2013 phase 1 clinical trial investigated the miRNA-34 drug mimic MRX34 in patients with either primary liver cancer, solid tumours, or hematologic malignancy. The study was terminated after some patients experienced serious adverse reactions to the investigational drug [162].

7. Overcoming Limitations of miRNA Biomarkers and Therapeutic Agents

A major limitation to the utility of miRNAs as biomarkers for the diagnosis and monitoring of disease progression in malignancy is the natural variability in miRNA expression levels across tissue types and disease stages [163,164,165]. This issue alone makes standardising sample collection methods for use in clinical correlations more challenging. Further to this, once a miRNA biomarker or target gene has been established, a significant limitation to the use of a therapeutic miRNA or epigenetic inhibitory drug is the risk of a serious adverse event due to toxicities like those seen in studies with the miRNA mimic MRX34 and HDAC-based inhibitor drugs [155,162]. As discussed above, this is where the role of rigorous clinical trials has greatly contributed to the recent advancements of miRNA therapeutics. Other significant limitations which are yet to be overcome in some specific miR-malignant cell applications include inefficient delivery, inefficient cellular uptake, short half-life, low intracellular release and low in vivo stability [23,166]. In a clinical setting, non-viral based therapeutic miRNAs may be favoured due to their stable composition, lower immunogenicity and ease of manufacturing [23]. Delivery mechanisms utilising nanoparticles, polymers and liposomes for mediated drug delivery have also been favoured in recent studies [167,168]. However, these applications require further optimisation before miRNAs can become embedded in standard cancer therapeutic development [169]. A possible solution to overcoming these inefficiencies may be direct intra-tumoral injections of miRNA drugs to enhance target efficacy and reduce adverse reactions [144,149,170]. As research continues to advance in this area, several new clinical trials are likely to be conducted to assess the efficiency and reliability of these miRNA therapeutics [11].
In addition, one puzzling phenomenon in cancer treatment is recurrence of cancer cells in a more aggressive manner and avoidance of apoptosis with higher metastatic potential [171]. The reversal of apoptosis is seen in some cancer cells and often results in more aggressive tumours and metastasis. The exact mechanism behind this event is still unknown; however, this reversal process is known to trigger a transition from non-stem cancer cells (NSCCs) to CSCs, which potentially could be suppressed through the use of DNA methylation or demethylation inhibitors before apoptosis induction [171]. A particular state of the tumour cells termed as polyploid giant cancer cells (PGCCs) has been suggested to be responsible for facilitating the escape from therapeutic-induced senescence [172,173]. Tumours can originate from a stem cell via dedifferentiation, therefore, the use of DNMT inhibitors may reactivate tumour suppressor genes, which could disrupt PGCC-mediated dedifferentiation and the development and progression of tumours [174,175]. The level of PGCC increases after exposure to chemotherapeutic drugs like 5-fluorouracil (5-FU) [175]. 5-fluorouracil is used to treat colorectal cancer (CRC), and, interestingly, resistance of CRC to 5-FU has been reported to be associated with the upregulation of nuclear factor-erythroid 2-related factor 2 (Nrf2) via DNA demethylase ten-eleven translocation (TET)-dependent DNA methylation [176]. Therefore, epi-miR based therapeutic strategies should consider the aftermath of any targeted treatment.

8. Conclusions

In this review, we have detailed the relationship between epigenetic alteration of miRNAs and miRNA-mediated epigenetic modifications in cancer. Aberrant DNA methylation generally causes miRNA dysregulation in cancer, and methylation of specific miRNA genes may be a valuable biomarker for cancer diagnosis and prognosis. Variation in the histone architecture also disrupts miRNA expression. Such dysregulation in miRNA expression leads to genome-wide epigenetic abnormalities. It is to be expected that further study of the association between epigenetic regulation of and by miRNAs will lead to the innovative identification and use of new biomarkers as well as therapeutic targets against cancer. The evaluation of the efficiency and reproducibility of miRNA biomarkers across cancer classifications, disease stages and tissues types are also likely to advance in the near future. Reactivation of epigenetic mechanisms of miRNA expression could be an encouraging novel approach of cancer therapy by targeting epigenetically regulated miRNA genes using drugs that inhibit methylation and/or histone modification. Computational analysis of gene regulation is a complex but reliable platform with which to study the epi-miRNA relationship with numerous cellular processes and cancer. These platforms have been successful in identifying epigenetically regulated miRNAs in many malignancies which have paved the way for the use of miRNA-based therapeutics for the treatment of both haematological and solid tumour malignancies. A strategic combination of laboratory-based experimental data, clinical resources and high-throughput computational applications will provide a significant body of knowledge on “epi–miR–epi” regulation for further development of diagnostic and therapeutic suitable for cancer treatments. Further optimisation of delivery applications such as nanoparticle and liposome-mediated delivery, along with additional clinical research trials are required before miRNAs therapeutics can become established as a standard of care therapy in common cancers. With this area of cancer research rapidly evolving, clinical validation of miRNA-based therapeutics may become established following the successful completion of further clinical trials.

Supplementary Materials

The following are available online at https://www.mdpi.com/2072-6694/12/10/2922/s1, Table S1: Epigenetic regulation of miRNAs in leukaemia and lymphomas, Table S2: Epigenetic regulations of miRNAs in solid cancers.

Author Contributions

E.K.E. and K.M.T.A. conceived the original idea, developed the conceptual framework, and drafted the manuscript. K.M.T.A. contributed to the acquisition of data, carried out the network analysis, and interpretation. L.M.H. provided critical advice on conceptualisation and revision and helped shape the manuscript and provided final approval of the manuscript. L.R.G. supervised the project, provided critical feedback on conceptualisation and manuscript preparation and provided final approval of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Funding for this study has been generously provided by a Herbert family philanthropic donation, Funding has also been provided by the GRC Genomics Lymphoma Project (GLP) fund. This research was also supported by infrastructure purchased with Australian Government EIF Super Science Funds as part of the Therapeutic Innovation Australia—Queensland Node project (LRG). K.M.T.A. was supported study by Queensland University of Technology Post-graduate Research Award (QUTPRA) and Tuition Fee scholarship schemes.

Acknowledgments

The authors would like to thank Ian Peall for fruitful discussions and suggestions for the manuscript.

Conflicts of Interest

The authors declare no conflict of interests.

References

  1. miRBase miRNA. Available online: http://www.mirbase.org/ (accessed on 1 August 2020).
  2. Di Lisio, L.; Sanchez-Beato, M.; Gomez-Lopez, G.; Rodriguez, M.E.; Montes-Moreno, S.; Mollejo, M.; Menarguez, J.; Martinez, M.A.; Alves, F.J.; Pisano, D.G.; et al. MicroRNA signatures in B-cell lymphomas. Blood Cancer J. 2012, 2, e57. [Google Scholar] [CrossRef] [Green Version]
  3. Khan, A.Q.; Ahmed, E.I.; Elareer, N.R.; Junejo, K.; Steinhoff, M.; Uddin, S. Role of miRNA-Regulated Cancer Stem Cells in the Pathogenesis of Human Malignancies. Cells 2019, 8, 840. [Google Scholar] [CrossRef] [Green Version]
  4. Calin, G.A.; Croce, C.M. MicroRNA signatures in human cancers. Nat. Rev. Cancer 2006, 6, 857–866. [Google Scholar] [CrossRef]
  5. Humphries, B.; Wang, Z.; Yang, C. MicroRNA Regulation of Epigenetic Modifiers in Breast Cancer. Cancers 2019, 11, 897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Iorio, M.V.; Piovan, C.; Croce, C.M. Interplay between microRNAs and the epigenetic machinery: An intricate network. Biochim. Biophys. Acta 2010, 1799, 694–701. [Google Scholar] [CrossRef] [PubMed]
  7. Yao, Q.; Chen, Y.; Zhou, X. The roles of microRNAs in epigenetic regulation. Curr. Opin. Chem. Biol. 2019, 51, 11–17. [Google Scholar] [CrossRef] [PubMed]
  8. Ma, F.; Liu, X.; Li, D.; Wang, P.; Li, N.; Lu, L.; Cao, X. MicroRNA-466l upregulates IL-10 expression in TLR-triggered macrophages by antagonizing RNA-binding protein tristetraprolin-mediated IL-10 mRNA degradation. J. Immunol. 2010, 184, 6053–6059. [Google Scholar] [CrossRef] [PubMed]
  9. Esquela-Kerscher, A.; Slack, F.J. Oncomirs—microRNAs with a role in cancer. Nat. Rev. Cancer 2006, 6, 259–269. [Google Scholar] [CrossRef]
  10. Calin, G.A.; Dumitru, C.D.; Shimizu, M.; Bichi, R.; Zupo, S.; Noch, E.; Aldler, H.; Rattan, S.; Keating, M.; Rai, K.; et al. Frequent deletions and down-regulation of micro- RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia. Proc. Natl. Acad. Sci. USA 2002, 99, 15524–15529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Abba, M.L.; Patil, N.; Leupold, J.H.; Moniuszko, M.; Utikal, J.; Niklinski, J.; Allgayer, H. MicroRNAs as novel targets and tools in cancer therapy. Cancer Lett. 2017, 387, 84–94. [Google Scholar] [CrossRef]
  12. Guo, J.; Miao, Y.; Xiao, B.; Huan, R.; Jiang, Z.; Meng, D.; Wang, Y. Differential expression of microRNA species in human gastric cancer versus non-tumorous tissues. J. Gastroenterol. Hepatol. 2009, 24, 652–657. [Google Scholar] [CrossRef] [PubMed]
  13. Getaneh, Z.; Asrie, F.; Melku, M. MicroRNA profiles in B-cell non-Hodgkin lymphoma. EJIFCC 2019, 30, 195–214. [Google Scholar] [PubMed]
  14. Bradshaw, G.; Sutherland, H.G.; Haupt, L.M.; Griffiths, L.R. Dysregulated MicroRNA Expression Profiles and Potential Cellular, Circulating and Polymorphic Biomarkers in Non-Hodgkin Lymphoma. Genes 2016, 7, 130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Tagawa, H.; Ikeda, S.; Sawada, K. Role of microRNA in the pathogenesis of malignant lymphoma. Cancer Sci. 2013, 104, 801–809. [Google Scholar] [CrossRef] [PubMed]
  16. Caramuta, S.; Lee, L.; Ozata, D.M.; Akcakaya, P.; Georgii-Hemming, P.; Xie, H.; Amini, R.M.; Lawrie, C.H.; Enblad, G.; Larsson, C.; et al. Role of microRNAs and microRNA machinery in the pathogenesis of diffuse large B-cell lymphoma. Blood Cancer J. 2013, 3, e152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Taslim, C.; Weng, D.Y.; Brasky, T.M.; Dumitrescu, R.G.; Huang, K.; Kallakury, B.V.; Krishnan, S.; Llanos, A.A.; Marian, C.; McElroy, J.; et al. Discovery and replication of microRNAs for breast cancer risk using genome-wide profiling. Oncotarget 2016, 7, 86457–86468. [Google Scholar] [CrossRef]
  18. Hwang, H.W.; Wentzel, E.A.; Mendell, J.T. Cell-cell contact globally activates microRNA biogenesis. Proc. Natl. Acad. Sci. USA 2009, 106, 7016–7021. [Google Scholar] [CrossRef] [Green Version]
  19. Fuso, A.; Raia, T.; Orticello, M.; Lucarelli, M. The complex interplay between DNA methylation and miRNAs in gene expression regulation. Biochimie 2020, 173, 12–16. [Google Scholar] [CrossRef]
  20. Suzuki, H.; Takatsuka, S.; Akashi, H.; Yamamoto, E.; Nojima, M.; Maruyama, R.; Kai, M.; Yamano, H.O.; Sasaki, Y.; Tokino, T.; et al. Genome-wide profiling of chromatin signatures reveals epigenetic regulation of MicroRNA genes in colorectal cancer. Cancer Res. 2011, 71, 5646–5658. [Google Scholar] [CrossRef] [Green Version]
  21. Abba, M.; Patil, N.; Leupold, J.H.; Allgayer, H. MicroRNAs-from metastasis prediction to metastasis prevention? Mol. Cell. Oncol. 2016, 3, e1074336. [Google Scholar] [CrossRef] [Green Version]
  22. Ebert, M.S.; Sharp, P.A. MicroRNA sponges: Progress and possibilities. RNA 2010, 16, 2043–2050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Li, Y.; Humphries, B.; Yang, C.; Wang, Z. Nanoparticle-Mediated Therapeutic Agent Delivery for Treating Metastatic Breast Cancer-Challenges and Opportunities. Nanomaterials 2018, 8, 361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Dai, E.; Yu, X.; Zhang, Y.; Meng, F.; Wang, S.; Liu, X.; Liu, D.; Wang, J.; Li, X.; Jiang, W. EpimiR: A database of curated mutual regulation between miRNAs and epigenetic modifications. Database 2014, 2014, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Fan, Y.; Siklenka, K.; Arora, S.K.; Ribeiro, P.; Kimmins, S.; Xia, J. miRNet—Dissecting miRNA-target interactions and functional associations through network-based visual analysis. Nucleic Acids Res. 2016, 44, W135–W141. [Google Scholar] [CrossRef] [PubMed]
  26. Tan, W.; Liu, B.; Qu, S.; Liang, G.; Luo, W.; Gong, C. MicroRNAs and cancer: Key paradigms in molecular therapy (Review). Oncol. Lett. 2018, 15, 2735–2742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Hayes, J.; Peruzzi, P.P.; Lawler, S. MicroRNAs in cancer: Biomarkers, functions and therapy. Trends Mol. Med. 2014, 20, 460–469. [Google Scholar] [CrossRef]
  28. Suzuki, H.; Maruyama, R.; Yamamoto, E.; Kai, M. Epigenetic alteration and microRNA dysregulation in cancer. Front. Genet. 2013, 4, 1–8. [Google Scholar] [CrossRef] [Green Version]
  29. Yang, N.; Coukos, G.; Zhang, L. MicroRNA epigenetic alterations in human cancer: One step forward in diagnosis and treatment. Int. J. Cancer 2008, 122, 963–968. [Google Scholar] [CrossRef]
  30. Suzuki, H.; Maruyama, R.; Yamamoto, E.; Kai, M. DNA methylation and microRNA dysregulation in cancer. Mol. Oncol. 2012, 6, 567–578. [Google Scholar] [CrossRef] [Green Version]
  31. Morales, S.; Monzo, M.; Navarro, A. Epigenetic regulation mechanisms of microRNA expression. Biomol. Concepts 2017, 8, 203–212. [Google Scholar] [CrossRef]
  32. Cora, D.; Re, A.; Caselle, M.; Bussolino, F. MicroRNA-mediated regulatory circuits: Outlook and perspectives. Phys. Biol. 2017, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Osella, M.; Riba, A.; Testori, A.; Corá, D.; Caselle, M. Interplay of microRNA and epigenetic regulation in the human regulatory network. Front. Genet. 2014, 5, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Malumbres, M. MiRNAs and cancer: An epigenetics view. Mol. Aspects Med. 2013, 34, 863–874. [Google Scholar] [CrossRef] [Green Version]
  35. Sato, F.; Tsuchiya, S.; Meltzer, S.J.; Shimizu, K. MicroRNAs and epigenetics. FEBS J. 2011, 278, 1598–1609. [Google Scholar] [CrossRef] [PubMed]
  36. Pogribny, I.P.; Filkowski, J.N.; Tryndyak, V.P.; Golubov, A.; Shpyleva, S.I.; Kovalchuk, O. Alterations of microRNAs and their targets are associated with acquired resistance of MCF-7 breast cancer cells to cisplatin. Int. J. Cancer 2010, 127, 1785–1794. [Google Scholar] [CrossRef]
  37. Garzon, R.; Liu, S.; Fabbri, M.; Liu, Z.; Heaphy, C.E.A.; Callegari, E.; Schwind, S.; Pang, J.; Yu, J.; Muthusamy, N.; et al. MicroRNA-29b induces global DNA hypomethylation and tumor suppressor gene reexpression in acute myeloid leukemia by targeting directly DNMT3A and 3B and indirectly DNMT1. Blood 2009, 113, 6411–6418. [Google Scholar] [CrossRef] [Green Version]
  38. Friedman, J.M.; Liang, G.; Liu, C.C.; Wolff, E.M.; Tsai, Y.C.; Ye, W.; Zhou, X.; Jones, P.A. The putative tumor suppressor microRNA-101 modulates the cancer epigenome by repressing the polycomb group protein EZH2. Cancer Res. 2009, 69, 2623–2629. [Google Scholar] [CrossRef] [Green Version]
  39. Tan, Y.; Zhang, B.; Wu, T.; Skogerbø, G.; Zhu, X.; Guo, X.; He, S.; Chen, R. Transcriptional inhibiton of Hoxd4 expression by miRNA-10a in human breast cancer cells. BMC Mol. Biol. 2009, 10. [Google Scholar] [CrossRef] [Green Version]
  40. Sandhu, R.; Rivenbark, A.G.; Coleman, W.B. Loss of post-transcriptional regulation of DNMT3b by microRNAs: A possible molecular mechanism for the hypermethylation defect observed in a subset of breast cancer cell lines. Int. J. Oncol. 2012, 41, 721–732. [Google Scholar] [CrossRef] [Green Version]
  41. Xu, Q.; Jiang, Y.; Yin, Y.; Li, Q.; He, J.; Jing, Y.; Qi, Y.T.; Xu, Q.; Li, W.; Lu, B.; et al. A regulatory circuit of miR-148a/152 and DNMT1 in modulating cell transformation and tumor angiogenesis through IGF-IR and IRS1. J. Mol. Cell Biol. 2013, 5, 3–13. [Google Scholar] [CrossRef] [Green Version]
  42. Ren, G.; Baritaki, S.; Marathe, H.; Feng, J.; Park, S.; Beach, S.; Bazeley, P.S.; Beshir, A.B.; Fenteany, G.; Mehra, R.; et al. Polycomb protein EZH2 regulates tumor invasion via the transcriptional repression of the metastasis suppressor RKIP in breast and prostate cancer. Cancer Res. 2012, 72, 3091–3104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ng, E.K.O.; Tsang, W.P.; Ng, S.S.M.; Jin, H.C.; Yu, J.; Li, J.J.; Röcken, C.; Ebert, M.P.A.; Kwok, T.T.; Sung, J.J.Y. MicroRNA-143 targets DNA methyltransferases 3A in colorectal cancer. Br. J. Cancer 2009, 101, 699–706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Wang, H.; Wu, J.; Meng, X.; Ying, X.; Zuo, Y.; Liu, R.; Pan, Z.; Kang, T.; Huang, W. Microrna-342 inhibits colorectal cancer cell proliferation and invasion by directly targeting dna methyltransferase 1. Carcinogenesis 2011, 32, 1033–1042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Nguyen, T.; Kuo, C.; Nicholl, M.B.; Sim, M.S.; Turner, R.R.; Morton, D.L.; Hoon, D.S.B. Downregulation of microRNA-29c is associated with hypermethylation of tumor-related genes and disease outcome in cutaneous melanoma. Epigenetics 2011, 6, 388–394. [Google Scholar] [CrossRef]
  46. Tsuruta, T.; Kozaki, K.I.; Uesugi, A.; Furuta, M.; Hirasawa, A.; Imoto, I.; Susumu, N.; Aoki, D.; Inazawa, J. miR-152 is a tumor suppressor microRNA that is silenced by DNA hypermethylation in endometrial cancer. Cancer Res. 2011, 71, 6450–6462. [Google Scholar] [CrossRef] [Green Version]
  47. Zhu, A.; Xia, J.; Zuo, J.; Jin, S.; Zhou, H.; Yao, L.; Huang, H.; Han, Z. MicroRNA-148a is silenced by hypermethylation and interacts with DNA methyltransferase 1 in gastric cancer. Med. Oncol. 2012, 29, 2701–2709. [Google Scholar] [CrossRef]
  48. Smits, M.; Nilsson, J.; Mir, S.E.; Van Der Stoop, P.M.; Hulleman, E.; Niers, J.M.; Hamer, P.C.D.W.; Marquez, V.E.; Cloos, J.; Krichevsky, A.M.; et al. miR-101 is down-regulated in glioblastoma resulting in EZH2- induced proliferation, migration, and angiogenesis AbstrAct: Background: Glioblastoma (GBM) is a malignant brain tumor with dismal. Oncotarget 2010, 1, 710–720. [Google Scholar] [CrossRef] [Green Version]
  49. Smits, M.; Mir, S.E.; Nilsson, J.R.A.; van der Stoop, P.M.; Niers, J.M.; Marquez, V.E.; Cloos, J.; Breakefield, X.O.; Krichevsky, A.M.; Noske, D.P.; et al. Down-regulation of miR-101 in endothelial cells promotes blood vessel formation through reduced repression of EZH2. PLoS ONE 2011, 6. [Google Scholar] [CrossRef]
  50. Zhang, Z.; Tang, H.; Wang, Z.; Zhang, B.; Liu, W.; Lu, H.; Xiao, L.; Liu, X.; Wang, R.; Li, X.; et al. MiR-185 Targets the DNA Methyltransferases 1 and Regulates Global DNA Methylation in human glioma. Mol. Cancer 2011, 10, 124. [Google Scholar] [CrossRef] [Green Version]
  51. Wei, X.; Xiang, T.; Ren, G.; Tan, C.; Liu, R.; Xu, X.; Wu, Z. MiR-101 is down-regulated by the hepatitis B virus x protein and induces aberrant DNA methylation by targeting DNA methyltransferase 3A. Cell. Signal. 2013, 25, 439–446. [Google Scholar] [CrossRef]
  52. Yuan, J.H.; Yang, F.; Chen, B.F.; Lu, Z.; Huo, X.S.; Zhou, W.P.; Wang, F.; Sun, S.H. The histone deacetylase 4/SP1/microrna-200a regulatory network contributes to aberrant histone acetylation in hepatocellular carcinoma. Hepatology 2011, 54, 2025–2035. [Google Scholar] [CrossRef] [PubMed]
  53. Ngo-Yin Fan, D.; Ho-Ching Tsang, F.; Hoi-Kam Tam, A.; Leung-Kuen Au, S.; Chak-Lui Wong, C.; Wei, L.; Man-Fong Lee, J.; He, X.; Oi-Lin Ng, I.; Wong, C.M. Histone lysine methyltransferase, suppressor of variegation 3-9 homolog 1, promotes hepatocellular carcinoma progression and is negatively regulated by microRNA-125b. Hepatology 2013, 57, 637–647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Huang, J.; Wang, Y.; Guo, Y.; Sun, S. Down-regulated microRNA-152 induces aberrant DNA methylation in hepatitis B virus-related hepatocellular carcinoma by targeting DNA methyltransferase 1. Hepatology 2010, 52, 60–70. [Google Scholar] [CrossRef] [PubMed]
  55. Chen, Y.; Luo, J.; Tian, R.; Sun, H.; Zou, S. miR-373 negatively regulates methyl-CpG-binding domain protein 2 (MBD2) in hilar cholangiocarcinoma. Dig. Dis. Sci. 2011, 56, 1693–1701. [Google Scholar] [CrossRef]
  56. Chen, Y.J.; Luo, J.; Yang, G.Y.; Yang, K.; Wen, S.Q.; Zou, S.Q. Mutual regulation between microRNA-373 and methyl-CpGbinding domain protein 2 in hilar cholangiocarcinoma. World J. Gastroenterol. 2012, 18, 3849–3861. [Google Scholar] [CrossRef]
  57. Braconi, C.; Huang, N.; Patel, T. Microrna-dependent regulation of DNA methyltransferase-1 and tumor suppressor gene expression by interleukin-6 in human malignant cholangiocytes. Hepatology 2010, 51, 881–890. [Google Scholar] [CrossRef] [Green Version]
  58. Wang, L.; Pal, S.; Sif, S. Protein Arginine Methyltransferase 5 Suppresses the Transcription of the RB Family of Tumor Suppressors in Leukemia and Lymphoma Cells. Mol. Cell. Biol. 2008, 28, 6262–6277. [Google Scholar] [CrossRef] [Green Version]
  59. Cho, H.M.; Jeon, H.S.; Lee, S.Y.; Jeong, K.J.; Park, S.Y.; Lee, H.Y.; Lee, J.U.; Kim, J.H.; Kwon, S.J.; Choi, E.; et al. microRNA-101 inhibits lung cancer invasion through the regulation of enhancer of zeste homolog 2. Exp. Ther. Med. 2011, 2, 963–967. [Google Scholar] [CrossRef] [Green Version]
  60. Asangani, I.A.; Harms, P.W.; Dodson, L.; Pandhi, M.; Kunju, L.P.; Maher, C.A.; Fullen, D.R.; Johnson, T.M.; Giordano, T.J.; Palanisamy, N.; et al. Genetic and epigenetic loss of microRNA-31 leads to feed-forward expression of EZH2 in melanoma. Oncotarget 2012, 3, 1011–1025. [Google Scholar] [CrossRef] [Green Version]
  61. Amodio, N.; Leotta, M.; Bellizzi, D.; Di Martino, M.T.; D’Aquila, P.; Lionetti, M.; Fabiani, F.; Leone, E.; Gullà, A.M.; Passarino, G.; et al. DNA-demethylating and anti-tumor activity of synthetic miR-29b mimics in multiple myeloma. Oncotarget 2012, 3, 1246–1258. [Google Scholar] [CrossRef] [Green Version]
  62. Fabbri, M.; Garzon, R.; Cimmino, A.; Liu, Z.; Zanesi, N.; Callegari, E.; Liu, S.; Alder, H.; Costinean, S.; Fernandez-Cymering, C.; et al. MicroRNA-29 family reverts aberrant methylation in lung cancer by targeting DNA methyltransferases 3A and 3B. Proc. Natl. Acad. Sci. USA. 2007, 104, 15805–15810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Li, G.; Zhao, J.; Peng, X.; Liang, J.; Deng, X.; Chen, Y. The mechanism involved in the loss of PTEN expression in NSCLC tumor cells. Biochem. Biophys. Res. Commun. 2012, 418, 547–552. [Google Scholar] [CrossRef] [PubMed]
  64. Samakoglu, S.; Deevi, D.S.; Li, H.; Wang, S.; Murphy, M.; Bao, C.; Bassi, R.; Prewett, M.; Tonra, J.R. Preclinical rationale for combining an EGFR antibody with cisplatin/gemcitabine for the treatment of NSCLC. Cancer Genom. Proteom. 2012, 9, 77–92. [Google Scholar] [PubMed]
  65. Xiang, Y.; Ma, N.; Wang, D.; Zhang, Y.; Zhou, J.; Wu, G.; Zhao, R.; Huang, H.; Wang, X.; Qiao, Y.; et al. MiR-152 and miR-185 co-contribute to ovarian cancer cells cisplatin sensitivity by targeting DNMT1 directly: A novel epigenetic therapy independent of decitabine. Oncogene 2014, 33, 378–386. [Google Scholar] [CrossRef] [Green Version]
  66. Majid, S.; Dar, A.A.; Saini, S.; Shahryari, V.; Arora, S.; Zaman, M.S.; Chang, I.; Yamamura, S.; Tanaka, Y.; Chiyomaru, T.; et al. miRNA-34b inhibits prostate cancer through demethylation, active chromatin modifications, and AKT pathways. Clin. Cancer Res. 2013, 19, 73–84. [Google Scholar] [CrossRef] [Green Version]
  67. Noonan, E.J.; Place, R.F.; Pookot, D.; Basak, S.; Whitson, J.M.; Hirata, H.; Giardina, C.; Dahiya, R. MiR-449a targets HDAC-1 and induces growth arrest in prostate cancer. Oncogene 2009, 28, 1714–1724. [Google Scholar] [CrossRef] [Green Version]
  68. Cao, P.; Deng, Z.; Wan, M.; Huang, W.; Cramer, S.D.; Xu, J.; Lei, M.; Sui, G. MicroRNA-101 negatively regulates Ezh2 and its expression is modulated by androgen receptor and HIF-1α/HIF-1β. Mol. Cancer 2010, 9, 1–12. [Google Scholar] [CrossRef] [Green Version]
  69. Liu, X.; Wang, C.; Chen, Z.; Jin, Y.; Wang, Y.; Kolokythas, A.; Dai, Y.; Zhou, X. MicroRNA-138 suppresses epithelial-mesenchymal transition in squamous cell carcinoma cell lines. Biochem. J. 2011, 440, 23–31. [Google Scholar] [CrossRef] [Green Version]
  70. Sasaki, D.; Imaizumi, Y.; Hasegawa, H.; Osaka, A.; Tsukasaki, K.; Lim Choi, Y.; Mano, H.; Marquez, V.E.; Hayashi, T.; Yanagihara, K.; et al. Overexpression of enhancer of zeste homolog 2 with trimethylation of lysine 27 on histone H3 in adult T-cell leukemia/lymphoma as a target for epigenetic therapy. Haematologica 2011, 96, 712–719. [Google Scholar] [CrossRef] [Green Version]
  71. Okano, M.; Bell, D.W.; Haber, D.A.; Li, E. DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 1999, 99, 247–257. [Google Scholar] [CrossRef] [Green Version]
  72. Roscigno, G.; Quintavalle, C.; Donnarumma, E.; Puoti, I.; Diaz-Lagares, A.; Iaboni, M.; Fiore, D.; Russo, V.; Todaro, M.; Romano, G.; et al. MiR-221 promotes stemness of breast cancer cells by targeting DNMT3b. Oncotarget 2016, 7, 580–592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Pang, Y.; Liu, J.; Li, X.; Xiao, G.; Wang, H.; Yang, G.; Li, Y.; Tang, S.C.; Qin, S.; Du, N.; et al. MYC and DNMT3A-mediated DNA methylation represses microRNA-200b in triple negative breast cancer. J. Cell Mol. Med. 2018, 22, 6262–6274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Ma, L.; Young, J.; Prabhala, H.; Pan, E.; Mestdagh, P.; Muth, D.; Teruya-Feldstein, J.; Reinhardt, F.; Onder, T.T.; Valastyan, S.; et al. miR-9, a MYC/MYCN-activated microRNA, regulates E-cadherin and cancer metastasis. Nat. Cell. Biol. 2010, 12, 247–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Selcuklu, S.D.; Donoghue, M.T.; Rehmet, K.; de Souza Gomes, M.; Fort, A.; Kovvuru, P.; Muniyappa, M.K.; Kerin, M.J.; Enright, A.J.; Spillane, C. MicroRNA-9 inhibition of cell proliferation and identification of novel miR-9 targets by transcriptome profiling in breast cancer cells. J. Biol. Chem. 2012, 287, 29516–29528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Lujambio, A.; Ropero, S.; Ballestar, E.; Fraga, M.F.; Cerrato, C.; Setien, F.; Casado, S.; Suarez-Gauthier, A.; Sanchez-Cespedes, M.; Git, A.; et al. Genetic unmasking of an epigenetically silenced microRNA in human cancer cells. Cancer Res. 2007, 67, 1424–1429. [Google Scholar] [CrossRef] [Green Version]
  77. Lehmann, U.; Hasemeier, B.; Christgen, M.; Muller, M.; Romermann, D.; Langer, F.; Kreipe, H. Epigenetic inactivation of microRNA gene hsa-mir-9-1 in human breast cancer. J. Pathol. 2008, 214, 17–24. [Google Scholar] [CrossRef]
  78. Bandres, E.; Agirre, X.; Bitarte, N.; Ramirez, N.; Zarate, R.; Roman-Gomez, J.; Prosper, F.; Garcia-Foncillas, J. Epigenetic regulation of microRNA expression in colorectal cancer. Int. J. Cancer 2009, 125, 2737–2743. [Google Scholar] [CrossRef]
  79. Chen, P.; Price, C.; Li, Z.; Li, Y.; Cao, D.; Wiley, A.; He, C.; Gurbuxani, S.; Kunjamma, R.B.; Huang, H.; et al. miR-9 is an essential oncogenic microRNA specifically overexpressed in mixed lineage leukemia-rearranged leukemia. Proc. Natl. Acad. Sci. USA 2013, 110, 11511–11516. [Google Scholar] [CrossRef] [Green Version]
  80. Wang, H.; An, X.; Yu, H.; Zhang, S.; Tang, B.; Zhang, X.; Li, Z. MiR-29b/TET1/ZEB2 signaling axis regulates metastatic properties and epithelial-mesenchymal transition in breast cancer cells. Oncotarget 2017, 8, 102119–102133. [Google Scholar] [CrossRef]
  81. Asmar, F.; Punj, V.; Christensen, J.; Pedersen, M.T.; Pedersen, A.; Nielsen, A.B.; Hother, C.; Ralfkiaer, U.; Brown, P.; Ralfkiaer, E.; et al. Genome-wide profiling identifies a DNA methylation signature that associates with TET2 mutations in diffuse large B-cell lymphoma. Haematologica 2013, 98, 1912–1920. [Google Scholar] [CrossRef] [Green Version]
  82. Chiba, S. Dysregulation of TET2 in hematologic malignancies. Int. J. Hematol. 2017, 105, 17–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Inomata, M.; Tagawa, H.; Guo, Y.M.; Kameoka, Y.; Takahashi, N.; Sawada, K. MicroRNA-17-92 down-regulates expression of distinct targets in different B-cell lymphoma subtypes. Blood 2009, 113, 396–402. [Google Scholar] [CrossRef] [PubMed]
  84. Paik, J.H.; Jang, J.Y.; Jeon, Y.K.; Kim, W.Y.; Kim, T.M.; Heo, D.S.; Kim, C.W. MicroRNA-146a downregulates NFκB activity via targeting TRAF6 and functions as a tumor suppressor having strong prognostic implications in NK/T cell lymphoma. Clin. Cancer Res. 2011, 17, 4761–4771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Hoffman, A.E.; Zheng, T.; Yi, C.; Leaderer, D.; Weidhaas, J.; Slack, F.; Zhang, Y.; Paranjape, T.; Zhu, Y. microRNA miR-196a-2 and breast cancer: A genetic and epigenetic association study and functional analysis. Cancer Res. 2009, 69, 5970–5977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Neves, R.; Scheel, C.; Weinhold, S.; Honisch, E.; Iwaniuk, K.M.; Trompeter, H.I.; Niederacher, D.; Wernet, P.; Santourlidis, S.; Uhrberg, M. Role of DNA methylation in miR-200c/141 cluster silencing in invasive breast cancer cells. BMC Res. Notes. 2010, 3, 7–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Tryndyak, V.P.; Beland, F.A.; Pogribny, I.P. E-cadherin transcriptional down-regulation by epigenetic and microRNA-200 family alterations is related to mesenchymal and drug-resistant phenotypes in human breast cancer cells. Int. J. Cancer 2010, 126, 2575–2583. [Google Scholar] [CrossRef] [PubMed]
  88. Jiménez-Wences, H.; Martínez-Carrillo, D.N.; Peralta-Zaragoza, O.; Campos-Viguri, G.E.; Hernández-Sotelo, D.; Jiménez-López, M.A.; Muñoz-Camacho, J.G.; Garzón-Barrientos, V.H.; Illades-Aguiar, B.; Fernández-Tilapa, G. Methylation and expression of miRNAs in precancerous lesions and cervical cancer with HPV16 infection. Oncol. Rep. 2016, 35, 2297–2305. [Google Scholar] [CrossRef] [Green Version]
  89. Liu, J.; Li, H.; Sun, L.; Shen, S.; Zhou, Q.; Yuan, Y.; Xing, C. Epigenetic Alternations of MicroRNAs and DNA Methylation Contribute to Liver Metastasis of Colorectal Cancer. Dig. Dis. Sci. 2019, 64, 1523–1534. [Google Scholar] [CrossRef]
  90. Deng, H.; Guo, Y.; Song, H.; Xiao, B.; Sun, W.; Liu, Z.; Yu, X.; Xia, T.; Cui, L.; Guo, J. MicroRNA-195 and microRNA-378 mediate tumor growth suppression by epigenetical regulation in gastric cancer. Gene 2013, 518, 351–359. [Google Scholar] [CrossRef]
  91. Wang, H.; Peng, R.; Wang, J.; Qin, Z.; Xue, L. Circulating microRNAs as potential cancer biomarkers: The advantage and disadvantage. Clin. Epigenetics 2018, 10, 1–10. [Google Scholar] [CrossRef] [Green Version]
  92. Tsai, K.W.; Wu, C.W.; Hu, L.Y.; Li, S.C.; Liao, Y.L.; Lai, C.H.; Kao, H.W.; Fang, W.L.; Huang, K.H.; Chan, W.C.; et al. Epigenetic regulation of miR-34b and miR-129 expression in gastric cancer. Int. J. Cancer 2011, 129, 2600–2610. [Google Scholar] [CrossRef]
  93. Alpini, G.; Glaser, S.S.; Zhang, J.P.; Francis, H.; Han, Y.; Gong, J.; Stokes, A.; Francis, T.; Hughart, N.; Hubble, L.; et al. Regulation of placenta growth factor by microRNA-125b in hepatocellular cancer. J. Hepatol. 2011, 55, 1339–1345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Tivnan, A.; Zhao, J.; Johns, T.G.; Day, B.W.; Stringer, B.W.; Boyd, A.W.; Tiwari, S.; Giles, K.M.; Teo, C.; McDonald, K.L. The tumor suppressor microRNA, miR-124a, is regulated by epigenetic silencing and by the transcriptional factor, REST in glioblastoma. Tumor. Biol. 2014, 35, 1459–1465. [Google Scholar] [CrossRef] [PubMed]
  95. Braconi, C.; Kogure, T.; Valeri, N.; Huang, N.; Nuovo, G.; Costinean, S.; Negrini, M.; Miotto, E.; Croce, C.M.; Patel, T. MicroRNA-29 can regulate expression of the long non-coding RNA gene MEG3 in hepatocellular cancer. Oncogene 2011, 30, 4750–4756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ng, H.Y.; Wan, T.S.; So, C.C.; Chim, C.S. Epigenetic inactivation of DAPK1, p14ARF, mir-34a and -34b/c in acute promyelocytic leukaemia. J. Clin. Pathol. 2014, 67, 626–631. [Google Scholar] [CrossRef] [PubMed]
  97. Chim, C.S.; Wong, K.Y.; Leung, C.Y.; Chung, L.P.; Hui, P.K.; Chan, S.Y.; Yu, L. Epigenetic inactivation of the hsa-miR-203 in haematological malignancies. J. Cell. Mol. Med. 2011, 15, 2760–2767. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Wong, K.Y.; So, C.C.; Loong, F.; Chung, L.P.; Lam, W.W.L.; Liang, R.; Li, G.K.H.; Jin, D.Y.; Chim, C.S. Epigenetic inactivation of the miR-124-1 in haematological malignancies. PLoS ONE 2011, 6. [Google Scholar] [CrossRef] [Green Version]
  99. Wong, K.Y.; Yim, R.L.H.; Kwong, Y.L.; Leung, C.Y.; Hui, P.K.; Cheung, F.; Liang, R.; Jin, D.Y.; Chim, C.S. Epigenetic inactivation of the MIR129-2 in hematological malignancies. J. Hematol. Oncol. 2013, 6, 1–11. [Google Scholar] [CrossRef] [Green Version]
  100. Rusek, M.M.; Abba, M.; Eljaszewicz, A.; Moniuszko, M.; Niklinski, J.; Allgayer, H. MicroRNA modulators of epigenetic regulation, the tumor microenvironment and the immune system in lung cancer. Mol. Cancer 2015, 14, 1–10. [Google Scholar] [CrossRef] [Green Version]
  101. Mazar, J.; DeBlasio, D.; Govindarajan, S.S.; Zhang, S.; Perera, R.J. Epigenetic regulation of microRNA-375 and its role in melanoma development in humans. FEBS Lett. 2011, 585, 2467–2476. [Google Scholar] [CrossRef] [Green Version]
  102. Azizi, M.; Teimoori-Toolabi, L.; Arzanani, M.K.; Azadmanesh, K.; Fard-Esfahani, P.; Zeinali, S. MicroRNA-148b and microRNA-152 reactivate tumor suppressor genes through suppression of DNA methyltransferase-1 gene in pancreatic cancer cell lines. Cancer Biol. Ther. 2014, 15, 419–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Hulf, T.; Sibbritt, T.; Wiklund, E.D.; Patterson, K.; Song, J.Z.; Stirzaker, C.; Qu, W.; Nair, S.; Horvath, L.G.; Armstrong, N.J.; et al. Epigenetic-induced repression of microRNA-205 is associated with MED1 activation and a poorer prognosis in localized prostate cancer. Oncogene 2013, 32, 2891–2899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Vrba, L.; Garbe, J.C.; Stampfer, M.R.; Futscher, B.W. Epigenetic regulation of normal human mammary cell type-specific miRNAs. Genome Res. 2011, 21, 2026–2037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Macdonald, S.; Oncology, R.; General, M. Breast Cancer Breast Cancer. J. R. Soc. Med. 2016, 70, 515–517. [Google Scholar] [CrossRef]
  106. Veeck, J.; Esteller, M. Breast cancer epigenetics: From DNA methylation to microRNAs. J. Mammary Gland Biol. Neoplasia 2010, 15, 5–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Pinto, R.; Pilato, B.; Ottini, L.; Lambo, R.; Simone, G.; Paradiso, A.; Tommasi, S. Different methylation and MicroRNA expression pattern in male and female familial breast cancer. J. Cell. Physiol. 2013, 228, 1264–1269. [Google Scholar] [CrossRef] [PubMed]
  108. Duru, N.; Gernapudi, R.; Eades, G.; Eckert, R.; Zhou, Q. Epigenetic Regulation of miRNAs and Breast Cancer Stem Cells. Curr. Pharmacol. Rep. 2015, 1, 161–169. [Google Scholar] [CrossRef] [PubMed]
  109. Rocha Simonini, P.D.S.; Breiling, A.; Gupta, N.; Malekpour, M.; Youns, M.; Omranipour, R.; Malekpour, F.; Volinia, S.; Croce, C.M.; Najmabadi, H.; et al. Epigenetically deregulated microRNA-375 is involved in a positive feedback loop with estrogen receptor α in breast cancer cells. Cancer Res. 2010, 70, 9175–9184. [Google Scholar] [CrossRef] [Green Version]
  110. de Brot, S.; Rutland, C.S.; Mongan, N.P.; James, V. Epigenetic Control. of MicroRNA Expression and Cancer; Elsevier Inc.: Atlanta, GA, USA, 2018; ISBN 9780128110225. [Google Scholar]
  111. Li, Y.; Kong, D.; Ahmad, A.; Bao, B.; Dyson, G.; Sarkar, F.H. Epigenetic deregulation of miR-29a and miR-1256 by isoflavone contributes to the inhibition of prostate cancer cell growth and invasion. Epigenetics 2012, 7, 940–949. [Google Scholar] [CrossRef] [Green Version]
  112. Li, Y.; Sarkar, F.H. Targeting Epigenetically Deregulated miRNA by Nutraceuticals: Focusing on Cancer Prevention and Treatment. Curr. Pharmacol. Rep. 2015, 1, 1–10. [Google Scholar] [CrossRef] [Green Version]
  113. Guil, S.; Esteller, M. DNA methylomes, histone codes and miRNAs: Tying it all together. Int. J. Biochem. Cell. Biol. 2009, 41, 87–95. [Google Scholar] [CrossRef] [PubMed]
  114. Sheervalilou, R.; Shirvaliloo, S.; Aval, S.F.; Khamaneh, A.M.; Sharifi, A.; Ansarin, K.; Zarghami, N. A new insight on reciprocal relationship between microRNA expression and epigenetic modifications in human lung cancer. Tumor. Biol. 2017, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. He-Yuan Qi; Zhao-Jun Zhang; Ya-Juan Li; Xian-Dong Fang No Title. Hereditas 2011, 33, 1291–1299.
  116. Li, B.; Carey, M.; Workman, J.L. The Role of Chromatin during Transcription. Cell 2007, 128, 707–719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Barski, A.; Cuddapah, S.; Cui, K.; Roh, T.Y.; Schones, D.E.; Wang, Z.; Wei, G.; Chepelev, I.; Zhao, K. High-resolution profiling of histone methylations in the human genome. Cell 2007, 129, 823–837. [Google Scholar] [CrossRef] [Green Version]
  118. Tao, B.B.; Liu, X.Q.; Zhang, W.; Li, S.; Dong, D.; Xiao, M.; Zhong, J. Evidence for the association of chromatin and microRNA regulation in the human genome. Oncotarget 2017, 8, 70958–70966. [Google Scholar] [CrossRef]
  119. Linares, A.; Dalenc, F.; Balaguer, P.; Boulle, N.; Cavailles, V. Manipulating protein acetylation in breast cancer: A promising approach in combination with hormonal therapies? J. Biomed. Biotechnol. 2011, 2011, 856985. [Google Scholar] [CrossRef] [Green Version]
  120. Tessarz, P.; Kouzarides, T. Histone core modifications regulating nucleosome structure and dynamics. Nat. Rev. Mol. Cell. Biol. 2014, 15, 703–708. [Google Scholar] [CrossRef]
  121. Zentner, G.E.; Henikoff, S. Regulation of nucleosome dynamics by histone modifications. Nat. Struct. Mol. Biol. 2013, 20, 259–266. [Google Scholar] [CrossRef]
  122. Shen, T.; Sanchez, H.N.; Zan, H.; Casali, P. Genome-Wide Analysis Reveals Selective Modulation of microRNAs and mRNAs by Histone Deacetylase Inhibitor in B Cells Induced to Undergo Class-Switch DNA Recombination and Plasma Cell Differentiation. Front. Immunol. 2015, 6, 627. [Google Scholar] [CrossRef] [Green Version]
  123. Bamodu, O.A.; Huang, W.C.; Lee, W.H.; Wu, A.; Wang, L.S.; Hsiao, M.; Yeh, C.T.; Chao, T.Y. Aberrant KDM5B expression promotes aggressive breast cancer through MALAT1 overexpression and downregulation of hsa-miR-448. BMC Cancer 2016, 16, 160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Wang, S.; Wu, W.; Claret, F.X. Mutual regulation of microRNAs and DNA methylation in human cancers. Epigenetics 2017, 12, 187–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Chuang, J.C.; Jones, P.A. Epigenetics and microRNAs. Pediatr. Res. 2007, 61, 24–29. [Google Scholar] [CrossRef]
  126. Nowacka-Zawisza, M.; Wiśnik, E. DNA methylation and histone modifications as epigenetic regulation in prostate cancer (review). Oncol. Rep. 2017, 38, 2587–2596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Akbari Moqadam, F.; Pieters, R.; Den Boer, M.L. The hunting of targets: Challenge in miRNA research. Leukemia 2013, 27, 16–23. [Google Scholar] [CrossRef]
  128. Tarang, S.; Weston, M.D. Macros in microRNA target identification: A comparative analysis of in silico, in vitro, and in vivo approaches to microRNA target identification. RNA Biol. 2014, 11, 324–333. [Google Scholar] [CrossRef] [Green Version]
  129. Carroll, A.P.; Goodall, G.J.; Liu, B. Understanding principles of miRNA target recognition and function through integrated biological and bioinformatics approaches. Wiley Interdiscip. Rev. RNA 2014, 5, 361–379. [Google Scholar] [CrossRef]
  130. Akhtar, M.M.; Micolucci, L.; Islam, M.S.; Olivieri, F.; Procopio, A.D. Bioinformatic tools for microRNA dissection. Nucleic Acids Res. 2016, 44, 24–44. [Google Scholar] [CrossRef] [Green Version]
  131. Oliveira, A.C.; Bovolenta, L.A.; Nachtigall, P.G.; Herkenhoff, M.E.; Lemke, N.; Pinhal, D. Combining results from distinct microRNA target prediction tools enhances the performance of analyses. Front. Genet. 2017, 8, 1–10. [Google Scholar] [CrossRef] [Green Version]
  132. Huang, Z.; Shi, J.; Gao, Y.; Cui, C.; Zhang, S.; Li, J.; Zhou, Y.; Cui, Q. HMDD v3.0: A database for experimentally supported human microRNA-disease associations. Nucleic Acids Res. 2019, 47, D1013–D1017. [Google Scholar] [CrossRef] [Green Version]
  133. Xie, B.; Ding, Q.; Han, H.; Wu, D. MiRCancer: A microRNA-cancer association database constructed by text mining on literature. Bioinformatics 2013, 29, 638–644. [Google Scholar] [CrossRef] [PubMed]
  134. He, X.; Chang, S.; Zhang, J.; Zhao, Q.; Xiang, H.; Kusonmano, K.; Yang, L.; Sun, Z.S.; Yang, H.; Wang, J. MethyCancer: The database of human DNA methylation and cancer. Nucleic Acids Res. 2008, 36, 836–841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Mi, S.; Lu, J.; Sun, M.; Li, Z.; Zhang, H.; Neilly, M.B.; Wang, Y.; Qian, Z.; Jin, J.; Zhang, Y.; et al. MicroRNA expression signatures accurately discriminate acute lymphoblastic leukemia from acute myeloid leukemia. Proc. Natl. Acad. Sci. USA 2007, 104, 19971–19976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Schotte, D.; Chau, J.C.; Sylvester, G.; Liu, G.; Chen, C.; van der Velden, V.H.; Broekhuis, M.J.; Peters, T.C.; Pieters, R.; den Boer, M.L. Identification of new microRNA genes and aberrant microRNA profiles in childhood acute lymphoblastic leukemia. Leukemia 2009, 23, 313–322. [Google Scholar] [CrossRef] [PubMed]
  137. Schotte, D.; De Menezes, R.X.; Akbari Moqadam, F.; Khankahdani, L.M.; Lange-Turenhout, E.; Chen, C.; Pieters, R.; Den Boer, M.L. MicroRNA characterize genetic diversity and drug resistance in pediatric acute lymphoblastic leukemia. Haematologica 2011, 96, 703–711. [Google Scholar] [CrossRef] [Green Version]
  138. Martin-Palanco, V.; Rodriguez, G.; Martin, C.; Rojas, R.; Torres, A.; Roman-Gomez, J. microRNA methylation profile has prognosis impact in acute lymphoblastic leukemia patients undergoing stem cell transplantation. Biol. Blood Marrow Transpl. 2011, 17, 745–748. [Google Scholar] [CrossRef] [Green Version]
  139. Esau, C.C.; Monia, B.P. Therapeutic potential for microRNAs. Adv. Drug Deliv. Rev. 2007, 59, 101–114. [Google Scholar] [CrossRef]
  140. Wang, V.; Wu, W. MicroRNA-based therapeutics for cancer. BioDrugs 2009, 23, 15–23. [Google Scholar] [CrossRef]
  141. Janssen, E.A.; Slewa, A.; Gudlaugsson, E.; Jonsdottir, K.; Skaland, I.; Soiland, H.; Baak, J.P. Biologic profiling of lymph node negative breast cancers by means of microRNA expression. Mod. Pathol. 2010, 23, 1567–1576. [Google Scholar] [CrossRef] [Green Version]
  142. Rio-Machin, A.; Ferreira, B.I.; Henry, T.; Gomez-Lopez, G.; Agirre, X.; Alvarez, S.; Rodriguez-Perales, S.; Prosper, F.; Calasanz, M.J.; Martinez, J.; et al. Downregulation of specific miRNAs in hyperdiploid multiple myeloma mimics the oncogenic effect of IgH translocations occurring in the non-hyperdiploid subtype. Leukemia 2013, 27, 925–931. [Google Scholar] [CrossRef] [Green Version]
  143. Otoukesh, B.; Abbasi, M.; Gorgani, H.O.; Farahini, H.; Moghtadaei, M.; Boddouhi, B.; Kaghazian, P.; Hosseinzadeh, S.; Alaee, A. MicroRNAs signatures, bioinformatics analysis of miRNAs, miRNA mimics and antagonists, and miRNA therapeutics in osteosarcoma. Cancer Cell Int. 2020, 20, 254. [Google Scholar] [CrossRef] [PubMed]
  144. Chen, Y.; Gao, D.Y.; Huang, L. In vivo delivery of miRNAs for cancer therapy: Challenges and strategies. Adv. Drug Deliv. Rev. 2015, 81, 128–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Luo, Q.; Gao, Y.; Long, Y.; Chen, L.; Huang, Y.; Fang, L. MiRNA-497 regulates cell growth and invasion by targeting cyclin E1 in breast cancer. Cancer Cell Int. 2013, 13, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Wang, Y.; Sun, H.H.; Sui, M.H.; Ma, J.J. miR-218 inhibits acute promyelocytic leukemia cell growth by targeting BMI-1. Oncol. Lett. 2017, 14, 8078–8083. [Google Scholar] [CrossRef] [PubMed]
  147. Dorrance, A.M.; Neviani, P.; Ferenchak, G.J.; Huang, X.; Nicolet, D.; Maharry, K.S.; Ozer, H.G.; Hoellarbauer, P.; Khalife, J.; Hill, E.B.; et al. Targeting leukemia stem cells in vivo with antagomiR-126 nanoparticles in acute myeloid leukemia. Leukemia 2015, 29, 2143–2153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Ors-Kumoglu, G.; Gulce-Iz, S.; Biray-Avci, C. Therapeutic microRNAs in human cancer. Cytotechnology 2019, 71, 411–425. [Google Scholar] [CrossRef] [PubMed]
  149. Mercatelli, N.; Fortini, D.; Palombo, R.; Paronetto, M.P. Small molecule inhibition of Ewing sarcoma cell growth via targeting the long non coding RNA HULC. Cancer Lett. 2020, 469, 111–123. [Google Scholar] [CrossRef]
  150. Jung, J.; Yeom, C.; Choi, Y.S.; Kim, S.; Lee, E.J.; Park, M.J.; Kang, S.W.; Kim, S.B.; Chang, S. Simultaneous inhibition of multiple oncogenic miRNAs by a multi-potent microRNA sponge. Oncotarget 2015, 6, 20370–20387. [Google Scholar] [CrossRef]
  151. Zhang, T.; Hu, Y.; Ju, J.; Hou, L.; Li, Z.; Xiao, D.; Li, Y.; Yao, J.; Wang, C.; Zhang, Y.; et al. Downregulation of miR-522 suppresses proliferation and metastasis of non-small cell lung cancer cells by directly targeting DENN/MADD domain containing 2D. Sci. Rep. 2016, 6, 1–12. [Google Scholar] [CrossRef] [Green Version]
  152. Xiao, J.; Yang, B.; Lin, H.; Lu, Y.; Luo, X.; Wang, Z. Novel approaches for gene-specific interference via manipulating actions of microRNAs: Examination on the pacemaker channel genes HCN2 and HCN4. J. Cell. Physiol. 2007, 212, 285–292. [Google Scholar] [CrossRef]
  153. Saito, Y.; Liang, G.; Egger, G.; Friedman, J.M.; Chuang, J.C.; Coetzee, G.A.; Jones, P.A. Specific activation of microRNA-127 with downregulation of the proto-oncogene BCL6 by chromatin-modifying drugs in human cancer cells. Cancer Cell 2006, 9, 435–443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Saldivar-Gonzalez, F.I.; Gomez-Garcia, A.; Chavez-Ponce de Leon, D.E.; Sanchez-Cruz, N.; Ruiz-Rios, J.; Pilon-Jimenez, B.A.; Medina-Franco, J.L. Inhibitors of DNA Methyltransferases From Natural Sources: A Computational Perspective. Front. Pharmacol. 2018, 9, 1144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. West, A.C.; Johnstone, R.W.; West, A.C.; Johnstone, R.W. New and emerging HDAC inhibitors for cancer treatment Find the latest version: Review series New and emerging HDAC inhibitors for cancer treatment. J. Clin. Investig. 2014, 124, 30–39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Tsai, H.C.; Li, H.; Van Neste, L.; Cai, Y.; Robert, C.; Rassool, F.V.; Shin, J.J.; Harbom, K.M.; Beaty, R.; Pappou, E.; et al. Transient Low Doses of DNA-Demethylating Agents Exert Durable Antitumor Effects on Hematological and Epithelial Tumor Cells. Cancer Cell 2012, 21, 430–446. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Agrawal, K.; Das, V.; Vyas, P.; Hajduch, M. Nucleosidic DNA demethylating epigenetic drugs—A comprehensive review from discovery to clinic. Pharmacol. Ther. 2018, 188, 45–79. [Google Scholar] [CrossRef]
  158. Azad, N.; Zahnow, C.A.; Rudin, C.M.; Baylin, S.B. The future of epigenetic therapy in solid tumours—Lessons from the past. Nat. Rev. Clin. Oncol. 2013, 10, 256–266. [Google Scholar] [CrossRef]
  159. Reid, G.; Kao, S.C.; Pavlakis, N.; Brahmbhatt, H.; MacDiarmid, J.; Clarke, S.; Boyer, M.; van Zandwijk, N. Clinical development of TargomiRs, a miRNA mimic-based treatment for patients with recurrent thoracic cancer. Epigenomics 2016, 8, 1079–1085. [Google Scholar] [CrossRef] [Green Version]
  160. van Zandwijk, N.; Pavlakis, N.; Kao, S.C.; Linton, A.; Boyer, M.J.; Clarke, S.; Huynh, Y.; Chrzanowska, A.; Fulham, M.J.; Bailey, D.L.; et al. Safety and activity of microRNA-loaded minicells in patients with recurrent malignant pleural mesothelioma: A first-in-man, phase 1, open-label, dose-escalation study. Lancet. Oncol. 2017, 18, 1386–1396. [Google Scholar] [CrossRef]
  161. de Gooijer, C.J.; Baas, P.; Burgers, J.A. Current chemotherapy strategies in malignant pleural mesothelioma. Transl. Lung Cancer Res. 2018, 7, 574–583. [Google Scholar] [CrossRef]
  162. National Institute of Health. U.S National Library of Medicine Clinical Trials. Available online: https://clinicaltrials.gov/ct2 (accessed on 10 August 2020).
  163. Huang, Z.; Zhang, L.; Zhu, D.; Shan, X.; Zhou, X.; Qi, L.W.; Wu, L.; Zhu, J.; Cheng, W.; Zhang, H.; et al. A novel serum microRNA signature to screen esophageal squamous cell carcinoma. Cancer Med. 2017, 6, 109–119. [Google Scholar] [CrossRef]
  164. Zhou, X.; Wen, W.; Shan, X.; Zhu, W.; Xu, J.; Guo, R.; Cheng, W.; Wang, F.; Qi, L.W.; Chen, Y.; et al. A six-microRNA panel in plasma was identified as a potential biomarker for lung adenocarcinoma diagnosis. Oncotarget 2017, 8, 6513–6525. [Google Scholar] [CrossRef] [PubMed]
  165. Monzo, M.; Santasusagna, S.; Moreno, I.; Martinez, F.; Hernández, R.; Muñoz, C.; Castellano, J.J.; Moreno, J.; Navarro, A. Exosomal microRNAs isolated from plasma of mesenteric veins linked to liver metastases in resected patients with colon cancer. Oncotarget 2017, 8, 30859–30869. [Google Scholar] [CrossRef] [PubMed]
  166. Dirin, M.; Winkler, J. Influence of diverse chemical modifications on the ADME characteristics and toxicology of antisense oligonucleotides. Expert. Opin. Biol. Ther. 2013, 13, 875–888. [Google Scholar] [CrossRef] [PubMed]
  167. Velpurisiva, P.; Gad, A.; Piel, B.; Jadia, R.; Rai, P. Nanoparticle Design Strategies for Effective Cancer Immunotherapy. J. Biomed. 2017, 2, 64–77. [Google Scholar] [CrossRef] [Green Version]
  168. Bai, Z.; Wei, J.; Yu, C.; Han, X.; Qin, X.; Zhang, C.; Liao, W.; Li, L.; Huang, W. Non-viral nanocarriers for intracellular delivery of microRNA therapeutics. J. Mater. Chem. B 2019, 7, 1209–1225. [Google Scholar] [CrossRef]
  169. Zhang, Y.; Wang, Z.; Gemeinhart, R.A. Progress in microRNA delivery. J. Control. Release 2013, 172, 962–974. [Google Scholar] [CrossRef] [Green Version]
  170. Mercatelli, N.; Coppola, V.; Bonci, D.; Miele, F.; Costantini, A.; Guadagnoli, M.; Bonanno, E.; Muto, G.; Frajese, G.V.; De Maria, R.; et al. The inhibition of the highly expressed miR-221 and miR-222 impairs the growth of prostate carcinoma xenografts in mice. PLoS ONE 2008, 3, e4029. [Google Scholar] [CrossRef]
  171. Xu, Y.; So, C.; Lam, H.M.; Fung, M.C.; Tsang, S.Y. Apoptosis Reversal Promotes Cancer Stem Cell-Like Cell Formation. Neoplasia 2018, 20, 295–303. [Google Scholar] [CrossRef]
  172. Coward, J.; Harding, A. Size does matter: Why polyploid tumor cells are critical drug targets in the war on cancer. Front. Oncol. 2014, 4, 1–15. [Google Scholar] [CrossRef]
  173. Liu, J. The dualistic origin of human tumors. Semin. Cancer Biol. 2018, 53, 1–16. [Google Scholar] [CrossRef]
  174. Zhang, L.; Ding, P.; Lv, H.; Zhang, D.; Liu, G.; Yang, Z.; Li, Y.; Liu, J.; Zhang, S. Number of polyploid giant cancer cells and expression of EZH2 are associated with VM formation and tumor grade in human ovarian tumor. BioMed Res. Int. 2014, 2014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Mirzayans, R.; Andrais, B.; Murray, D. Roles of polyploid/multinucleated giant cancer cells in metastasis and disease relapse following anticancer treatment. Cancers 2018, 10, 118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Kang, K.A.; Piao, M.J.; Kim, K.C.; Kang, H.K.; Chang, W.Y.; Park, I.C.; Keum, Y.S.; Surh, Y.J.; Hyun, J.W. Epigenetic modification of Nrf2 in 5-fluorouracil-resistant colon cancer cells: Involvement of TET-dependent DNA demethylation. Cell Death Dis. 2014, 5, 1–13. [Google Scholar] [CrossRef] [PubMed]
Figure 1. miRNA epigenetic crosstalk. Association of miRNAs with epigenetic regulators involved in the processes of DNA methylation and histone modification. miRNA-mediated alteration of these regulators causes aberrant DNA methylation and chromatin modification. These distorted conditions modify the expression of genes that are involved in modulating the epigenetic machinery and can also affect miRNA expression.
Figure 1. miRNA epigenetic crosstalk. Association of miRNAs with epigenetic regulators involved in the processes of DNA methylation and histone modification. miRNA-mediated alteration of these regulators causes aberrant DNA methylation and chromatin modification. These distorted conditions modify the expression of genes that are involved in modulating the epigenetic machinery and can also affect miRNA expression.
Cancers 12 02922 g001
Figure 2. A miRNA-epigenetic network analysis. The diagram represents a network analysis outcome conducted in miRNet suite among 555 miRNAs with 48 epigenetic modifiers hosted by the tool itself. Among those modifiers, DNA methyltransferases (DNMTs) and histone deacetylases (HDACs) showed a strong connection with miRNAs and other modifiers. The enlisted 48 modifiers used were as follows: AML1-ETO oncoprotein/HDAC1/DNMTs/MeCP2, AML1/ETO protein, CARM1, CBP, DNMT-1, DNMT-1/DNMT-3a/DNMT-3b, DNMT-1/DNMT-3b, DNMT-1/DNMT-3B, DNMT-1/MeCP2, DNMT-1/MLL, DNMT-3a, DNMT-3a/DNMT-3b, DNMT-3a/DNMT-3b/DNMT-1, DNMT-3b, DNMTs, Egr2/Jardi1b, EVI1/DNMT-3b, EZH2, EZH2/G9a/HDAC, HDAC1, HDAC1/AP-1, HDAC1/HDAC2, HDAC1/HDAC2/HDAC4, HDAC1/HDAC3/EP300, HDAC3/Myc, HDAC4, HDAC4/HDAC5, HDAC4/SP1, HDACs, HER2, IL-6, JARID1B, Kindlin 2/DNMT-3, MBD1, MBD2, MeCP2, SUZ12/BMI1, Mel-18/DNMT-1, MLL fusion proteins, MYC/HDAC3/EZH2, MYST3, p50 p53, PRC2, PRMT5, RNAPII, sp1, and SUV39H1.
Figure 2. A miRNA-epigenetic network analysis. The diagram represents a network analysis outcome conducted in miRNet suite among 555 miRNAs with 48 epigenetic modifiers hosted by the tool itself. Among those modifiers, DNA methyltransferases (DNMTs) and histone deacetylases (HDACs) showed a strong connection with miRNAs and other modifiers. The enlisted 48 modifiers used were as follows: AML1-ETO oncoprotein/HDAC1/DNMTs/MeCP2, AML1/ETO protein, CARM1, CBP, DNMT-1, DNMT-1/DNMT-3a/DNMT-3b, DNMT-1/DNMT-3b, DNMT-1/DNMT-3B, DNMT-1/MeCP2, DNMT-1/MLL, DNMT-3a, DNMT-3a/DNMT-3b, DNMT-3a/DNMT-3b/DNMT-1, DNMT-3b, DNMTs, Egr2/Jardi1b, EVI1/DNMT-3b, EZH2, EZH2/G9a/HDAC, HDAC1, HDAC1/AP-1, HDAC1/HDAC2, HDAC1/HDAC2/HDAC4, HDAC1/HDAC3/EP300, HDAC3/Myc, HDAC4, HDAC4/HDAC5, HDAC4/SP1, HDACs, HER2, IL-6, JARID1B, Kindlin 2/DNMT-3, MBD1, MBD2, MeCP2, SUZ12/BMI1, Mel-18/DNMT-1, MLL fusion proteins, MYC/HDAC3/EZH2, MYST3, p50 p53, PRC2, PRMT5, RNAPII, sp1, and SUV39H1.
Cancers 12 02922 g002
Figure 3. An overview of the therapeutic application of miRNA-based agents used in the prevention of tumour progression. These agents target and regulate the miRNA of interest, resulting in either inhibition of an OncomiR or upregulation of a tumour suppressor gene. Five commonly utilised therapeutic miRNA agents are displayed above showing (A) anti-miRs, miRNA sponges and miRNA masks used to suppress oncomiRs by inhibiting miRNA binding to mRNA. miRNA mimics and molecule inhibitor drugs can be used to inhibit DNMT enzymatic activity and trigger its degradation, thereby reducing oncomiR expression and development of malignancy. Similarly, histone deacetylase inhibitors (HDACis) can be used to block tumorigenesis. (B) miRNA mimics can also be utilised to mimic the activity of tumour suppressor miRNA in malignancies to reduce or suppress tumorigenesis.
Figure 3. An overview of the therapeutic application of miRNA-based agents used in the prevention of tumour progression. These agents target and regulate the miRNA of interest, resulting in either inhibition of an OncomiR or upregulation of a tumour suppressor gene. Five commonly utilised therapeutic miRNA agents are displayed above showing (A) anti-miRs, miRNA sponges and miRNA masks used to suppress oncomiRs by inhibiting miRNA binding to mRNA. miRNA mimics and molecule inhibitor drugs can be used to inhibit DNMT enzymatic activity and trigger its degradation, thereby reducing oncomiR expression and development of malignancy. Similarly, histone deacetylase inhibitors (HDACis) can be used to block tumorigenesis. (B) miRNA mimics can also be utilised to mimic the activity of tumour suppressor miRNA in malignancies to reduce or suppress tumorigenesis.
Cancers 12 02922 g003
Table 1. miRNAs regulate epigenetic modifications in cancer.
Table 1. miRNAs regulate epigenetic modifications in cancer.
CancermiRNAExpressionEpi Regulator Epi ModificationEpi TargetRef
Acquired resistance of breast cancermiR-29a
miR-29b-3p
HighDNMT-3aDNA Methylationglobal DNA methylation[36]
miR-132HighMeCP2DNA Methylationglobal DNA methylation
Acute myeloid leukaemiamiR-29b-3pHighDNMT-3a/DNMT-3b/DNMT-1DNA MethylationESR1/cyclin-dependent kinase inhibitor 2B[37]
Bladder transitional cell carcinomamiR-101-3pLowEZH2H3K27me3-[38]
Breast cancermiR-10aLow-H3K27me3/DNA MethylationHOXD4[39]
miR-29c
miR-26b
miR-148b
HighDNMT-3bDNA MethylationCEACAM6/CST6/SCNN1A[40]
miR-148aHighDNMT-1DNA MethylationmiR-148a[41]
miR-152HighDNMT-1DNA MethylationmiR-152
Breast/ProstatemiR-101-3pHighEZH2H3K27me3miR-101[42]
Colorectal cancermiR-143LowDNMT-3aDNA Methylation-[43]
miR-342HighDNMT-1DNA MethylationADAM23/Hint1/RASSF1A/RECK[44]
Cutaneous melanomamiR-29a-3pHighDNMT-3a/DNMT-3bDNA MethylationRASSF1A/TFPI-2/RAR-/SOCS/GATA4[45]
Endometrial cancermiR-152LowDNMT-1DNA Methylation-[46]
Gastric cancermiR-148aHighDNMT-1DNA MethylationmiR-148a[47]
GlioblastomamiR-101-3pLowEZH2H3K27me3-[48]
miR-152HighDNMT-1DNA MethylationmiR-152[49]
GliomamiR-185HighDNMT-1DNA MethylationANKDD1A/GAD1/HIST1H3E/PCDHA8/PCDHA13/PHOX2B/SIX3/SST[50]
Hepatitis B virus (HBV)-related hepatocellular carcinomamiR-101-3pHighDNMT-3aDNA MethylationRASSF1/PRDM2/GSTP1/RUNX3[51]
Hepatocellular carcinomamiR-200aHighHDAC4H3acmiR-200a[52]
miR-125b-5pHighSUV39H1H3K9me3Ki67[53]
miR-152HighDNMT-1DNA MethylationGSTP1/CDH1[54]
Hilar cholangiocarcinomamiR-373HighMBD2DNA MethylationMBD2 regulate RASSF1A[55]
HighMBD2DNA MethylationmiR-373[56]
Human malignant cholangiocytesmiR-148a
miR-152
HighDNMT-1DNA MethylationRassf1a/p16INK4a[57]
Leukaemia and lymphoma cellsmiR-19a
miR-25
miR-32
miR-92b
miR-96
HighPRMT5DNA MethylationH3R8/H4R3[58]
Lung cancermiR-101-1HighEZH2H3K27me3CDH1[59]
MelanomamiR-31HighEZH2DNA MethylationmiR-31[60]
Multiple myelomamiR-29b-3pLowDNMT-3a/DNMT-3bDNA Methylation-[61]
Non-small-cell lung cancermiR-29a
miR-29b-3p
miR-29c
HighDNMT-3a/DNMT-3bDNA MethylationFHIT/WWOX[62]
miR-29b-3pHighDNMT-1/DNMT-3a/DNMT-3bDNA MethylationPTEN[63]
miR-29b-3pHighDNMT-3bDNA MethylationCADM1/RASSF1/FHIT[64]
Ovarian cancermiR-185
miR-152
LowDNMT-1DNA Methylation-[65]
Prostate cancermiR-34bHighHDAC1/HDAC2/HDAC4H3K4me3miR-34b[66]
DNMT-1DNA Methylation
miR-449aLowHDAC1Histone Acetylation-[67]
miR-101-3pLowEZH2H3K27me3-[68]
Squamous cell carcinomamiR-138-5pHighEZH2H3K27me3E-cad[69]
T-cell leukaemiamiR-101-3p
miR-128
LowEZH2H3K27me3-[70]
Table 2. Epi-miRNA databases and networking tools.
Table 2. Epi-miRNA databases and networking tools.
DatabaseFunctionsLinkReference
EpimiRContains 1974 regulatory relationships between 19 different epigenetic modifications and 617 miRNAs across Homo sapiens and 6 more species. The records are divided into two sections: Epi2miR and miR2Epi. http://www.jianglab.cn/EpimiR/[24]
HMDD The Human microRNA Disease Database (HMDD) collects curated experiment-supported evidence for disease-associated human miRNAs classified into 6 evidence classes (genetics, epigenetics, target, circulation, tissue and other) and 20 evidence codes. It also provides a disease-associated miRNA-target network visualisation function.https://www.cuilab.cn/hmdd[132]
miRCancerThe database presently records 878 interactions between 236 miRNAs and 79 human cancers through the processing of more than 26,000 published articles. http://mircancer.ecu.edu/browse.jsp[133]
MethyCancerMethyCancer introduces highly integrated DNA methylation data, cancer-related gene, mutation and cancer evidence from numerous resources, and the CpG island (CGI) clones derived from large-scale sequencing.http://methycancer.psych.ac.cn/[134]
miRNetmiRNet enables statistical analysis and functional interpretation of a variety of data produced from existing miRNA studies. The key features include: (i) integration of high-quality miRNA-target interaction data from 11 databases; (ii) differential expression analysis of data from microarray, RNA-seq and quantitative PCR; (iii) flexible options for data filtering, refinement and customisation during network creation; and (iv) a network visualisation system coupled with enrichment analysis.https://www.mirnet.ca/miRNet/home.xhtml[25]
Table 3. miRNet epi-miRNA network analysis outcome.
Table 3. miRNet epi-miRNA network analysis outcome.
miRNAHitsp-ValueAdj p-Value
mir-291400
mir-515581.275 × 10−92.55 × 10−11
mir-17201.786667 × 10−85.36 × 10−10
mir-2690.000042750.00000171
mir-30130.000240.000012
mir-25100.000270.0000162
mir-1980.00077285710.0000541
mir-19360.00181250.000145
mir-29070.0069454550.000764
mir-970.0069454550.000764
mir-3470.0069454550.000764
mir-19660.021916670.00263
mir-66450.022071430.00309
mir-12950.022071430.00309
mir-10120.022266670.00334
mir-1580.035588240.0101
mir-10140.035588240.0113
mir-122830.035588240.0121
mir-45430.035588240.0121
mir-43230.035588240.0121
mir-48330.035588240.0121
mir-42530.035588240.0121
mir-32230.035588240.0121
mir-37330.035588240.0121
mir-12730.035588240.0121
mir-12630.035588240.0121
mir-19130.035588240.0121
mir-14230.035588240.0121
mir-14030.035588240.0121
mir-18330.035588240.0121
mir-13930.035588240.0121
mir-3130.035588240.0121
mir-2230.035588240.0121
mir-2130.035588240.0121
mir-2850.054285710.019
mir-36540.066428570.0279
mir-14640.066428570.0279
mir-22140.066428570.0279
mir-21840.066428570.0279
mir-13240.066428570.0279
mir-2740.066428570.0279
mir-2440.066428570.0279
mir-12450.080697670.0347
mir-66330.091363640.0402
let-7130.11333330.051
mir-1260a20.11478260.0528
mir-13540.18659570.0877
mir-13050.23508770.119
mir-19940.23508770.13
mir-87620.23508770.134
mir-54220.23508770.134
mir-57420.23508770.134
mir-20220.23508770.134
mir-32420.23508770.134
mir-34220.23508770.134
mir-33020.23508770.134
mir-36120.23508770.134
mir-850.34827590.202
mir-54420.380.228
mir-14720.380.228
mir-10340.38196720.233
mir-730.41632650.272
mir-50040.41632650.29
mir-36840.41632650.29
mir-19420.41632650.325
mir-19020.41632650.325
mir-15320.41632650.325
mir-21930.41632650.343
mir-14830.41632650.343
mir-125610.41632650.407
mir-120510.41632650.407
mir-120410.41632650.407
mir-120310.41632650.407
mir-120210.41632650.407
mir-94410.41632650.407
mir-94310.41632650.407
mir-93510.41632650.407
mir-76010.41632650.407
mir-76510.41632650.407
mir-132310.41632650.407
mir-64110.41632650.407
mir-63810.41632650.407
mir-63610.41632650.407
mir-63010.41632650.407
mir-60210.41632650.407
mir-60110.41632650.407
mir-59610.41632650.407
mir-57210.41632650.407
mir-56810.41632650.407
mir-56410.41632650.407
mir-55910.41632650.407
mir-49810.41632650.407
mir-44810.41632650.407
mir-32610.41632650.407
mir-37510.41632650.407
mir-18410.41632650.407
mir-13710.41632650.407
mir-30240.41632650.408
mir-94130.4180.414
mir-64220.4180.418

Share and Cite

MDPI and ACS Style

Arif, K.M.T.; Elliott, E.K.; Haupt, L.M.; Griffiths, L.R. Regulatory Mechanisms of Epigenetic miRNA Relationships in Human Cancer and Potential as Therapeutic Targets. Cancers 2020, 12, 2922. https://doi.org/10.3390/cancers12102922

AMA Style

Arif KMT, Elliott EK, Haupt LM, Griffiths LR. Regulatory Mechanisms of Epigenetic miRNA Relationships in Human Cancer and Potential as Therapeutic Targets. Cancers. 2020; 12(10):2922. https://doi.org/10.3390/cancers12102922

Chicago/Turabian Style

Arif, K. M. Taufiqul, Esther K. Elliott, Larisa M. Haupt, and Lyn R. Griffiths. 2020. "Regulatory Mechanisms of Epigenetic miRNA Relationships in Human Cancer and Potential as Therapeutic Targets" Cancers 12, no. 10: 2922. https://doi.org/10.3390/cancers12102922

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