Skip to main content

REVIEW article

Front. Oncol., 18 November 2020
Sec. Molecular and Cellular Oncology
This article is part of the Research Topic Repurposed Drugs Targeting Cancer Signaling Pathways: Clinical Insights to Improve Oncologic Therapies View all 11 articles

Epidrug Repurposing: Discovering New Faces of Old Acquaintances in Cancer Therapy

  • 1Unidad de Investigación Biomédica en Cáncer, Instituto Nacional de Cancerología-Instituto de Investigaciones Biomédicas, UNAM, Mexico City, Mexico
  • 2Departamento de Bioquímica, Facultad de Química, Universidad Nacional Autónoma de México, Mexico City, Mexico
  • 3Instituto Nacional de Medicina Genómica, Mexico City, Mexico

Gene mutations are strongly associated with tumor progression and are well known in cancer development. However, recently discovered epigenetic alterations have shown the potential to greatly influence tumoral response to therapy regimens. Such epigenetic alterations have proven to be dynamic, and thus could be restored. Due to their reversible nature, the promising opportunity to improve chemotherapy response using epigenetic therapy has arisen. Beyond helping to understand the biology of the disease, the use of modern clinical epigenetics is being incorporated into the management of the cancer patient. Potential epidrug candidates can be found through a process known as drug repositioning or repurposing, a promising strategy for the discovery of novel potential targets in already approved drugs. At present, novel epidrug candidates have been identified in preclinical studies and some others are currently being tested in clinical trials, ready to be repositioned. This epidrug repurposing could circumvent the classic paradigm where the main focus is the development of agents with one indication only, while giving patients lower cost therapies and a novel precision medical approach to optimize treatment efficacy and reduce toxicity. This review focuses on the main approved epidrugs, and their druggable targets, that are currently being used in cancer therapy. Also, we highlight the importance of epidrug repurposing by the rediscovery of known chemical entities that may enhance epigenetic therapy in cancer, contributing to the development of precision medicine in oncology.

Introduction

Since the turn of the century, epigenetics has become an important research area in human diseases study, where genetic mutations have been classically understood as the main cause in the development of human pathologies (1). The term epigenetics involves a wide variety of mechanisms that cells use to regulate the transcription of their DNA without changing its genetic material (2). Whether an epigenetic modification has a facilitating or inhibiting role in the gene expression depends on the chemical nature of the mark that is placed over the chromatin, and the type of modification that is set down on the proximal environment of these genes (3). Thus, epigenetics shapes a regulatory complex that bridges the gap between genetic sequences and actionable mutations. Due to current knowledge about these epigenetic mechanisms, the importance of this regulatory system has become more evident and it has led to the understanding that epigenetic alterations are some of the main mechanisms underlying many human diseases such as cancer, which arises through aberrant genetic and epigenetic alterations, both of which have a key role in malignant transformation, tumor progression and prognosis (4).

Nowadays, it is known that as cancer progresses, there are genetic aberrations that make tumors highly prone to developing resistance to therapies (5). Emerging data on cancer-associated epigenetic alterations have shown that epigenetic modifications leading to drug resistance may be the cue for individual variation in chemotherapy response, having the potential to be reversible using epigenetic therapy (6). The possibility to reprogram the cancer epigenome is becoming a promising target therapy for both, treatment development and reversibility of drug resistance. Which focuses on the development of pharmacological compounds that can reprogram the epigenetic landscape to enhance chemotherapy response (7).

For a few years, the design of therapeutic strategies has been a growing field of query for single-target epigenetic drugs (epidrugs); however, the traditional epidrug discovery pathway is time-consuming and expensive (8, 9). Hence, a promising strategy for epidrug development is based on tracing novel potential epi-targets in previously approved drugs through a process called drug repositioning or repurposing (10, 11). Epidrug repurposing allows exploring a wide diversity of molecular combinations in multifactorial diseases such as cancer, where combinational epigenetic therapies are likely to be more effective than monotherapy to overcome chemotherapy resistance (9). This review focuses on the emerging area of ​​epidrug repurposing, highlighting strategies to enhance cancer therapy. To further understand this, we will discuss the main mechanisms and elements involved in epigenetic alterations in cancer and its relevance in cancer therapy response.

Background in Epigenetics

Epigenetics is the term coined by Conrad Hal Waddington seventy-six years ago, to refer to the molecular mechanisms that may exert their influence on gene expression that do not involve alterations in its gene code. Through these, an organism can develop and adapt its phenotype to environmental changes (12). Over time, many definitions of Epigenetics have arisen (13); however, we can understand epigenetics as reversible chemical modifications of DNA and histone proteins (epimarks) that regulate specific functions in chromatin remodeling without altering the DNA sequence (14). Epimarks are associated with the transcription and function of a gene, that may change the cellular phenotype or its functional patterns in response to a particular context, across different developmental stages, cellular differentiation, or maintenance of tissue-specific cell lineages (15).

At the molecular level, epigenetic machinery is composed mainly of three interconnected components working synergistically in the chromatin organization levels, which include DNA methylation, histone post-translational modifications, and regulatory non-coding RNAs (ncRNAs) (14, 16). In the nucleus, chromatin can exist in two physical and functional states: heterochromatin (condensed chromatin), which is associated with transcriptional repression; and euchromatin (relaxed chromatin), associated with transcriptional activation (17) (Figure 1). The organizational states of the chromatin are highly regulated by epigenetic mechanisms involving nucleosome, which is the basic packaging unit of chromatin, composed by an octamer of histone proteins (two dimers of H2A-H2B and a tetramer of H3-H4 histones) (Figure 1A), that constitutes a compact structure with 147 base pairs of DNA turned almost twice around it (17, 18). N-terminal tails of histone proteins can acquire post-translational modifications through multiple mechanisms including phosphorylation, ubiquitination, methylation/demethylation, and acetylation, the latter being the most studied. Histone and direct DNA modifications constitute “the epigenetic code”: an interplay between epigenetic factors and positive and negative feedback mechanisms that regulate it (18). Therefore, understanding the main mechanisms in the field of epigenetic research and their role in disease development is essential in its application in cancer therapy.

FIGURE 1
www.frontiersin.org

Figure 1 Overview of the epigenetic landscape. Different compaction levels of chromatin are depicted, from naked DNA to the metaphasic chromosome. (A) Two dimers of H2A-H2B and a tetramer of H3-H4 histones are required for nucleosome assembly, the chromatin’s basic packaging unit (B) DNA methylation is a process carried out by DMNTs in CpG dinucleotides, particularly on CpG islands. This dynamic epigenetic mark can be reversed by enzymatic conversion. (C) Histone acetylation is performed on lysine residues by HAT enzyme complexes. In contrast, histone lysine deacetylation is carried out by HDACs enzyme complexes. (D) Histone lysine methylation is carried out by HMT complexes. Lysines can be processively methylated from mono to di and trimethylation.

DNA Methylation

Methylation on DNA’s cytosine is the most broadly studied epigenetic modification in humans. It encompasses a reaction defined as “the covalent transfer of a methyl group to the C-5 position of a cytosine ring of DNA” (15, 19). Generally, in mammals, DNA methylation occurs predominantly—but not exclusively—in the context of genomic regions called CpG islands, which are formed by clusters of CpG dinucleotides, and it’s catalyzed by a group of enzymes called DNA methyltransferases (DNMTs). These enzymes transfer a methyl group from the donor molecule S-adenosylmethionine (SAM) to the fifth carbon of a cytosine residue to form 5-methylcytosine (5mC) (18, 19) (Figure 1B). This covalent modification is able to inhibit DNA transcription; either through the steric hindrance imposed by the methyl group which prevents transcription factors from binding DNA (1820), or by the recruitment of proteins with methyl-CpG-binding domains (MBD). These proteins also contain domains able to recruit histone-modifying and chromatin-remodeling complexes to the methylated sites, forming repressor complexes that enhance the silencing state on that chromatin region (21). Three different DNMTs generate and maintain methylation patterns. DNMT1 is the methyltransferase enzyme specialized in the maintenance of previously placed methylation patterns, and DNMT3a & DNMT3b are instead involved in the establishment of de novo methylation patterns over DNA (18, 22, 23).

DNA methylation patterns occur in different regions of the genome. Alterations in these patterns lead to diseases (18). For instance, gene promoters which are mainly embedded in CpG islands (70%) are normally unmethylated, thus allowing transcription. Aberrant hypermethylation patterns of these gene regulatory elements lead to transcriptional inactivation and are tumor-type specific as well as a common hallmark of cancer (9). Alternatively, during diseases, other alterations occur, like the demethylation of the gene body. Such alteration allows transcription to be initiated at several incorrect sites. In consequence, DNA hypomethylation at specific regions can activate the aberrant expression of genes, some of which could behave as proto-oncogenes (18). Finally, as aforementioned, alterations of hypermethylated patterns in repetitive sequences can promote the activation of transposable elements and chromosomal instability, both phenomena being also correlated with carcinogenesis and metastasis (6).

However, the reactions that lead to altered patterns of DNA methylation can potentially be reversible and restored through DNMT inhibitors (DNMTi: see below) that contain nucleoside derivatives and non-nucleoside analogs, some of them have been highly researched and shown promise in cancer therapies (24).

Histone Post-Translational Modifications

Another axis of the epigenetic machinery, closely associated with DNA methylation, are the covalent post-translational modifications of nucleosomal histones. Through the addition of chemical groups at specific sites within the amino- or carboxy-terminus of each histone, different functional consequences influencing chromosome structure can be elicited. Chromatin is functionally divided into actively transcribed euchromatin and transcriptionally inactive heterochromatin, which finally regulates the accessibility to genomic DNA and has a role in the control of gene expression (18, 25). The principal histone proteins modifications include methylation, acetylation, phosphorylation, ubiquitylation, sumoylation, and ribosylation, from which methylation and acetylation are the most common and characterized, and generally occur in the proximity of promoter and enhancer genomic regions (26). Each histone residue can undergo one or more modifications, which have different effects depending on which residue is modified, giving rise to crosstalk between the different marks, constituting “the histone code” altogether (18).

Multiple enzymes catalyze histone post-translational modifications with specific catalytic activity based on each histone tail’s amino acids that can act as their substrates. Most of these modifications are reversible. There are specialized enzymes that can remove each type of covalent modification. Histone acetyltransferases (HATs) and deacetylases (HDACs) control acetylation, as well as histone methyltransferases (HMTs) and demethylases (HDMs) coordinate histone methylation. Acetylation and deacetylation of histones are among the most studied reversible, followed by methylation and demethylation of histone lysines (17, 27).

Due to the importance of histone epimarks in gene regulation and cellular function, aberrant histone post-translational modifications may change gene expression patterns and cause human pathologies (6). Thus, it is of great importance to understand the reversible nature of these marks as an advantageous alternative for the treatment of diseases where epigenome deregulation is one of the hallmarks. 

Histone Acetylation and Deacetylation

Histone acetylation has a key role in many biological processes (cell cycle regulation, alternative splicing, nuclear transport, among others) (28). It can promote relaxed states of the chromatin (euchromatin) and favor gene transcription, while deacetylation exerts the opposite effect, generating heterochromatin domains that can inhibit transcription (2). Two families of enzymes with reverse functions control the feedback regulation between acetylation/deacetylation of histones: histones acetyltransferases (HATs or KATs) and histones deacetylases (HDACs) (2). HATs catalyze the transfer of acetyl groups to lysine-amino-terminal residues using acetyl-CoA as a donor; this reaction neutralizes the positive charge of the Lys (17, 29) (Figure 1C). As a result, the interaction between the histone and the DNA is weakened, forming an opening domain in chromatin, leading to exposure of DNA sequences and their transcription (2, 28). HATs are divided in three families based on their catalytic domain’s functional and structural identity, which bears the acetyltransferase activity for the recognition of acetyl-lysine residues (17). Several HATs associate with other protein complexes and subunits to selectively modify the different histones; however, p300/CBP is probably the most extensively studied HAT, since it is capable of acetylating all four histones along with many other coactivator or corepressor transcriptional complexes (30).

In contrast, HDACs remove acetyl groups from lysine residues through different reactions that reestablish the positive charges on histone tails, increasing its interaction with DNA and stabilizing the chromatin in place (2, 28) (Figure 1C). The histone deacetylase family includes 18 members (31), divided into two groups based on their enzymatic activity: Zn2+-dependent enzymes, which include classes I, II, and IV HDACs, exert their function through hydrolytic catalysis; and NAD+ cofactor-dependent enzymes, that include class III sirtuins (SIRTs), with a catalytic mechanism of nucleophilic substitution for histone deacetylation (28).

Both HATs and HDACs play a key role in the maintenance and regulation of chromatin accessibility, leading gene expression regulation, among other mechanisms. Histone acetylation global imbalance is one of the prominent alterations in the diseased state and a hallmark of many tumor types, where HDACs have been found overexpressed (32) or mutated (33). Additionally, abnormal genomic events such as translocations, mutations, or deletions in HAT- and acetylation readers-related genes may occur during cancer development (18). As a result, aberrant acetylation-related proteins contribute to the progression of the disease. For instance, germline mutations and overexpression of HDACs have been observed in various cancers, resulting in a global loss of histone acetylation and the consequent silencing of tumor suppressor genes (34). Also, it has been observed that reduced lysine 16 acetylation (H4K16ac), as well as the loss of acetylation of histone 3 (H3ac) are also hallmarks of human cancer (35, 36). Furthermore, HATs and HDACs are targeted to transcriptionally-active genes by phosphorylated RNA polymerase II through the recruitment of effector proteins with specialized reader domains (18), suggesting that the mechanistic switch between acetylation/deacetylation can be manipulated and restored by specific drugs inhibiting key enzymes by targeting their catalytic reaction (HATi and HDACi; see below).

Histone Methylation and Demethylation

Histone methylation occurs on arginine (R) and lysine (K) residues, and it is catalyzed by HMTs (or KMTs and RMTs) that use S-adenosyl-l-methionine (SAM) as a methyl donor group (Figure 1C). Lysine methyltransferases are divided into two broad groups based on the presence or the absence of a SET domain (Su(var)3-9, Enhancer-of-zeste, and Trithorax): SET-domain containing methyltransferase family and DOT1-domain lysine N-methyltransferase (37, 38).

KMTs can transfer three methyl groups onto lysine residues, prompting mono, di, and, trimethylation (me1, me2 and, me3 respectively) (17) (Figure 1D). The association of an active or repressive transcriptional state depends on the number of methyl groups and in the position of the lysine residue in the histone amino acid sequence. A repressed chromatin state (heterochromatin, constitutive, or facultative), correlates with methylation of H3K9me2,3, H3K27me2,3, and H4K20me3, while methylation of H3K4me2,3, H3K9me1, H3K27me1, H3K20me1, and H3K36me1 are associated with transcriptionally active chromatin (euchromatin) (17, 39). Besides, histone methylation also has an important role in DNA repair, DNA replication, alternative splicing, and chromosome condensation (18). Histone demethylases HDMs (or KDMs) can revert these modifications (Figure 1D), divided into two different families with distinct enzymatic mechanisms: KDM1A/LSD1 amine oxidase family, dependent on flavin adenine dinucleotide (FAD) as a cofactor; and the KDM2A/B dioxygenase family, which contain a Jumonji C (JmjC) domain and are iron Fe (II) and α-ketoglutarate-dependent to accomplish histone demethylation through methyl groups oxidation (40). The readers of methylated lysine residues consist of various proteins with specialized domains that can recognize these modifications (17).

Besides the global loss of acetylation and DNA hypomethylation, the deregulation of histone methylation/demethylation can lead to chromosome instability (18). It has been suggested that the aberrant expression of both histone methyltransferases and demethylases genes is the main cause of an altered distribution of histone methylation marks. Deregulation of histone methylation patterns can become a driver for mutations in many types of tumors (15). For instance, cancer cells have a global loss of activation marks, such as H4K20me3; along with a gain of methylation in repressive marks, such as H3K9me and H3K27me, as well as the monomethylation of H3K4me (35, 36) which are associated with DNA hypermethylation of silenced genes. The basal patterns of histone methylation are essential for establishing a permissive euchromatic state, allowing the expression of tumor suppressor genes. Therefore, its alteration results in the repression of some of these genes and oncogene aberrant expression (18, 35). Instead, instability of the methylation/demethylation mechanistic switch can promote proliferation and neoplastic transformation in several cancer types (4143).

Epigenetic Alterations in Cancer and Cancer Therapy

As mentioned before, the cancer epigenome is characterized by global changes in DNA methylation, disruptions in histone posttranslational modification patterns, and alterations of normal chromatin-modifying enzymes expression (18, 36) (Figure 2) [see review (44)]. Accordingly, these changes can promote the disruption of cellular homeostasis in precancerous cells through the deregulation of genes implicated in cancer initiation and progression (4); for instance, those genes associated with apoptosis resistance, proliferation, invasive potential, and genomic instability, as well as genes correlated to therapeutic response (45, 46). Thus, the relationship between genetic disruptions and epigenetic abnormalities are mutually beneficial in order to drive cancer development and could be playing a key role in individual differences displayed by patients in the way they respond to therapies in both toxicity or treatment efficacy (15, 46, 47). Multiple studies demonstrate that reversing epigenetic patterns through de novo epidrugs and epidrug repurposing can resensitize cancer cells to chemotherapy (4850).

FIGURE 2
www.frontiersin.org

Figure 2 Epigenetic alterations in cancer cells. In non-neoplasic cells, CpG islands of tumor suppressor gene promoters are generally unmethylated and acetylated, resulting in transcriptional activation and expression. In contrast, non-coding regions and repetitive elements are hypermethylated, ensuring chromosome stability. Gene bodies are normally methylated, enhancing transcription. Neoplasic cells are characterized by global hypomethylation and local CpG island hypermethylation, especially at tumor suppressor gene promoters, resulting in aberrant transcription and genomic instability.

Principles of Epigenetic Therapy

Increasing understanding of epigenetic mechanisms and their importance in disease has led to the development of therapeutic interventions targeting epigenetic modulatory mechanisms. Due to the chemical reversibility nature of DNA methylation and histone posttranslational modifications, epigenetic proteins can be druggable targets by means of small-enzymatic inhibitors that aim for the restoration of the aberrant epigenetic machinery and hold the potential for reverting epigenetic signatures in cancer (14).

Epigenetic drugs (epidrugs) are chemical agents that modify the structure of DNA and chromatin, facilitating disruption of transcriptional and post-transcription changes, primarily by controlling the enzymes required for their establishment and maintenance, reactivating the tumor suppressor and DNA repair genes that are epigenetically silenced (51). Lately, epigenetic therapy has taken relevance in the field of oncology, where epidrugs have been successfully used in treatment, mostly in combination with standard chemotherapy (52).

Epidrugs (with one-target, as well as repurposed epidrugs; see below) that are designed based on these principles can exert direct cytotoxic effects over malignant cells (14, 46), function as sensitizers in complementary therapies (53, 54), or can be used to overcome epigenetically-acquired drug resistance against the limits of chemotherapy efficacy, as there are the dynamic associations between epigenetic pattern changes and resistance to therapeutic regimes for cancer (50, 52, 55). New epidrugs compounds are continually being tested for cytotoxicity, pharmacological parameters, and a better understanding of their mode of action; in both preclinical research (in vitro and in vivo) as well as in clinical trials. Epigenetics therapy is enhanced by a combination of laboratory and clinical data. The US Food and Drug Administration (FDA) has approved many epigenetic treatments and used them for treating cancer (6).

Epidrug Generations

Historically, molecules designed to inhibit the catalytic function of epigenetic factors have not only resulted in the reduction of the targeted enzymatic activity but also the appearance of indirect modifications of the transcription of large gene sets (56). Several epigenetic protein families have similar cofactors and co-substrates, similar epidrugs could target several epigenetic protein families. Some compounds can inhibit the functionality of a whole family of epigenetic proteins (Table 1).

TABLE 1
www.frontiersin.org

Table 1 Current inhibition assays performed for different epigenetic factors.

The quest for finding epigenetic inhibitors led to the first generation of epidrugs, characterized by a meager degree of selectivity (57). Epidrugs of the first generation include DNMTi and HDACi, some of which have already been approved to treat hematological malignancies (58). DNMTi are pyrimidine analogs incorporated into DNA during replication and form covalent DNA adducts that cause DNA damage response activation and eventually lead to apoptosis. This was not without cytotoxic implications (3, 59). On the other hand, first generation HDACi are molecules that inhibit the Zn2+ dependent HDAC enzymes, except for sirtuin inhibitors, which inhibit a specific class of histone deacetylases that depend on NAD+ to perform their catalytic activity (59).

First-generation inhibitors represented many undesirable pharmacokinetic properties and poor target selectivity, resulting in the need for the creation of second-generation epidrugs, which included DNMTi (such as zebularine and guadecitabine), and HDACi (including hydroxamic acid, belinostat and panobinostat, tucidinostat and valproic acid) with improved physiological properties (59).

The second generation of epidrugs was characterized by strong academic research accompanied by industrial drug discovery to find molecules that resembled first generation epidrugs. The hypothesis was that molecules with more potent inhibitor action and fewer side-effects could be found. Another thing to consider was pharmacokinetics: first generation epidrugs had poor bioavailability, were more active within non pH physiological ranges, and were targets of cellular deaminases, which ultimately meant a short half-life for these compounds.

Ultimately, the third generation of epidrugs reflected that epigenetic factors could write, delete, or read epigenetic marks in the form of protein complexes. Therefore, a deeper understanding of epigenetic protein’s interactome is essential for the design of highly selective epidrugs (57). Epi-drugs of third generation includes, among others, histone methyltransferase inhibitors (HMTi), histone demethylase inhibitors (HDMi), and bromodomain and extra-terminal domain inhibitors (BETi) (59).

DNMT Inhibitors

DNA methylation inhibitors intercalate between DNA base pairs and suppress the CpG dinucleotide’s methylation, especially important at CpG islands. These inhibitors can be classified as DNMTi nucleoside analogs and non-nucleoside analogs (60) (Figure 3). DNMTi cytidine analogs are usually chemically unstable, and because of their similarity to cytidine, DNA and RNA polymerases identify both compounds and add them into growing nucleic acid chains, therefore hampering their selectivity (61).

FIGURE 3
www.frontiersin.org

Figure 3 Classification of epigenetic inhibitors. Epigenetic inhibitors are classified as DNMTi, HDACi, HMTi, HDMI, and BETi. The chemical nature of each inhibitor defines the affinity of its targets.

Since the first DNMTi discovery (azacytidine), the number of inhibitors of DNMT has increased exponentially. The CHEMBL database reports 841 compounds tested for DNMT1 inhibition (CHEMBL1993), 258 compounds for DNMT3A (CHEMBL1992), and 80 compounds for DNMT3B (CHEMBL6095) (62) (Table 1, DNMTi section).

Among azacytidine derivatives, 5-aza-2’-deoxycytidine gained importance in the clinic, commonly known as “Decitabine”. Decitabine contains DNA sugar deoxyribose and is only integrated into DNA, while azacytidine allows for both RNA and DNA incorporation (14). Of note, Azacitidine and decitabine have both the same action mechanism. They both behave as a suicide substrate, trapping DNMTs after metabolic conversion and incorporation into DNA (3).

Guadecitabine is a hypomethylating agent of the second generation whose active metabolite is decitabine. Guadecitabine holds an amazing property: it is not a cytidine deaminase substrate, thus improving its selectivity. This drug has shown promise in treatments and recently tested in a Phase II clinical trial for treating non-intensive chemotherapy candidates with AML (63).

In 2004, azacytidine became the first medication approved by the FDA for all stages of myelodysplastic syndrome, a bone marrow disorder with a high risk of AML progression, characterized by irregular blood cell development, followed by decitabine in 2006 (64). These two drugs are currently used as first-line MDS therapy when other therapies are insufficient (14) (Table 2, DNMTi section).

TABLE 2
www.frontiersin.org

Table 2 Overview of epigenetic inhibitors currently in clinical trials for cancer therapies.

As mentioned before, DNMTs have two substrates, the methyl group donor cofactor SAM and the methylated cytosine. Non-nucleoside DNMTi includes analogs of the methyl donor S-adenosyl-L-methionine (SAM) and small molecules that interact with the active site of the enzyme DNMT (Figure 3). Indeed, it is possible to obtain potent DNMT inhibitors by designing substrate analogs and connecting them (65). This strategy has resulted in the most effective way to inhibit DNMTs and reactivate genes in cancer cells by promoting demethylation (60). Many forms of these derivatives have shown remarkable results in many models of cancer and other human diseases. These include hydralazine, EGCG, RG108, MG98, and disulfiram (6671) (Table 2, DNMTi section). MG98 is a second-generation phosphorothioate antisense oligodeoxynucleotide that inhibits translation effects of DNMT1 mRNA but has no apparent impact on tumors (72).

Despite preclinical evidence indicating a potentiating chemotherapy cytotoxic activity of HDAC inhibitors and DNMT inhibitors, clinical outcomes have been discouraging: three of the five main combination randomized trials were stopped because of ineffectiveness or disadvantaged toxicity profiles compared to chemotherapy alone (59). The possible role of DNMT inhibitors remains unclear, but in conjunction with other therapies, these agents may theoretically still be of use.

There is a good scientific justification for combining DNMT inhibitors with HDAC inhibitors since both hypermethylated DNA and hypoacetylated histones are associated with closed chromatin states that repress gene expression by independent mechanisms. Further studies should be carried out into the efficacy of this combination at different dosages and durations of treatment. To date, hundreds of clinical trials have studied the effects of anti-DNA methylation therapy on different cancers.

HDAC Inhibitors

The development of the first HDACi commenced with the finding that erythroleukemia murine cells differentiated in the presence of dimethyl sulfoxide (DMSO). Later, chemical analogs that could make similar interactions as DMSO were studied (56). This was the case of vorinostat (SAHA), a molecule capable of metal coordination and hydrogen bonding. Interestingly, natural compounds inhibitors of HDACs (trichostatin A and trapoxin A) were found to chemically resemble vorinostat at the hydroxamic acid moiety. The mechanism of action of these compounds inhibits HDACs by reversibly binding to Zn2+ in the enzyme’s active site. Since the discovery of vorinostat, a lot of new activity assays are performed every day with inhibitor compounds (62) (Table 1, HDACi section).

Zinc binding is essential for the inactivation of most HDACs (56). As mentioned before, the Zn-binding hydroxamic moiety has proven to be one of the most successful inhibitors, and thousands of synthetic HDAC inhibitors with this moiety have been reported. Many of these inhibitors have focused primarily on optimizing the pharmacokinetics of vorinostat and trichostatin A (Figure 3; Table 2, HDACi section).

Currently, vorinostat therapy clinical applications have been applied to neurological conditions and, surprisingly, to reactivating chronic viral infection (73). Therapies for HIV-1 patients do not kill the virus entirely because it may be latent in reservoirs of CD4 + cells (74). Epigenetic mechanisms regulate viral latency, and so, clinical trials to test the effect of vorinostat therapy in reactivation of HIV-1 viral latency are currently being performed.

This optimizing focus led to the design of the hydroxamic acid containing HDACi, such as belinostat, dacinostat, givinostat, and panobinostat. The latter being the only HDACi with approval within the EU. As single agents, these molecules have shown limited efficacy, but when in combination with DNMTi, they have shown to be more effective, especially in patients with solid tumors (75, 76). Other metal-binding functional groups have been of great interest to this group. This is the case of thiols, benzamides, and carboxylic acids (56). Examples of these functional groups can be found in the drugs: romidepsin, entinostat, mocetinostat, and short-chain fatty acids, such as sodium butyrate, Pivanex, phenylbutyric acid, and valproic acid (Figure 3; Table 2, HDACi section).

Unlike hydroxamic acid analogs, short-chain fatty acids occupy an acetate escape tunnel, which may have a zinc-binding function or compete with an acetate group released in the deacetylation reaction. These are the least potent type of HDACi (77). The benzamide inhibitor class consists of a chemical moiety capable of contacting specific amino acids in the HDAC core tube active site, with or without zinc ion binding (78). These inhibitors are active at micromolar levels. The antiproliferative and cytotoxic activity has been shown by entinostat against several tumor cell lines in vitro. Entinostat is a clinical trial available orally active inhibitor (79) (Figure 3; Table 2, HDACi section).

Currently, the discovery of sirtuin inhibitors (SIRTi) is an ongoing quest in which most compounds are still under preclinical investigation (80). Most efforts have been driven toward the discovery of SIRT1 and SIRT2 inhibitors. SIRT1 inhibitors have been proposed for treating cancer, for they have shown to inhibit TNBC cell growth, survival, and tumorigenesis (56, 81). Nicotinamide is the only inhibitor of sirtuin currently used in solid tumor clinics (82). SIRTi can be categorized as β-naphthols (sirtinol, splitomicin, salermide, and cambinol), indoles (EX-527 and oxindole), and urea (suramin and tenovin) (83) (Figure 3; Table 2, SIRTi section).

HDACi have many biological effects due to changes in patterns of histone acetylation and many non-histone proteins, including proteins involved in gene expression control, extrinsic and intrinsic apoptosis pathways, the progression of the cell cycle, redox pathways, mitotic division, DNA repair, cell migration and angiogenesis (56). Whether selective inhibition of HDACs will be beneficial as anti-cancer agents over broader-acting HDACi is a question that remains unanswered (56).

Histone Methyltransferase Inhibitors

HMTs are enzymes that add up to three methyl groups to lysine (KMTs) or arginine (RMTs) residues in histone proteins (84). Lysine methylation may either activate or silence gene transcription depending on the lysine residue involved (85). Nearly 100 KMTs have been described which use the SAM molecule as the methyl donor (14). SAM-like molecules, such as sinefungin, compete with SAM for its binding site (Figure 3). These molecules are inhibitors of all SAM using enzymes, like HMTs (14). KMT drug discovery heavily relies on their cofactor binding pocket, which has structural characteristics convenient for inhibitor interaction and makes these enzymes appealing for the design of small molecular inhibitors for interference (80). Examples of HMTi can be found in drugs such as EPZ004777, EPZ-5676, DZNep, pinometostat, and tazemetostat. Pinometostat and tazemetostat are selective DOT1L and EZH2 inhibitors, respectively (Table 2, HMTi section).

Both inhibitors are of interest in some types of cancer because DOT1L is a KMT involved in abnormal methylation of H3K79 and expression of HOX genes that cause leukemia (Copeland et al., 2013), while elevated expression of the KMT, EZH2, is associated with many forms of cancer due to hypermethylation of H3K27 which facilitates transcriptional silencing (80). Also, in B-cell-lymphoma patients, EZH2 mutations occur with a frequency of approximately 15-20 percent in either tumor type, particularly in diffuse large-B cell-lymphomas and follicular lymphomas (86, 87). These modifications contribute to the more effective trimethylation of H3K27 by the mutant form of this protein (88). Preclinical studies showed that EZH2 inhibitors induced the arrest of proliferation, differentiation, and eventual apoptosis of DLBCL cells. These results were stronger in DLBCL cells that bear EZH2 mutations, but they also occurred in EZH2-wild-type DLBCL cells (89).

While several small molecule inhibitors have been developed for PRMTs with adequate potency, most PRMT inhibitors’ selectivity remains to be improved. Therefore, the detection of PRMT inhibitors involves further analysis of novel approaches (i.e., allosteric control) (90). Three PRMT inhibitors, including PRMT5 inhibitor GSK3326595 (Table 2, HRMTi section), and JNJ-64619178 as well as PRMT1 inhibitor GSK3368715 have entered clinical trials so far. PRMT inhibitors with novel action mechanisms and strong drug-like properties will shed new light on developments in drug discovery and development of PRMTi (87, 90). The number of inhibitor assays reported on CHEMBL database against the enzymatic activity of the HMTs increases everyday (62) (Table 1, HMTi section).

Histone Demethylase Inhibitors

Significant progress has been made in the development of JmjC-KDM inhibitors since the first inhibitors were identified in 2008 (91). The vast majority enter the catalytic domain and inhibit the enzyme’s activity by chelating the active site Fe (II), interfering with the 2OG binding. Because of the similarity between JmjC-KDMs’ active site pockets, it has proved difficult to achieve selectiveness in the broad superfamily of 2OG dioxygenases (92). The recent availability of JmjC-KDM crystal structures has encouraged medicinal chemistry efforts and has made it possible for the JmjC-KDMs to produce many chemical candidates. Examples of these inhibitors include hydroxamate derivatives, pyridinedicarboxylate derivatives, N-oxalyl amino acid derivatives, and agents which interfere with metal binding (71) (Figure 3; Table 2, HDMi section).

In 2004, Professor Yang Shi first described LSD1 and discovered that it had significant biological functions in a wide variety of biological processes, including cancer (93). During carcinogenesis, in AML and SCLC, elevated levels of LSD1 were observed (94). Pharmacological LSD1 inhibition with small molecules has shown that it suppresses the division, proliferation, invasion, and migration of cancer cells (95). LSD1 thus becomes an evolving clinical target for anticancer therapy. Many LSD1 inhibitors, including natural products, peptides, and synthetic compounds, have been identified.

The similarity of LSD demethylases with monoamine oxidases (MAOs) has started the quest for repurposing MAO inhibitors to find inhibitors for these types of enzymes. Initially approved by the FDA for the treatment of mood and anxiety disorders (96), the MAO inhibitor tranylcypromine (TCP) was found to be able to inhibit its homolog LSD1 moderately by forming covalent adducts (97). As a result, many MAO inhibitors (MAOi) such as pargyline, phenelzine, and tranylcypromine have been shown to inhibit HDM KDM1A (80) (Figure 3; Table 2, HDMi section). New studies are now ongoing in clinical trials with some TCP-based LSD1 inhibitors alone or combined therapy with other therapeutic agents for treating cancer (98).

Bromo and Extra Terminal Domain Inhibitors

Bromodomains are protein motifs present in several epigenetic readers including BET family, that recognize and bind to acetylated lysine residues located on histone tails. BETs consist of two bromodomains and an extra-terminal region. The BET family includes the Bromodomain testis-specific protein (BRDT), BRD2, BRD3, and BRD4 (99). BETs lead to malignancies production and progression by stimulating and enhancing the expression of main oncogenes such as MYC (100). Indeed, when treated with the inhibitor JQ1, BET inhibition resulted in MYC downregulation, which resulted in decreased levels of mRNA and protein in mouse MLL-fusion leukemia cells (101).

In various forms of cancers, including breast, neuroendocrine, ovarian, rhabdomyosarcoma, and glioma, preclinical studies of BET inhibitors have shown their efficacy (87). They disrupt the recognition by BET-containing reader proteins of acetylated lysine residues in histones, a mark associated with active transcription (102). The mechanism of BETi relies on the fact that the region that binds acetyl-lysine is hydrophobic and can be taken up by small hydrophobic molecules that specifically target this catalytic site. Examples of these inhibitors can be found in Thienotriazolodiazepines (JQ1, CPI-203, OTX015) and Benzodiazepines (CPI-0610 and molibresib) (Figure 3; Table 2, BETi section).

Preliminary clinical trials have demonstrated that BET inhibitors cannot induce long-lasting cytotoxic effects in human cancers when administered as single agents (103). Nevertheless, the potential of combinations with other epigenetic therapies is important (104). Although BET inhibitors’ toxicity may reduce such combinations, HDACi studies indicate that combinations with reduced doses may be effective, possibly reducing toxicity. This also reflects on the number of inhibitor assays for BRDs (62) (Table 1, BETi section).

The Basis for Drug Repurposing

Although epigenetic therapy has proven to be remarkably effective, epidrug discovery remains as a traditional “de novo” drug discovery pathway, which has significant disadvantages such as high costs, time consuming, and low success rate (105, 106) (Figure 4). An answer that addresses these problems and could speed up epidrugs in the clinic has arisen from the relatively recent idea of using known drugs for new targets, commonly known as drug repurposing (DR). This approach has gained considerable popularity, emerging as an interesting approach in cancer therapy research and many fields within medicine (107).

FIGURE 4
www.frontiersin.org

Figure 4 Advantages of pharmacological epi-drug repurposing in clinical applications. Drug repurposing serves as a shortcut reducing the time of incorporating a drug into the clinic; since the preclinical phase has already been carried out previously, giving a second chance to old drugs. Initially, it reduces the cost of development and toxicity research, which leads to greater cost-benefit efficiency for the pharmaceutical industry by generating a new cancer therapy. The repositioning of epi-drugs is a promise for the generation of new drugs of precision medicine.

DR is the discovery process of finding new medical uses of a preexisting drug which was previously approved for another indication, withdrawn from the market due to adverse effects or disapproved for failing to prove its efficacy and safety (11, 107) (Figure 4).

This approach includes the selection of drugs with promising repurposing potential and it also has important advantages over the “de novo” drug discovery processes. Previously assessed drug safety significantly reduces both costs and time for making these drugs readily available for use in the clinic (108, 109).

Historically speaking, repurposing of medications was mainly fortuitous; if an off-target effect or newly discovered target was detected, it was sure for it to be targeted for commercial usage. Examples of this are shown in drugs like sildenafil citrate, whose repurposing for erectile dysfunction was not based on a systemic approach, nor was thalidomide repurposing for erythema nodosum leprosum (ENL) and multiple myeloma, which are still the most promising examples of DR (107). Sildenafil was first formulated as an antihypertensive medication. However, after Pfizer reprofiled it for erectile dysfunction therapy and sold it as Viagra, it held the lead market share in erectile dysfunction medications in 2012, with global sales totaling more than 2 billion (110). Thalidomide, an antiemetic first sold in 1957, was discontinued within four years due to its notorious association with teratogenic defects in infants born to mothers who took the drug during their first trimester of pregnancy (107). However, the efficacy of thalidomide, first in ENL and decades later in multiple myeloma has been successfully demonstrated. Ever since, thalidomide has achieved considerable market success for treating multiple myeloma and has also contributed to the production and authorization of many more effective formulations, such as lenalidomide, which had $8.2 billion in worldwide revenues in 2017 (111).

These achievements have led to the implementation of systematic approaches to detect repurposable substances (109). The field of DR is fascinating, and its importance reflects in the vast number of drug projects of pharmaceutical companies that already have several candidate molecules that, although successful in phase I, they did not prosper in Phase II or III clinical trials. This gives rise to the existence of several known molecules, which are relatively safe to use in the clinic. Hence, this large reservoir of molecules provides a vast niche for the search for repositionable drugs, which is much larger than the set of approved drugs (112).

A DR approach usually consists of three phases before the target drug is taken into further development: The selection of a target molecule for a specific indication, analysis of the drug impact in preclinical models, and the evaluation of the effectiveness in clinical trials in phase II, when enough adequate safety results are available from phase I tests. These methods can be classified into computational approaches and experimental approaches, which are now both being widely used synergistically. DR is encompassed within these two large fields, focused on clinical evidence (109).

Experimental approaches include binding assays for the identification of novel target interactions. These types of assays come from proteomic methods, like affinity chromatography and mass spectrometry are used to detect novel targets of existing drugs (113); and phenotypic screening, which are approaches based on in vitro or in vivo models of disease screening of compounds can indicate clinical potential (114). These approaches offer testing in a relevant biochemical context by performing in vitro assays with live cells (115, 116). The evolution of in vitro screening has led to systematize drug discovery, allowing ultra-high-throughput screening, analyzing up to 10,000 compounds per day (116, 117); however, the main limitation of these methodologies are the high costs of the required infrastructure, as well as nonspecific results (8).

Computational methods include the study of large sets of data (e.g., gene expression, chemical composition, genotype or proteomic data or electronic health records) that lead to the development of reprofiling hypotheses (118). Computational approaches include: signature matching, which results for comparing a drug signature such as its transcriptomic, structural or adverse effect profile to that of another pharmaceutical product or disease phenotype (119); molecular docking, a structural computational strategy focused to predict complementarity of the binding site between a drug and a receptor (120); genetic association, a high throughput analysis of genes associated with a disease which can turn out to be potential targets for drugs (121); pathway mapping, another approach that analyses biological pathways in order to develop networks of drugs or disorders based on patterns in gene expression, disease biology, protein interactions or GWAS data to better classify repurposable candidates (122); retrospective clinical analysis, a systematic review of electronic health records, data from clinical trials and surveillances post-marketing could be useful identifying repurposable drugs; and novel sources, which is the combination of large-scale in-vitro drug screens with genomic data, electronic health records and self-reported patient data represents new ways to repurpose drugs (123, 124).

In sum, these approaches allow multiple manners for conducting DR. However, these methodologies applications need to be taken with caution, as many of them seem to be reductionist (117, 125). Numerous strategies are now coupling drug networks with computational analysis to characterize different diseases’ metabolic pathways. These efforts aim to identify drugs acting not only on a single target but also on a whole network of proteins (126, 127). In every computational approach, experimental validation is compulsory since the actual methods are not 100% accurate.

HTS (High-Throughput Screening) is the most common approach in DR of epidrugs, and most of them are designed to inhibit catalytic sites of epigenetic writer enzymes (128). Computational methods, such as virtual screening, aim to efficiently discover novel active compounds against epigenetic factors (8). The increasing attention on epigenetic targets as an opportunity for DR provides high expectations. Next, we will summarize the current efforts in epidrug repurposing for cancer therapy.

Available Databases Focused on Exploration and Recompilation of DR Research

Nowadays, there is a large amount of information available focused on the search and annotation of drugs to be repurposed and the drugs that currently have research that supports their proposed new uses. Some public databases such as ChemBL, DrugBank, and DrugCentral are repositories of bioactivity data and drug chemical structures. These databases summarize multiple indications and chemical drug-target interactions. More specifically, the FDA-approved epidrugs are gathered in several databases focused on tested epidrugs and provides information about annotation tools (Table 3, Section Epidrugs). These databases are useful because they facilitate the integration of epidrug datasets obtained from experimental and computational approaches, reducing the manual search of information, and helping to increase collaboration on the field.

TABLE 3
www.frontiersin.org

Table 3 Some databases and tools that summarize the current knowledge on DR.

Other databases that aim to summarize the current efforts and latest frontiers in DR research are the REPOHub, repoDB, and the Project Repethio; these include clinical trials, pre-clinical tools for annotations, and information resources. Unlike the previous ones, these databases focus on gathering and matching the results from both predictive tools and experimental or clinical trials, resulting in faster results on drugs that could be repurposed (Table 3, Section Drug Repurposing). Tanoli et al., 2020 summarize the types of data available through multi-database exploration focused on DR (142). Currently, the ReDO project (Repurposing Drugs in Oncology) is probably the only database focused on assembling DR for cancer targets. And it has played a crucial role in the development of research for new drugs to cancer therapy with the DR approach.

Epidrug Repurposing in Cancer (Epi-DR)

The interest in oncological DR has emerged as a response to the declining productivity of oncological drug development (143) and as a source of low-cost treatments to meet the increased demands for novel treatments, in efforts to overcome chemoresistance and reduce the development time of de novo drugs (144).

Some widely used and well-known drugs for cancer therapy are examples of epi-DR, with an effect on epigenetic targets, and are either currently FDA-approved or under clinical development (145). The first repurposed drugs as an anticancer epidrug in the field were the 5-azacytidine and 5‐aza‐2′‐deoxycytidine (decitabine) (146). At first, these drugs were both approved by the FDA to treat myelodysplastic syndromes due to their antimetabolic effects on in vitro assays in cancer cells (146). However, the toxicity shown by 5-azacytidine led to other chemotherapeutic regimens being preferred (146); later, it was found that azacytidine and decitabine could both inhibit DNA methylation and were incorporated by tumor cells and also in myelodysplastic syndromes (146148).

DNMT Inhibitors

The natural compound Harmine downregulates the expression of DNMT1, which results in reactivation of the p15 tumor suppressor gene in AML. Future studies are expected to assess if Harmine can be considered a potential therapy for AML and if it can be used as a single agent or adjuvant (149). Chlorogenic acid is a polyphenol coffee that has been found to suppress DNMT1. Its inhibitory activity derives from a chemical change resulting in increased S-adenosyl-L-homocysteine (SAH) production. Chlorogenic acid has been shown to inhibit DNMT1, using breast cancer cell lines, which lowers DNA methylation (150).

Laccaic acid A is a direct, competing DNMT1 natural compound inhibitor that reactivates genes silenced by promoter DNA methylation synergistically with 5-azadC in breast cancer cells (151). Procaine is a promising treatment with growth-inhibiting and DNA-hypomethylation effects in cancer cells. Especially in gastric cancer where its antiproliferative and apoptotic effects have been proven (152). Its well-defined, safe use as a local anesthetic, with well-known pharmacology, should promote procaine to pre-clinical trials (153). Procainamide, a derivative of procaine, hinders the enzymatic activity of DNMT1 by directly reducing the enzyme affinity for both DNA and S-adenosyl-L-methionine. It would be important to analyze whether procainamide, a fairly stable non-nucleoside inhibitor of DNMT1, will prevent cancer from arising (154).

A computer-based search for similarities between a database of approved drugs and 5-aza-2’-deoxycytidine has recently been detected as an ideal candidate for DR. Mahanine, a plant derived alkaloid, was shown to induce DNMT1 and DNMT3B proteasomal degradation by inactivating Akt, which in turn restored RASSF1A expression in prostate cancer cells. Mahanine then represents a possible therapeutic agent for advanced prostate cancer when RASSF1A expression is inhibited (155).

Hydralazine, approved as an antihypertensive, is a non-nucleoside DNMTi that interacts with the binding domain of DNMTs, and can decrease DNMT1 and DNMT3A mRNA expression and protein levels in T cell leukemia cell lines (156). In advanced cervical cancer, bladder, and cervical cancer cell lines, respectively (157, 158), hydralazine induces DNA demethylation and decreases DNMT activity. Also, hydralazine, combined with magnesium valproate, is an opportunity to reverse imatinib resistance in patients with several malignancies, including lung (NCT00996060), cervical (NCT00404326), and locally advanced breast (NCT00395655) cancers, as well as different solid tumors which are refractory to current therapies (159161) (NCT00404508). Olsalazine, an FDA approved anti-inflammatory agent, has proven its hypomethylating and very low cytotoxicity effects in cell-based screen tests (162).

Mithramycin A, an antibiotic with potent antitumor activity, binds to sequences of GC-rich or CG-rich DNA and upregulates tumor suppressor genes’ expression by reducing the methylation of their promoters through binding and depleting the DNMT1 protein in lung cancer cells (163). Nanaomycin A, an anthracycline antibiotic, has demonstrated selectivity to DNMT3B in biochemical assays. Dock modeling strategies suggest that nanaomycin A is capable of binding DNMT3B’s catalytic site. Treatment of the human tumor lines of the colon, lung, bone marrow with nanaomycin A demonstrated substantial genomic demethylation. While it is unclear if anthracyclines will be a successful choice for clinical DR due to certain long-term cardiotoxicity concerns, Nanaomycin A is the first non-SAH DNMT3B-selective compound that offers valuable biochemical properties for additional studies (164).

Disulfiram is an alcohol aversive drug that has been approved by the FDA for more than 60 years for treating alcohol abuse. It allows acetaldehyde to accumulate in the blood by inhibiting ALDH (165). Disulfiram’s anticancer activity is mediated by its ability to suppress DNMT1 and through the reactivation of epigenetically silenced genes such as APC and RARB in prostate cancer cell lines (70) (Table 4, Section 1).

TABLE 4
www.frontiersin.org

Table 4 Current DNMTi and DNMT-HDAC dual inhibitors repurposed drugs with applications in cancer therapy [*modified from Moreira-Silva et al. (9)].

Peptides are small proteins made up of fewer than 50 amino acids. Such compounds have several roles in the human body and can modulate epigenetic pathways, raising the exciting possibility of peptide-based therapy. Such peptides may be endogenous, or food derived. Amyloid beta (Aβ), the central component of Alzheimer’s senile plaque (AD), reduces global DNA methylation but increases DNA methylation in the Neprilysin gene promoting region, an Aβ-degrading enzyme (189). Soluble Aβ oligomers decrease intracellular glutathione levels by hampering cysteine uptake, followed by a global decrease in DNA methylation (174). BCM7 and GM7 are food derived peptides produced by hydrolytic casein and gliadin digestion. They decrease cysteine absorption through opioid receptor activation in neuronal and gastrointestinal cells. This reduction is followed by an increase of oxidized glutathione and an increase in DNA methylation (175, 176) (Table 4, Section 1).

Dual DNMT and HDAC Inhibitors

In most cancer types, altered DNMT and HDAC activity is observed (190). Therefore, some repurposed drugs that inhibit both DNMT and HDAC enzymes could improve efficacy over one-target agents (Table 4, Section 2).

Berberine, an isoquinoline alkaloid derived from Berberis vulgaris (191) and used to treat bacterial, parasitic, and fungal infections, has been repurposed as a DNMT and HDAC dual inhibitor (192). In multiple myeloma cell lines, berberine treatment showed downregulated DNMT1 and DNMT3A expression, restoring p53 expression through DNA hypomethylation (193). Berberine also inhibits Class I and II HDACs in lung cancer cell lines, down-regulates gene expression, and increases histone H3 and H4 acetylation (194). EGCG is a polyphenol found in green tea (Camellia sinensis) and is a known anti-inflammatory compound (195). It has recently been proposed as an inhibitor of DNMT by direct interaction with the catalytic site of DNMT (186188). EGCG reduces cell growth and increases apoptosis in renal carcinoma cells through the upregulation of TFPI-2. In skin carcinoma cells, EGCG increases the levels of acetylation of histone H3 and histone H4 lysine residues through HDAC inhibition, leading to the upregulation of tumor-suppressor genes (188) (Table 4, Section 2). Resveratrol is a natural polyphenolic compound found in grapes and berries (196), and it has been proposed as a dual inhibitor of both DNMTs and HDACs. In breast cancer cell lines, resveratrol inhibits both HDAC and DNMT1 activity, decreases histone H3 lysine 27 methylation, and increases its acetylation (182184). In thyroid cancer cell lines, treatment with resveratrol showed resensitization to therapy when in combination with retinoic acid through the demethylation of CpG sites at promoter regions of CRABP2 gene (185); the effect of resveratrol as a repurposed cancer drug was also investigated in clinical trials (NCT00256334, NCT01476592, NCT00433576). Finally, parthenolide is a terpenoid compound, isolated from Tanacetum parthenium, with anti‐inflammatory properties. Parthenolide downregulates HDAC1 gene expression (179) and increases histone acetylation (177, 180). It reverses drug resistance in some cancer cell lines (178) and restores silenced gene expression through a decrease in DNA methylation levels (181) (Table 4, Section 2).

HDAC Inhibitors

As previously mentioned, the use of HDACi among the chemotherapeutic agents is growing (Table 5, HDACi). Hydroxamic and carboxylic acids are being studied as potential HDACi; for instance, drugs like Vorinostat (SAHA), approved for psoriasis treatment, and Valproic acid (anticonvulsant) are currently included in several clinical trials against different types of cancers (236). A complete overview about clinical trials in some of the most studied HDACi repurposed, such as Vorinostat, Valproate, Belinostat, Panobinostat, and cyclic peptide Romidepsin is available (236) (Table 5, HDACi).

TABLE 5
www.frontiersin.org

Table 5 Current HDACi repurposed drugs with applications in cancer therapy [*modified from Moreira-Silva et al. (9)].

Compounds with HDACi potential have been found in plants. Ginseng (Panax ginseng) is a popular plant extract commonly used in South Korea and traditional Chinese medicine, which contains several compounds (ginsenosides) with pharmacological properties (144). Platycodi radix (Platycodon grandiflorum), commonly known as balloon flower, is used to treat many diseases related to obesity in East Asia (237). Recently, Byun and cols. demonstrated that ginseng and platycodi have significant HDACi activity in Lung Carcinoma cell lines, thus upregulating p21 gene expression and promoting cell death (204). HC toxin is a cyclic tetrapeptide derived from a plant-fungal parasitic-association between Helminthosporium carbonum (ascomycetes) and its host, (commonly Poaceae plants family). It was reported as a Maize Histone Deacetylase inhibitor (238) and proposed as an analog of Apicidin and Artemisin, a fungal metabolite (239), and antimalarial drug, respectively; with antiprotozoal HDACi activity proved for Malaria (Plasmodium berghei) in mice. However, recently HC-toxin has been rediscovered and identified as HDACi in different cancer cell lines (205). In breast cancer and neuroblastoma cell lines, HC toxin inhibited HDAC activity and promoted cell proliferation inhibition, cellular death, and induced H4 acetylation (205, 206). Artemisin has been repurposed as an HDAC1, HDAC2, and HDAC6 inhibitor in the breast cancer cell line MCF‐7 (203) (Table 5, HDACi).

Psammaplin A (PsA) is a phenolic compound that derives from the marine sponge-association, Poecillastra sp. and Jaspis sp., (Pseudoceratina purpurea) whose active substances are monomers of thiol groups with enzymatic inhibition activity (210, 240). These monomers play a key role for both HDACi and DNMTi activity (241). In endometrial cancer cells, PsA showed HDAC1 and HDAC6 inhibition, reduction of HDAC1 expression the elevation of histone H3 and H4 acetylation, induction of cell cycle arrest, and apoptosis (208, 209). Burkholdacs A and B, with a structure similar to Thailandepsin A, was identified as a novel HDACi through the systematic overexpression of transcription factors associated with Burkholderia thilandensis (227). They are bicyclic depsipeptide compounds, proposed as potent HDACi in brain cancer cells, but also in other cancer cell lines (226). Using a panel of 39 human cancer cell lines, burkholdacs have shown superior HDACi activity over Ramidopsine (approved HDACi) in at least six cancer cell lines (226). Burkholdacs’ affinity for HDAC1 is greater than that for HDAC6. Structural changes in burkholdacs A and B structures may increase their activity and selectivity, giving rise to isoform selective inhibition of HDACs therapeutical potential (226) (Table 5, HDACi). Other depsipeptides have also been studied for repurposing. Spiruchostatin A, and Plitidepsin (Aplidin) are natural depsipeptides derived from Pseudomonas sp. (228) and Aplidium albicans (242), respectively. In cancer cell lines, reduced spiruchostatin A effectively inhibited HDAC1, an effect not observed when oxidized, and it showed an increase in the acetylation levels of specific lysine residues of histones H3 and H4 (228). Plitidepsin is currently in clinical trials to treat multiple myeloma (243, 244) but it has also displayed interesting properties against hematological malignancies (245). Some depsipeptides display a greater affinity for HDAC1 than HDAC6 and class II HDACs, but this does not appear to limit their activity as anti-cancer agents judging by in vitro effects in cancer cells (208, 226, 228). Structure-function studies on depsipeptides can lead to the generation of chemical analogs with enhanced selectivity as HDACi drugs (Table 5, HDACi).

HAT, HMT, HDM, and BET Inhibitors

Recently, HATi, HMTi, HDMi, and BETi have become of great interest for personalized cancer treatment. Multiple studies have consistently shown the enormous potential of known drugs and compounds for DR as epigenetic modulators (Table 6, HATi, HMTi, HDMi, and BETi).

TABLE 6
www.frontiersin.org

Table 6 Current HAT HMT, HDM, and BET inhibitors repurposed with epigenetic applications in cancer therapy [*modified from Moreira-Silva et al., (9)].

HAT Inhibitors

Anacardic acid, a small molecule obtained from cashew nutshell liquid with known antitumor activity, inhibits the p300’s and PCAF’s HAT activity. Anacardic Acid is not specific to any particular HAT group, but it can be used to synthesize other specific HAT activity modulators based on this molecule (246). Plumbagin is an in vivo, potent acetyltransferase inhibitor, hydroxynaphthoquinone isolated from the roots of Plumbago Rosea. A single hydroxyl group in Plumbagin confers its HATi properties. Replacing this group with other chemical moieties results in complete loss of its inhibitory activity. Plumbagin has also been reported to suppress the activation of NFκ-B, leading to apoptosis potentiation. Plumbagin may be a potential anticancer agent, but its cell toxicity properties could be the main limitation of its use as a therapeutic molecule (253). Garcinol is a potent inhibitor of the p300 and PCAF HATs. It inhibits in vivo histone acetylation in HeLa cells but does not affect histone deacetylation. Garcinol suppresses chromatin transcription dependent on HAT p300 but does not affected transcription of DNA (249). Lunasin is a 43 amino acid peptide found in soybean, barley, wheat, and rye. Previous studies have shown that lunasin can suppress the proliferation and migration of cancer cells with no effect on wild-type cells. Lunasin is a competitive inhibitor of HATs. It inhibits histone acetylation and regulates the cell cycle. This binding is probably achieved through its helical structure, similar to chromatin-binding protein structures (267) (Table 6, HATi).

HMT Inhibitors

Allantodapsone was recovered from a virtual screening based on the PRMT1 structure. Allantodapsone inhibits H4R3 methylation in the hepatocellular carcinoma cell line HepG2 while leaving H3K4 methylation unaffected (255). Ribavirin is an antiviral drug that has become of interest as a therapeutic agent in cancer. Ribavirine selectively inhibits pediatric osteosarcoma and improves chemosensitivity (256). It also possesses in vitro growth inhibitory effects against various malignant cell lines at clinically reasonable concentrations; also, ribavirin treatment results in the reduction of EZH2 at RNA and protein levels, inhibition of EZH2 enzyme activity, and reduction of H3K27 methylation (257). The anti-malarial drug, hydroxychloroquine, has also been effective in treating rheumatoid lupus, arthritis, and porphyria cutanea tarda. Structural experiments have shown that hydroxychloroquine inhibits the allosteric binding of PRC2 to EED within the H3K27me3-binding region, thereby antagonizing the catalytic function of the PRC2. These findings suggest a new epigenetic function of hydroxychloroquine with possible therapeutic repositioning (258) (Table 6, HMTi).

HDM Inhibitors

Clorgyline is a selective MAO A inhibitor- used as an antidepressant until severe dietary adverse effects are commonly known as the “cheese effect” were reported for this drug (268). As a member of MAO inhibitors, clorgyline can also inhibit LSD1, and it has been demonstrated to have cell-type dependent synergic effects when combined with DNMTi (259). Geranylgeranoic acid, an acyclic diterpenoid present in medicinal plants, has recently been found to be a potent inhibitor of recombinant LSD1. Geranylgeranoic acid inhibits the proliferation and induces a neuronal phenotype through increasing the abundance of H3K4me2 of NTRK2 gene promoter in human SH-SY5Y-derived neuroblastoma cells (260). Pargyline, a MAO B selective inhibitor with antidepressant activity, affects the transition from androgen-dependent to androgen-independent in prostate cancer. Inhibition of LSD1 with a concomitant reduction of H3K4me2 and H3K9me2 levels have been reported for pargyline. Pargyline, in combination with androgen deprivation therapy, could be an effective adjunctive treatment for advanced prostate cancer (261). Unlike selective MAO inhibitors such as pargyline, non-selective MAO inhibitors strongly repress the nucleosomal demethylation of histone H3K4. Tranylcypromine, a drug used in treating severe depression, has demonstrated strong LSD1 inhibitory effects with an IC50 of less than 2 mM (262). Tranylcypromine contributes to GBM cell synergistic apoptosis in association with other HDAC inhibitors (263). Recently, molecular docking studies have highlighted the potential of approved drugs such as decitabine, entecavir, abacavir, penciclovir, and DZNep as KDM5B inhibitors. Their role as HDMi could be of great importance in lung cancer, melanoma, hepatocellular carcinoma, gastric cancer, and prostate cancer, among others. Decitabine is a DNMTi used in myelodysplastic syndrome (MDS), abacavir, entecavir, and penciclovir are antivirals used in the treatment of HIV, hepatitis B, and herpes infections, respectively. DZNep is a specific HMTi with promising results in cancer immunotherapy (269). Finally, Polymyxin B and polymyxin E are antibiotics used in multidrug resistant bacterial infections. These compounds were shown to inhibit LSD1 by competition with its substrate at the enzyme’s cleft entry. Polymyxins have significant side effects that limit their application to untreated infections, but they could still be the target of drug repurposing for other diseases, such as leukemia (264) (Table 6, HDMi).

BET Inhibitors

Azelastine, a selective H1 antagonist, was found to be a promising BETi, displaying a stronger binding affinity than BETi control JQ1 for human BRD4 by docking-based methodologies. These findings highlight the importance of computational methods for molecular drug design and will uncover new BRD4 inhibition candidates (265). The antibiotic approved by FDA, nitroxoline, disrupts the association of BRD4 bromodomain with acetylated H4. Nitroxoline has shown strong selectivity at inhibiting all BET family members compared with non-BET proteins. By causing cell cycle arrest and apoptosis, nitroxoline successfully prevents the proliferation of MLL leukemia cells. The possible use of nitroxoline and its derivatives as BET inhibitors in BET related diseases is now under investigation (266) (Table 6, BETi).

Concluding Remarks

Drug repositioning has emerged as a viable strategy to increase drug discovery’s overall productivity, resulting in a new and cheaper way to generate alternative therapies for various diseases, including cancer. The drug repositioning approach is growing due to a broad range of reposition candidate molecules that already have clinical and toxicity profiling developments. One factor that has strongly driven this approach is the increasing availability of biomedical data, including genomic data, which covers various aspects of cellular mechanisms, opening a search that is not restricted to biological factors involved in a disease. This omic perspective allows the deduction of complex interactions that can be inhibited or treated to cure or reverse a pathological condition. Advances in complementary bioinformatic analytical methods provide critical substrate candidates that enable their systematic evaluation. Therefore, a window of opportunity opens where the reuse of previously synthesized drugs can be investigated and given a new direction. Epi-DR has already shown a profit in epigenetics and cancer treatment, where it has proven its efficacy. Indeed, many epidrugs emerged this way, such as 5-azacytidine and 5-aza-2′-deoxycytidine (decitabine) (146), Hydralazine (156), Vorinostat (SAHA), and Valproic acid (236).

Epigenetic alterations are considered to be among the earliest and most comprehensive genomic aberrations occurring during carcinogenesis, and therefore it has been classified as a hallmark of cancer (270). The impact of epigenetics in understanding cancer has been of great interest in recent years, and even more due to the advancement of the genomic era. Several works demonstrate the importance of epigenetic biomarkers that can predict the response or prognosis in various types of cancer. The promoter methylation of the MGMT gene in gliomas is a clear example, where it helps to indicate the use of precision medicine through the drug temozolomide (271). Another example is found in EHZ2 enzyme alterations, which indicate a poor prognosis in breast, prostate, and other types of cancers.

Epigenetic mechanisms have great flexibility to respond to environmental changes and modify gene expression. Consequently, search for artificial ways to induce epigenetic remodeling, which could improve therapy in the event of a disease as cancer. Therefore, the implementation of epigenetic therapies opens a new panorama for the fight against cancer. Epidrugs show enormous potential for clinical use, especially in cancer, because in these diseases, an epigenetic imbalance is a well-known characteristic that is both of origin, development, and severity of tumors.

Even though there are already some epidrugs approved by the FDA and the current knowledge about various mechanisms involved in gene regulation, promoted by the advancement of technologies that expand the information on specific epigenetic mechanisms, challenges remain in identifying epigenetic modifications of cancer and targeting them for therapeutic purposes. Among them stands out that epigenetic changes can be diverse in the types of cancer and between the different clinical phases and those that are dependent on environmental conditions. Therefore, we must distinguish between the dysregulation of driver genes and those whose changes are secondary to these. Also, the generation of epigenetic therapies as well as the molecular mechanisms that coordinate them is subject to understanding, and much research is still required of several of them to safely transport them to the clinic. However, identifying epigenetic alterations that affect the tumor’s fate and behavior finding drugs that target them are some of the promises of epigenetic therapy in cancer.

In this sense, the concept of reusing a medicine offers a broad scope to investigate the hidden potential behind the medicine and to recycle it. The reincorporation of a drug with the potential to remodel epigenetic characteristics, which are beneficial for cancer management, is of great interest to the field. Offering great advantages in drug development times could lead to precision medicine therapy with new and clearly encouraging prospects for the future (Figure 4).

Author Contributions

MM-C and MM-R wrote and designed the manuscript. RG-B coordinated, wrote, and designed the manuscript. RG-B and VJ-G revised the manuscript. VJ-G elaborated on the figures. CA-C wrote and revised the manuscript, and LH coordinated and directed the review development. All authors contributed to the article and approved the submitted version.

Funding

This work was supported by the Consejo Nacional de Ciencia y Tecnología (CONACyT) by the Fondo Sectorial de Investigación en Salud y Seguridad Social (FOSISS, grant no. SALUD-2017-2-290041). Marco Antonio Meraz-Rodriguez is a masters student in the “Programa de Maestría y Doctorado en Ciencias Bioquímicas, UNAM”, and received a fellowship from CONACyT (CVU 659273, no. 481908).

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank the National Cancer Institute of Mexico (INCan) for support to the present work. We thank Juan F. Duarte-Campos and Hugo R. Barajas for their critical comments and review.

References

1. Arrowsmith CH, Bountra C, Fish PV, Lee K, Schapira M. Epigenetic protein families: a new frontier for drug discovery. Nat Rev Drug Discovery (2012) 11(5):384–400. doi: 10.1038/nrd3674

CrossRef Full Text | Google Scholar

2. Fardi M, Solali S, Farshdousti Hagh M. Epigenetic mechanisms as a new approach in cancer treatment: An updated review. Genes Dis (2018) 5(4):304–11. doi: 10.1016/j.gendis.2018.06.003

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Kelly TK, De Carvalho DD, Jones PA. Epigenetic Modifications as Therapeutic Targets. Nat Biotechnol (2010) 28(10):1069–78. doi: 10.1038/nbt.1678

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Jerónimo C, Bastian PJ, Bjartell A, Carbone GM, Catto JWF, Clark SJ, et al. Epigenetics in Prostate Cancer: Biologic and Clinical Relevance. Eur Urol (2011) 60(4):753–66. doi: 10.1016/j.eururo.2011.06.035

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Strauss J, Figg WD. Using Epigenetic Therapy to Overcome Chemotherapy Resistance. Anticancer Res (2016) 36(1):1–4.

PubMed Abstract | Google Scholar

6. Miranda Furtado CL, Dos Santos Luciano MC, Da Silva Santos R, Furtado GP, Moraes MO, Pessoa C. Epidrugs: targeting epigenetic marks in cancer treatment. Epigenetics (2019) 14(12):1164–76. doi: 10.1080/15592294.2019.1640546

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Rodríguez-Paredes M, Esteller M. Cancer epigenetics reaches mainstream oncology. Nat Med (2011) 17(3):330–9. doi: 10.1038/nm.2305

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Naveja JJ, Dueñas-González A, Medina-Franco JL. Drug Repurposing for Epigenetic Targets Guided by Computational Methods. In: Epi-Informatics. Elsevier (2016). p. 327–57. doi: 10.1016/B978-0-12-802808-7.00012-5

CrossRef Full Text | Google Scholar

9. Moreira-Silva F, Camilo V, Gaspar V, Mano JF, Henrique R, Jerónimo C. Repurposing Old Drugs into New Epigenetic Inhibitors: Promising Candidates for Cancer Treatment? Pharmaceutics (2020) 12(5):410. doi: 10.3390/pharmaceutics12050410

CrossRef Full Text | Google Scholar

10. Blatt J, Corey SJ. Drug repurposing in pediatrics and pediatric hematology oncology. Drug Discovery Today (2013) 18(1–2):4–10. doi: 10.1016/j.drudis.2012.07.009

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Shim JS, Liu JO. Recent Advances in Drug Repositioning for the Discovery of New Anticancer Drugs. Int J Biol Sci (2014) 10(7):654–63. doi: 10.7150/ijbs.9224

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Waddington CH. The epigenotype. 1942. Int J Epidemiol (2012) 41(1):10–3. doi: 10.1093/ije/dyr184

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Deans C, Maggert KA. What do you mean, ‘epigenetic’? Genetics (2015) 199(4):887–96. doi: 10.1534/genetics.114.173492

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Ganesan A, Arimondo PB, Rots MG, Jeronimo C, Berdasco M. The timeline of epigenetic drug discovery: from reality to dreams. Clin Epigenet (2019) 11:1–2. doi: 10.1186/s13148-019-0776-0

CrossRef Full Text | Google Scholar

15. Roberti A, Valdes AF, Torrecillas R, Fraga MF, Fernandez AF. Epigenetics in cancer therapy and nanomedicine. Clin Epigenet (2019) 11(1):81. doi: 10.1186/s13148-019-0675-4

CrossRef Full Text | Google Scholar

16. el Bahhaj F, Dekker FJ, Martinet N, Bertrand P. Delivery of epidrugs. Drug Discovery Today (2014) 19(9):1337–52. doi: 10.1016/j.drudis.2014.03.017

PubMed Abstract | CrossRef Full Text | Google Scholar

17. Biswas S, Rao CM. Epigenetic tools (The Writers, The Readers and The Erasers) and their implications in cancer therapy. Eur J Pharmacol (2018) 837:8–24. doi: 10.1016/j.ejphar.2018.08.021

PubMed Abstract | CrossRef Full Text | Google Scholar

18. Portela A, Esteller M. Epigenetic modifications and human disease. Nat Biotechnol (2010) 28(10):1057–68. doi: 10.1038/nbt.1685

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Moore LD, Le T, Fan G. DNA Methylation and Its Basic Function. Neuropsychopharmacology (2013) 38(1):23–38. doi: 10.1038/npp.2012.112

PubMed Abstract | CrossRef Full Text | Google Scholar

20. Kuroda A, Rauch TA, Todorov I, Ku HT, Al-Abdullah IH, Kandeel F, et al. Insulin Gene Expression Is Regulated by DNA Methylation. PloS One (2009) 4(9):e6953. doi: 10.1371/journal.pone.0006953 Maedler K (ed.).

PubMed Abstract | CrossRef Full Text | Google Scholar

21. Esteller M. Epigenetic gene silencing in cancer: the DNA hypermethylome. Hum Mol Genet (2007) 16:R50–59. doi: 10.1093/hmg/ddm018

PubMed Abstract | CrossRef Full Text | Google Scholar

22. Suzuki MM, Bird A. DNA methylation landscapes: provocative insights from epigenomics. Nat Rev Genet (2008) 9(6):465–76. doi: 10.1038/nrg2341

PubMed Abstract | CrossRef Full Text | Google Scholar

23. Smith BC, Denu JM. Chemical mechanisms of histone lysine and arginine modifications. Biochim Biophys Acta (2009) 1789(1):45–57. doi: 10.1016/j.bbagrm.2008.06.005

PubMed Abstract | CrossRef Full Text | Google Scholar

24. Gros C, Fahy J, Halby L, Dufau I, Erdmann A, Gregoire J-M, et al. DNA methylation inhibitors in cancer: recent and future approaches. Biochimie (2012) 94(11):2280–96. doi: 10.1016/j.biochi.2012.07.025

PubMed Abstract | CrossRef Full Text | Google Scholar

25. Peterson CL, Laniel M-A. Histones and histone modifications. Curr Biol: CB (2004) 14(14):R546–551. doi: 10.1016/j.cub.2004.07.007

CrossRef Full Text | Google Scholar

26. Wang Z, Schones DE, Zhao K. Characterization of human epigenomes. Curr Opin Genet Dev (2009) 19(2):127–34. doi: 10.1016/j.gde.2009.02.001

PubMed Abstract | CrossRef Full Text | Google Scholar

27. Swygert SG, Peterson CL. Chromatin dynamics: interplay between remodeling enzymes and histone modifications. Biochim Et Biophys Acta (2014) 1839(8):728–36. doi: 10.1016/j.bbagrm.2014.02.013

CrossRef Full Text | Google Scholar

28. de Lera AR, Ganesan A. Epigenetic polypharmacology: from combination therapy to multitargeted drugs. Clin Epigenet (2016) 8:4–9. doi: 10.1186/s13148-016-0271-9

CrossRef Full Text | Google Scholar

29. Hodawadekar SC, Marmorstein R. Chemistry of acetyl transfer by histone modifying enzymes: structure, mechanism and implications for effector design. Oncogene (2007) 26(37):5528–40. doi: 10.1038/sj.onc.1210619

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Dancy BM, Cole PA. Protein Lysine Acetylation by p300/CBP. Chem Rev (2015) 115(6):2419–52. doi: 10.1021/cr500452k

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Zhang Y, Fang H, Jiao J, Xu W. The structure and function of histone deacetylases: the target for anti-cancer therapy. Curr Medicinal Chem (2008) 15(27):2840–9. doi: 10.2174/092986708786242796

CrossRef Full Text | Google Scholar

32. Zhu P, Martin E, Mengwasser J, Schlag P, Janssen K-P, Göttlicher M. Induction of HDAC2 expression upon loss of APC in colorectal tumorigenesis. Cancer Cell (2004) 5(5):455–63. doi: 10.1016/s1535-6108(04)00114-x

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Ropero S, Fraga MF, Ballestar E, Hamelin R, Yamamoto H, Boix-Chornet M, et al. A truncating mutation of HDAC2 in human cancers confers resistance to histone deacetylase inhibition. Nat Genet (2006) 38(5):566–9. doi: 10.1038/ng1773

PubMed Abstract | CrossRef Full Text | Google Scholar

34. You JS, Jones PA. Cancer genetics and epigenetics: two sides of the same coin? Cancer Cell (2012) 22(1):9–20. doi: 10.1016/j.ccr.2012.06.008

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Fraga MF, Ballestar E, Villar-Garea A, Boix-Chornet M, Espada J, Schotta G, et al. Loss of acetylation at Lys16 and trimethylation at Lys20 of histone H4 is a common hallmark of human cancer. Nat Genet (2005) 37(4):391–400. doi: 10.1038/ng1531

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Sharma S, Kelly TK, Jones PA. Epigenetics in cancer. Carcinogenesis (2010) 31(1):27–36. doi: 10.1093/carcin/bgp220

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Feng Q, Wang H, Ng HH, Erdjument-Bromage H, Tempst P, Struhl K, et al. Methylation of H3-Lysine 79 Is Mediated by a New Family of HMTases without a SET Domain. Curr Biol (2002) 12(12):1052–8. doi: 10.1016/S0960-9822(02)00901-6

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Nguyen AT, Zhang Y. The diverse functions of Dot1 and H3K79 methylation. Genes Dev (2011) 25(13):1345–58. doi: 10.1101/gad.2057811

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Karlić R, Chung H-R, Lasserre J, Vlahovicek K, Vingron M. Histone modification levels are predictive for gene expression. Proc Natl Acad Sci U S A (2010) 107(7):2926–31. doi: 10.1073/pnas.0909344107

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Shi Y, Whetstine JR. Dynamic regulation of histone lysine methylation by demethylases. Mol Cell (2007) 25(1):1–14. doi: 10.1016/j.molcel.2006.12.010

PubMed Abstract | CrossRef Full Text | Google Scholar

41. Dalgliesh GL, Furge K, Greenman C, Chen L, Bignell G, Butler A, et al. Systematic sequencing of renal carcinoma reveals inactivation of histone modifying genes. Nature (2010) 463(7279):360–3. doi: 10.1038/nature08672

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Hamamoto R, Furukawa Y, Morita M, Iimura Y, Silva FP, Li M, et al. SMYD3 encodes a histone methyltransferase involved in the proliferation of cancer cells. Nat Cell Biol (2004) 6(8):731–40. doi: 10.1038/ncb1151

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Kondo Y, Shen L, Suzuki S, Kurokawa T, Masuko K, Tanaka Y, et al. Alterations of DNA methylation and histone modifications contribute to gene silencing in hepatocellular carcinomas. Hepatol Res: Off J Japan Soc Hepatol (2007) 37(11):974–83. doi: 10.1111/j.1872-034X.2007.00141.x

CrossRef Full Text | Google Scholar

44. Füllgrabe J, Kavanagh E, Joseph B. Histone onco-modifications. Oncogene (2011) 30(31):3391–403. doi: 10.1038/onc.2011.121

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Berdasco M, Esteller M. Clinical epigenetics: seizing opportunities for translation. Nat Rev Genet (2019) 20(2):109–27. doi: 10.1038/s41576-018-0074-2

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Lv J-F, Hu L, Zhuo W, Zhang C-M, Zhou H-H, Fan L. Epigenetic alternations and cancer chemotherapy response. Cancer Chemother Pharmacol (2016) 77(4):673–84. doi: 10.1007/s00280-015-2951-0

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Evans WE, Johnson JA. Pharmacogenomics: the inherited basis for interindividual differences in drug response. Annu Rev Genomics Hum Genet (2001) 2:9–39. doi: 10.1146/annurev.genom.2.1.9

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Sharma SV, Lee DY, Li B, Quinlan MP, Takahashi F, Maheswaran S, et al. A chromatin-mediated reversible drug-tolerant state in cancer cell subpopulations. Cell (2010) 141(1):69–80. doi: 10.1016/j.cell.2010.02.027

PubMed Abstract | CrossRef Full Text | Google Scholar

49. El-Khoury V, Breuzard G, Fourré N, Dufer J. The histone deacetylase inhibitor trichostatin A downregulates human MDR1 (ABCB1) gene expression by a transcription-dependent mechanism in a drug-resistant small cell lung carcinoma cell line model. Br J Cancer (2007) 97(4):562–73. doi: 10.1038/sj.bjc.6603914

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Kumar S, Kushwaha PP, Gupta S. Emerging targets in cancer drug resistance. Cancer Drug Resist (2019) 2(2):161–77. doi: 10.20517/cdr.2018.27

CrossRef Full Text | Google Scholar

51. Salarinia R, Sahebkar A, Peyvandi M, Reza Mirzaei H, Reza Jaafari M, Matbou Riahi M, et al. Epi-Drugs and Epi-miRs: Moving Beyond Current Cancer Therapies. Curr Cancer Drug Targets (2016) 16(9):773–88. doi: 10.2174/1568009616666151207110143

PubMed Abstract | CrossRef Full Text | Google Scholar

52. Lauschke VM, Barragan I, Ingelman-Sundberg M. Pharmacoepigenetics and Toxicoepigenetics: Novel Mechanistic Insights and Therapeutic Opportunities. Annu Rev Pharmacol Toxicol (2018) 58(1):161–85. doi: 10.1146/annurev-pharmtox-010617-053021

PubMed Abstract | CrossRef Full Text | Google Scholar

53. Brown R, Curry E, Magnani L, Wilhelm-Benartzi CS, Borley J. Poised epigenetic states and acquired drug resistance in cancer. Nat Rev Cancer (2014) 14(11):747–53. doi: 10.1038/nrc3819

PubMed Abstract | CrossRef Full Text | Google Scholar

54. Kumar R, Harilal S, Gupta SV, Jose J, Thomas Parambi DG, Uddin M, et al. Exploring the new horizons of drug repurposing: A vital tool for turning hard work into smart work. Eur J Medicinal Chem (2019) 182:111602. doi: 10.1016/j.ejmech.2019.111602

CrossRef Full Text | Google Scholar

55. Raynal NJ-M, Da Costa EM, Lee JT, Gharibyan V, Ahmed S, Zhang H, et al. Repositioning FDA-approved drugs in combination with epigenetic drugs to reprogram colon cancer epigenome. Mol Cancer Ther (2017) 16(2):397–407. doi: 10.1158/1535-7163.MCT-16-0588

PubMed Abstract | CrossRef Full Text | Google Scholar

56. Nebbioso A, Carafa V, Benedetti R, Altucci L. Trials with ‘epigenetic’ drugs: an update. Mol Oncol (2012) 6(6):657–82. doi: 10.1016/j.molonc.2012.09.004

PubMed Abstract | CrossRef Full Text | Google Scholar

57. Cartron P-F, Cheray M, Bretaudeau L. Epigenetic protein complexes: the adequate candidates for the use of a new generation of epidrugs in personalized and precision medicine in cancer. Epigenomics (2019) 12(2):171–7. doi: 10.2217/epi-2019-0169

PubMed Abstract | CrossRef Full Text | Google Scholar

58. Bennett RL, Licht JD. Targeting Epigenetics in Cancer. Annu Rev Pharmacol Toxicol (2018) 58:187–207. doi: 10.1146/annurev-pharmtox-010716-105106

PubMed Abstract | CrossRef Full Text | Google Scholar

59. Morel D, Jeffery D, Aspeslagh S, Almouzni G, Postel-Vinay S. Combining epigenetic drugs with other therapies for solid tumours - past lessons and future promise. Nat Rev Clin Oncol (2020) 17(2):91–107. doi: 10.1038/s41571-019-0267-4

PubMed Abstract | CrossRef Full Text | Google Scholar

60. Erdmann A, Halby L, Fahy J, Arimondo PB. Targeting DNA methylation with small molecules: what’s next? J Medicinal Chem (2015) 58(6):2569–83. doi: 10.1021/jm500843d

CrossRef Full Text | Google Scholar

61. Leone G, Teofili L, Voso MT, Lübbert M. DNA methylation and demethylating drugs in myelodysplastic syndromes and secondary leukemias. Haematologica (2002) 87(12):1324–41.

PubMed Abstract | Google Scholar

62. Gaulton A, Hersey A, Nowotka M, Bento AP, Chambers J, Mendez D, et al. The ChEMBL database in 2017. Nucleic Acids Res (2017) 45(D1):D945–54. doi: 10.1093/nar/gkw1074

PubMed Abstract | CrossRef Full Text | Google Scholar

63. Sébert M, Renneville A, Bally C, Peterlin P, Beyne-Rauzy O, Legros L, et al. A phase II study of guadecitabine in higher-risk myelodysplastic syndrome and low blast count acute myeloid leukemia after azacitidine failure. Haematologica (2019) 104(8):1565–71. doi: 10.3324/haematol.2018.207118

PubMed Abstract | CrossRef Full Text | Google Scholar

64. Azad N, Zahnow CA, Rudin CM, Baylin SB. The future of epigenetic therapy in solid tumours–lessons from the past. Nat Rev Clin Oncol (2013) 10(5):256–66. doi: 10.1038/nrclinonc.2013.42

PubMed Abstract | CrossRef Full Text | Google Scholar

65. Zhang J, Zheng YG. SAM/SAH Analogs as Versatile Tools for SAM-Dependent Methyltransferases. ACS Chem Biol (2016) 11(3):583–97. doi: 10.1021/acschembio.5b00812

PubMed Abstract | CrossRef Full Text | Google Scholar

66. Asgatay S, Champion C, Marloie G, Drujon T, Senamaud-Beaufort C, Ceccaldi A, et al. Synthesis and evaluation of analogues of N-phthaloyl-l-tryptophan (RG108) as inhibitors of DNA methyltransferase 1. J Medicinal Chem (2014) 57(2):421–34. doi: 10.1021/jm401419p

CrossRef Full Text | Google Scholar

67. Candelaria M, de la Cruz-Hernandez E, Taja-Chayeb L, Perez-Cardenas E, Trejo-Becerril C, Gonzalez-Fierro A, et al. DNA Methylation-Independent Reversion of Gemcitabine Resistance by Hydralazine in Cervical Cancer Cells. PloS One (2012) 7(3):e29181. doi: 10.1371/journal.pone.0029181

PubMed Abstract | CrossRef Full Text | Google Scholar

68. Ceccaldi A, Rajavelu A, Champion C, Rampon C, Jurkowska R, Jankevicius G, et al. C5-DNA Methyltransferase Inhibitors: From Screening to Effects on Zebrafish Embryo Development. ChemBioChem (2011) 12(9):1337–45. doi: 10.1002/cbic.201100130

PubMed Abstract | CrossRef Full Text | Google Scholar

69. Davis AJ, Gelmon KA, Siu LL, Moore MJ, Britten CD, Mistry N, et al. and pharmacologic study of the human DNA methyltransferase antisense oligodeoxynucleotide MG98 given as a 21-day continuous infusion every 4 weeks. Invest New Drugs (2003) 21(1):85–97. doi: 10.1023/a:1022976528441

PubMed Abstract | CrossRef Full Text | Google Scholar

70. Lin J, Haffner MC, Zhang Y, Lee BH, Brennen WN, Britton J, et al. Disulfiram is a DNA demethylating agent and inhibits prostate cancer cell growth. Prostate (2011) 71(4):333–43. doi: 10.1002/pros.21247

PubMed Abstract | CrossRef Full Text | Google Scholar

71. Lu Y, Chan Y-T, Tan H-Y, Li S, Wang N, Feng Y. Epigenetic regulation in human cancer: the potential role of epi-drug in cancer therapy. Mol Cancer (2020) 19:6–8. doi: 10.1186/s12943-020-01197-3

PubMed Abstract | CrossRef Full Text | Google Scholar

72. Yang X, Lay F, Han H, Jones PA. Targeting DNA methylation for epigenetic therapy. Trends Pharmacol Sci (2010) 31(11):536–46. doi: 10.1016/j.tips.2010.08.001

PubMed Abstract | CrossRef Full Text | Google Scholar

73. Archin NM, Kirchherr JL, Sung JAM, Clutton G, Sholtis K, Xu Y, et al. Interval dosing with the HDAC inhibitor vorinostat effectively reverses HIV latency. J Clin Invest (2020) 127(8):3126–35. doi: 10.1172/JCI92684

CrossRef Full Text | Google Scholar

74. Jiang C, Lian X, Gao C, Sun X, Einkauf KB, Chevalier JM, et al. Distinct viral reservoirs in individuals with spontaneous control of HIV-1. Nature (2020) 585:1–7. doi: 10.1038/s41586-020-2651-8

CrossRef Full Text | Google Scholar

75. Connolly RM, Li H, Jankowitz RC, Zhang Z, Rudek MA, Jeter SC, et al. Combination Epigenetic Therapy in Advanced Breast Cancer with 5-Azacitidine and Entinostat: A Phase II National Cancer Institute/Stand Up to Cancer Study. Clin Cancer Res: Off J Am Assoc Cancer Res (2017) 23(11):2691–701. doi: 10.1158/1078-0432.CCR-16-1729

CrossRef Full Text | Google Scholar

76. Juergens RA, Wrangle J, Vendetti FP, Murphy SC, Zhao M, Coleman B, et al. Combination epigenetic therapy has efficacy in patients with refractory advanced non-small cell lung cancer. Cancer Discovery (2011) 1(7):598–607. doi: 10.1158/2159-8290.CD-11-0214

PubMed Abstract | CrossRef Full Text | Google Scholar

77. Mai A, Altucci L. Epi-drugs to fight cancer: from chemistry to cancer treatment, the road ahead. Int J Biochem Cell Biol (2009) 41(1):199–213. doi: 10.1016/j.biocel.2008.08.020

PubMed Abstract | CrossRef Full Text | Google Scholar

78. Pontiki E, Hadjipavlou-Litina D. Histone deacetylase inhibitors (HDACIs). Structure–activity relationships: history and new QSAR perspectives. Medicinal Res Rev (2012) 32(1):1–165. doi: 10.1002/med.20200

CrossRef Full Text | Google Scholar

79. Jaboin J, Wild J, Hamidi H, Khanna C, Kim CJ, Robey R, et al. MS-27-275, an inhibitor of histone deacetylase, has marked in vitro and in vivo antitumor activity against pediatric solid tumors. Cancer Res (2002) 62(21):6108–15.

PubMed Abstract | Google Scholar

80. Prachayasittikul V, Prathipati P, Pratiwi R, Phanus-Umporn C, Malik AA, Schaduangrat N, et al. Exploring the epigenetic drug discovery landscape. Expert Opin Drug Discovery (2017) 12(4):345–62. doi: 10.1080/17460441.2017.1295954

CrossRef Full Text | Google Scholar

81. Yardley DA, Ismail-Khan RR, Melichar B, Lichinitser M, Munster PN, Klein PM, et al. Randomized phase II, double-blind, placebo-controlled study of exemestane with or without entinostat in postmenopausal women with locally recurrent or metastatic estrogen receptor-positive breast cancer progressing on treatment with a nonsteroidal aromatase inhibitor. J Clin Oncol: Off J Am Soc Clin Oncol (2013) 31(17):2128–35. doi: 10.1200/JCO.2012.43.7251

CrossRef Full Text | Google Scholar

82. Tomaselli D, Lucidi A, Rotili D, Mai A. Epigenetic polypharmacology: A new frontier for epi-drug discovery. Medicinal Res Rev (2020) 40(1):190–244. doi: 10.1002/med.21600

CrossRef Full Text | Google Scholar

83. Mahajan SS, Leko V, Simon JA, Bedalov A. Sirtuin modulators. Handb Exp Pharmacol (2011) 206:241–55. doi: 10.1007/978-3-642-21631-2_11

PubMed Abstract | CrossRef Full Text | Google Scholar

84. Copeland RA, Moyer MP, Richon VM. Targeting genetic alterations in protein methyltransferases for personalized cancer therapeutics. Oncogene (2013) 32(8):939–46. doi: 10.1038/onc.2012.552

PubMed Abstract | CrossRef Full Text | Google Scholar

85. Han D, Huang M, Wang T, Li Z, Chen Y, Liu C, et al. Lysine methylation of transcription factors in cancer. Cell Death Dis (2019) 10(4):1–11. doi: 10.1038/s41419-019-1524-2

CrossRef Full Text | Google Scholar

86. Morin RD, Johnson NA, Severson TM, Mungall AJ, An J, Goya R, et al. Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin. Nat Genet (2010) 42(2):181–5. doi: 10.1038/ng.518

PubMed Abstract | CrossRef Full Text | Google Scholar

87. Sermer D, Pasqualucci L, Wendel H-G, Melnick A, Younes A. Emerging epigenetic-modulating therapies in lymphoma. Nat Rev Clin Oncol (2019) 16(8):494–507. doi: 10.1038/s41571-019-0190-8

PubMed Abstract | CrossRef Full Text | Google Scholar

88. Sneeringer CJ, Scott MP, Kuntz KW, Knutson SK, Pollock RM, Richon VM, et al. Coordinated activities of wild-type plus mutant EZH2 drive tumor-associated hypertrimethylation of lysine 27 on histone H3 (H3K27) in human B-cell lymphomas. Proc Natl Acad Sci (2010) 107(49):20980–5. doi: 10.1073/pnas.1012525107

PubMed Abstract | CrossRef Full Text | Google Scholar

89. Béguelin W, Popovic R, Teater M, Jiang Y, Bunting KL, Rosen M, et al. EZH2 is required for germinal center formation and somatic EZH2 mutations promote lymphoid transformation. Cancer Cell (2013) 23(5):677–92. doi: 10.1016/j.ccr.2013.04.011

PubMed Abstract | CrossRef Full Text | Google Scholar

90. Li X, Wang C, Jiang H, Luo C. A patent review of arginine methyltransferase inhibitors (2010-2018). Expert Opin Ther Patents (2019) 29(2):97–114. doi: 10.1080/13543776.2019.1567711

CrossRef Full Text | Google Scholar

91. Rose NR, Ng SS, Mecinović J, Liénard BMR, Bello SH, Sun Z, et al. Inhibitor Scaffolds for 2-Oxoglutarate-Dependent Histone Lysine Demethylases. J Medicinal Chem (2008) 51(22):7053–6. doi: 10.1021/jm800936s

CrossRef Full Text | Google Scholar

92. Joberty G, Boesche M, Brown JA, Eberhard D, Garton NS, Humphreys PG, et al. Interrogating the Druggability of the 2-Oxoglutarate-Dependent Dioxygenase Target Class by Chemical Proteomics. ACS Chem Biol (2016) 11(7):2002–10. doi: 10.1021/acschembio.6b00080

PubMed Abstract | CrossRef Full Text | Google Scholar

93. Shi Y, Lan F, Matson C, Mulligan P, Whetstine JR, Cole PA, et al. Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell (2004) 119(7):941–53. doi: 10.1016/j.cell.2004.12.012

PubMed Abstract | CrossRef Full Text | Google Scholar

94. Niebel D, Kirfel J, Janzen V, Höller T, Majores M, Gütgemann I. Lysine-specific demethylase 1 (LSD1) in hematopoietic and lymphoid neoplasms. Blood (2014) 124(1):151–2. doi: 10.1182/blood-2014-04-569525

PubMed Abstract | CrossRef Full Text | Google Scholar

95. Yang G-J, Lei P-M, Wong S-Y, Ma D-L, Leung C-H. Pharmacological Inhibition of LSD1 for Cancer Treatment. Molecules (2018) 23(12):9–11. doi: 10.3390/molecules23123194

CrossRef Full Text | Google Scholar

96. Shih JC, Chen K, Ridd MJ. Monoamine oxidase: from genes to behavior. Annu Rev Neurosci (1999) 22:197–217. doi: 10.1146/annurev.neuro.22.1.197

PubMed Abstract | CrossRef Full Text | Google Scholar

97. Yang M, Culhane JC, Szewczuk LM, Jalili P, Ball HL, Machius M, et al. Structural basis for the inhibition of the LSD1 histone demethylase by the antidepressant trans-2-phenylcyclopropylamine. Biochemistry (2007) 46(27):8058–65. doi: 10.1021/bi700664y

PubMed Abstract | CrossRef Full Text | Google Scholar

98. Zheng Y-C, Yu B, Chen Z-S, Liu Y, Liu H-M. TCPs: privileged scaffolds for identifying potent LSD1 inhibitors for cancer therapy. Epigenomics (2016) 8(5):651–66. doi: 10.2217/epi-2015-0002

PubMed Abstract | CrossRef Full Text | Google Scholar

99. Shortt J, Ott CJ, Johnstone RW, Bradner JE. A chemical probe toolbox for dissecting the cancer epigenome. Nat Rev Cancer (2017) 17(4):268. doi: 10.1038/nrc.2017.26

PubMed Abstract | CrossRef Full Text | Google Scholar

100. Doroshow DB, Eder JP, LoRusso PM. BET inhibitors: a novel epigenetic approach. Ann Oncol: Off J Eur Soc Med Oncol (2017) 28(8):1776–87. doi: 10.1093/annonc/mdx157

CrossRef Full Text | Google Scholar

101. Zuber J, Shi J, Wang E, Rappaport AR, Herrmann H, Sison EA, et al. RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia. Nature (2011) 478(7370):524–8. doi: 10.1038/nature10334

PubMed Abstract | CrossRef Full Text | Google Scholar

102. Morel D, Almouzni G, Soria J-C, Postel-Vinay S. Targeting chromatin defects in selected solid tumors based on oncogene addiction, synthetic lethality and epigenetic antagonism. Ann Oncol: Off J Eur Soc Med Oncol (2017) 28(2):254–69. doi: 10.1093/annonc/mdw552

CrossRef Full Text | Google Scholar

103. Winkler J, Raina K, Altieri M, Dong H, Wang J, Chen X, et al. PROTAC BET degraders are more broadly effective than BET inhibitors. Eur J Cancer (2016) 69:S10. doi: 10.1016/S0959-8049(16)32621-1

CrossRef Full Text | Google Scholar

104. Kummar S, Chen HX, Wright J, Holbeck S, Millin MD, Tomaszewski J, et al. Utilizing targeted cancer therapeutic agents in combination: novel approaches and urgent requirements. Nat Rev Drug Discovery (2010) 9(11):843–56. doi: 10.1038/nrd3216

CrossRef Full Text | Google Scholar

105. Li YY, Jones SJ. Drug repositioning for personalized medicine. Genome Med (2012) 4(3):27. doi: 10.1186/gm326

PubMed Abstract | CrossRef Full Text | Google Scholar

106. Hopkins AL. Drug discovery: Predicting promiscuity. Nature (2009) 462(7270):167–8. doi: 10.1038/462167a

PubMed Abstract | CrossRef Full Text | Google Scholar

107. Ashburn TT, Thor KB. Drug repositioning: identifying and developing new uses for existing drugs. Nat Rev Drug Discovery (2004) 3(8):673–83. doi: 10.1038/nrd1468

CrossRef Full Text | Google Scholar

108. Pantziarka P. Scientific advice - is drug repurposing missing a trick? Nature Reviews. Clin Oncol (2017) 14(8):455–6. doi: 10.1038/nrclinonc.2017.69

CrossRef Full Text | Google Scholar

109. Pushpakom S, Iorio F, Eyers PA, Escott KJ, Hopper S, Wells A, et al. Drug repurposing: progress, challenges and recommendations. Nat Rev Drug Discovery (2019) 18(1):41–58. doi: 10.1038/nrd.2018.168

CrossRef Full Text | Google Scholar

110. YCharts. Pfizer’s Expiring Viagra Patent Adversely Affects Other Drugmakers Too. Jersey City, NJ: Forbes (2020).

Google Scholar

111. Urquhart L. Market watch: Top drugs and companies by sales in 2017. Nat Rev Drug Discovery (2018) 17(4):232. doi: 10.1038/nrd.2018.42

CrossRef Full Text | Google Scholar

112. Tartaglia LA. Complementary new approaches enable repositioning of failed drug candidates. Expert Opin Invest Drugs (2006) 15(11):1295–8. doi: 10.1517/13543784.15.11.1295

CrossRef Full Text | Google Scholar

113. Brehmer D, Greff Z, Godl K, Blencke S, Kurtenbach A, Weber M, et al. Cellular Targets of Gefitinib. Cancer Res (2005) 65(2):379–82.

PubMed Abstract | Google Scholar

114. Moffat JG, Vincent F, Lee JA, Eder J, Prunotto M. Opportunities and challenges in phenotypic drug discovery: an industry perspective. Nat Rev Drug Discovery (2017) 16(8):531–43. doi: 10.1038/nrd.2017.111

CrossRef Full Text | Google Scholar

115. Chong CR, Xu J, Lu J, Bhat S, Sullivan DJ, Liu JO. Inhibition of angiogenesis by the antifungal drug itraconazole. ACS Chem Biol (2007) 2(4):263–70. doi: 10.1021/cb600362d

PubMed Abstract | CrossRef Full Text | Google Scholar

116. Sundberg SA. High-throughput and ultra-high-throughput screening: solution- and cell-based approaches. Curr Opin Biotechnol (2000) 11(1):47–53. doi: 10.1016/s0958-1669(99)00051-8

PubMed Abstract | CrossRef Full Text | Google Scholar

117. Reaume AG. Drug repurposing through nonhypothesis driven phenotypic screening. Drug Discovery Today: Ther Strategies (2011) 8(3–4):85–8. doi: 10.1016/j.ddstr.2011.09.007

CrossRef Full Text | Google Scholar

118. Hurle MR, Yang L, Xie Q, Rajpal DK, Sanseau P, Agarwal P. Computational drug repositioning: from data to therapeutics. Clin Pharmacol Ther (2013) 93(4):335–41. doi: 10.1038/clpt.2013.1

PubMed Abstract | CrossRef Full Text | Google Scholar

119. Hieronymus H, Lamb J, Ross KN, Peng XP, Clement C, Rodina A, et al. Gene expression signature-based chemical genomic prediction identifies a novel class of HSP90 pathway modulators. Cancer Cell (2006) 10(4):321–30. doi: 10.1016/j.ccr.2006.09.005

PubMed Abstract | CrossRef Full Text | Google Scholar

120. Kitchen DB, Decornez H, Furr JR, Bajorath J. Docking and scoring in virtual screening for drug discovery: methods and applications. Nat Rev Drug Discovery (2004) 3(11):935–49. doi: 10.1038/nrd1549

CrossRef Full Text | Google Scholar

121. Sanseau P, Agarwal P, Barnes MR, Pastinen T, Richards JB, Cardon LR, et al. Use of genome-wide association studies for drug repositioning. Nat Biotechnol (2012) 30(4):317–20. doi: 10.1038/nbt.2151

PubMed Abstract | CrossRef Full Text | Google Scholar

122. Greene CS, Voight BF. Pathway and network-based strategies to translate genetic discoveries into effective therapies. Hum Mol Genet (2016) 25(R2):R94–8. doi: 10.1093/hmg/ddw160

PubMed Abstract | CrossRef Full Text | Google Scholar

123. Wei W-Q, Denny JC. Extracting research-quality phenotypes from electronic health records to support precision medicine. Genome Med (2015) 7(1):2–8. doi: 10.1186/s13073-015-0166-y

PubMed Abstract | CrossRef Full Text | Google Scholar

124. Wicks P, Vaughan TE, Massagli MP, Heywood J. Accelerated clinical discovery using self-reported patient data collected online and a patient-matching algorithm. Nat Biotechnol (2011) 29(5):411–4. doi: 10.1038/nbt.1837

PubMed Abstract | CrossRef Full Text | Google Scholar

125. Medina-Franco JL, Yoo J, Dueñas-González A. DNA Methyltransferase Inhibitors for Cancer Therapy. In: Epigenetic Technological Applications. Elsevier (2015). p. 265–90. doi: 10.1016/B978-0-12-801080-8.00013-2

CrossRef Full Text | Google Scholar

126. Carrella D, Napolitano F, Rispoli R, Miglietta M, Carissimo A, Cutillo L, et al. Mantra 2.0: an online collaborative resource for drug mode of action and repurposing by network analysis. Bioinf (Oxford England) (2014) 30(12):1787–8. doi: 10.1093/bioinformatics/btu058

CrossRef Full Text | Google Scholar

127. Chang RL, Xie L, Xie L, Bourne PE, Palsson BØ. Drug off-target effects predicted using structural analysis in the context of a metabolic network model. PloS Comput Biol (2010) 6(9):e1000938. doi: 10.1371/journal.pcbi.1000938

PubMed Abstract | CrossRef Full Text | Google Scholar

128. Knapp S, Weinmann H. Small-Molecule Modulators for Epigenetics Targets. ChemMedChem (2013) 8(11):1885–91. doi: 10.1002/cmdc.201300344

PubMed Abstract | CrossRef Full Text | Google Scholar

129. Wang Y, Bryant SH, Cheng T, Wang J, Gindulyte A, Shoemaker BA, et al. PubChem BioAssay: 2017 update. Nucleic Acids Res (2017) 45(D1):D955–63. doi: 10.1093/nar/gkw1118

PubMed Abstract | CrossRef Full Text | Google Scholar

130. Ursu O, Holmes J, Bologa CG, Yang JJ, Mathias SL, Stathias V, et al. DrugCentral 2018: an update. Nucleic Acids Res (2019) 47(D1):D963–70. doi: 10.1093/nar/gky963

PubMed Abstract | CrossRef Full Text | Google Scholar

131. Tanoli Z, Alam Z, Vähä-Koskela M, Ravikumar B, Malyutina A, Jaiswal A, et al. Drug Target Commons 2.0: a community platform for systematic analysis of drug-target interaction profiles. Database: J Biol Database Curation (2018) 2018:1–13. doi: 10.1093/database/bay083

CrossRef Full Text | Google Scholar

132. Wishart DS, Knox C, Guo AC, Shrivastava S, Hassanali M, Stothard P, et al. DrugBank: a comprehensive resource for in silico drug discovery and exploration. Nucleic Acids Res (2006) 34(Database issue):D668–672. doi: 10.1093/nar/gkj067

PubMed Abstract | CrossRef Full Text | Google Scholar

133. Qi Y, Wang D, Wang D, Jin T, Yang L, Wu H, et al. HEDD: the human epigenetic drug database(2016). (Accessed Accessed: 10th October 2020). doi: 10.1093/database/baw159

CrossRef Full Text | Google Scholar

134. Shah SG, Mandloi T, Kunte P, Natu A, Rashid M, Reddy D, et al. HISTome2: a database of histone proteins, modifiers for multiple organisms and epidrugs. Epigenet Chromatin (2020) 13(1):31. doi: 10.1186/s13072-020-00354-8

CrossRef Full Text | Google Scholar

135. Singh Nanda J, Kumar R, Raghava GPS. dbEM: A database of epigenetic modifiers curated from cancerous and normal genomes. Sci Rep (2016) 6(1):19340. doi: 10.1038/srep19340

PubMed Abstract | CrossRef Full Text | Google Scholar

136. von Eichborn J, Murgueitio MS, Dunkel M, Koerner S, Bourne PE, Preissner R. PROMISCUOUS: a database for network-based drug-repositioning. Nucleic Acids Res (2011) 39(Database issue):D1060–6. doi: 10.1093/nar/gkq1037

PubMed Abstract | CrossRef Full Text | Google Scholar

137. Corsello SM, Bittker JA, Liu Z, Gould J, McCarren P, Hirschman JE, et al. The Drug Repurposing Hub: a next-generation drug library and information resource. Nat Med (2017) 23(4):405–8. doi: 10.1038/nm.4306

PubMed Abstract | CrossRef Full Text | Google Scholar

138. Shameer K, Glicksberg BS, Hodos R, Johnson KW, Badgeley MA, Readhead B, et al. Systematic analyses of drugs and disease indications in RepurposeDB reveal pharmacological, biological and epidemiological factors influencing drug repositioning. Briefings Bioinf (2018) 19(4):656–78. doi: 10.1093/bib/bbw136

CrossRef Full Text | Google Scholar

139. Brown AS, Patel CJ. A standard database for drug repositioning. Sci Data (2017) 4(1):170029. doi: 10.1038/sdata.2017.29

PubMed Abstract | CrossRef Full Text | Google Scholar

140. Himmelstein DS, Lizee A, Hessler C, Brueggeman L, Chen SL, Hadley D, et al. Systematic integration of biomedical knowledge prioritizes drugs for repurposing. eLife (2017) 6:e26726. doi: 10.7554/eLife.26726

PubMed Abstract | CrossRef Full Text | Google Scholar

141. Pantziarka P, Bouche G, Meheus L, Sukhatme V, Sukhatme VP. The Repurposing Drugs in Oncology (ReDO) Project (Accessed 10th October 2020).

Google Scholar

142. Tanoli Z, Seemab U, Scherer A, Wennerberg K, Tang J, Vähä-Koskela M. Exploration of databases and methods supporting drug repurposing: a comprehensive survey. Briefings Bioinf (2020) 1–23. doi: 10.1093/bib/bbaa003

CrossRef Full Text | Google Scholar

143. Hay M, Thomas DW, Craighead JL, Economides C, Rosenthal J. Clinical development success rates for investigational drugs. Nat Biotechnol (2014) 32(1):40–51. doi: 10.1038/nbt.2786

PubMed Abstract | CrossRef Full Text | Google Scholar

144. Chen S, Wang Z, Huang Y, O’Barr SA, Wong RA, Yeung S, et al. Ginseng and Anticancer Drug Combination to Improve Cancer Chemotherapy: A Critical Review. Evidence-Based Complement Altern Med (2014) 2014:1–13. doi: 10.1155/2014/168940

CrossRef Full Text | Google Scholar

145. Lötsch J, Schneider G, Reker D, Parnham MJ, Schneider P, Geisslinger G, et al. Common non-epigenetic drugs as epigenetic modulators. Trends Mol Med (2013) 19(12):742–53. doi: 10.1016/j.molmed.2013.08.006

PubMed Abstract | CrossRef Full Text | Google Scholar

146. Christman JK. 5-Azacytidine and 5-aza-2’-deoxycytidine as inhibitors of DNA methylation: mechanistic studies and their implications for cancer therapy. Oncogene (2002) 21(35):5483–95. doi: 10.1038/sj.onc.1205699

PubMed Abstract | CrossRef Full Text | Google Scholar

147. Fenaux P, Mufti GJ, Hellstrom-Lindberg E, Santini V, Finelli C, Giagounidis A, et al. Efficacy of azacitidine compared with that of conventional care regimens in the treatment of higher-risk myelodysplastic syndromes: a randomised, open-label, phase III study. Lancet Oncol (2009) 10(3):223–32. doi: 10.1016/S1470-2045(09)70003-8

PubMed Abstract | CrossRef Full Text | Google Scholar

148. Kantarjian H, Issa J-PJ, Rosenfeld CS, Bennett JM, Albitar M, DiPersio J, et al. Decitabine improves patient outcomes in myelodysplastic syndromes: results of a phase III randomized study. Cancer (2006) 106(8):1794–803. doi: 10.1002/cncr.21792

PubMed Abstract | CrossRef Full Text | Google Scholar

149. Oodi A, Norouzi H, Amirizadeh N, Nikougoftar M, Vafaie Z. Harmine, a Novel DNA Methyltransferase 1 Inhibitor in the Leukemia Cell Line. Indian J Hematol Blood Transfusion: Off J Indian Soc Hematol Blood Transfusion (2017) 33(4):509–15. doi: 10.1007/s12288-016-0770-z

CrossRef Full Text | Google Scholar

150. Lee WJ, Zhu BT. Inhibition of DNA methylation by caffeic acid and chlorogenic acid, two common catechol-containing coffee polyphenols. Carcinogenesis (2006) 27(2):269–77. doi: 10.1093/carcin/bgi206

PubMed Abstract | CrossRef Full Text | Google Scholar

151. Fagan RL, Cryderman DE, Kopelovich L, Wallrath LL, Brenner C. Laccaic Acid A Is a Direct, DNA-competitive Inhibitor of DNA Methyltransferase 1. J Biol Chem (2013) 288(33):23858–67. doi: 10.1074/jbc.M113.480517

PubMed Abstract | CrossRef Full Text | Google Scholar

152. Li Y-C, Wang Y, Li D-D, Zhang Y, Zhao T-C, Li C-F. Procaine is a specific DNA methylation inhibitor with anti-tumor effect for human gastric cancer. J Cell Biochem (2018) 119(2):2440–9. doi: 10.1002/jcb.26407

PubMed Abstract | CrossRef Full Text | Google Scholar

153. Villar-Garea A, Fraga MF, Espada J, Esteller M. Procaine is a DNA-demethylating agent with growth-inhibitory effects in human cancer cells. Cancer Res (2003) 63(16):4984–9.

PubMed Abstract | Google Scholar

154. Lee BH, Yegnasubramanian S, Lin X, Nelson WG. Procainamide Is a Specific Inhibitor of DNA Methyltransferase 1. J Biol Chem (2005) 280(49):40749–56. doi: 10.1074/jbc.M505593200

PubMed Abstract | CrossRef Full Text | Google Scholar

155. Agarwal S, Amin KS, Jagadeesh S, Baishay G, Rao PG, Barua NC, et al. Mahanine restores RASSF1A expression by down-regulating DNMT1 and DNMT3B in prostate cancer cells. Mol Cancer (2013) 12(1):99. doi: 10.1186/1476-4598-12-99

PubMed Abstract | CrossRef Full Text | Google Scholar

156. Deng C, Lu Q, Zhang Z, Rao T, Attwood J, Yung R, et al. Hydralazine may induce autoimmunity by inhibiting extracellular signal-regulated kinase pathway signaling. Arthritis Rheum (2003) 48(3):746–56. doi: 10.1002/art.10833

PubMed Abstract | CrossRef Full Text | Google Scholar

157. Segura-Pacheco B, Perez-Cardenas E, Taja-Chayeb L, Chavez-Blanco A, Revilla-Vazquez A, Benitez-Bribiesca L, et al. Global DNA hypermethylation-associated cancer chemotherapy resistance and its reversion with the demethylating agent hydralazine. J Trans Med (2006) 4(1):32. doi: 10.1186/1479-5876-4-32

CrossRef Full Text | Google Scholar

158. Song Y, Zhang C. Hydralazine inhibits human cervical cancer cell growth in vitro in association with APC demethylation and re-expression. Cancer Chemother Pharmacol (2009) 63(4):605–13. doi: 10.1007/s00280-008-0773-z

PubMed Abstract | CrossRef Full Text | Google Scholar

159. Candelaria M, Herrera A, Labardini J, González-Fierro A, Trejo-Becerril C, Taja-Chayeb L, et al. Hydralazine and magnesium valproate as epigenetic treatment for myelodysplastic syndrome. Preliminary results of a phase-II trial. Ann Hematol (2011) 90(4):379–87. doi: 10.1007/s00277-010-1090-2

PubMed Abstract | CrossRef Full Text | Google Scholar

160. Soto H, Sanchez K, Escobar JY, Constanzo A, Fernandez Z, Melendez C. Cost-Effectiveness Analysis of Hydralazine and Magnesium Valproate LP Associated With Treatment for Adult Patients with Metastatic Recurrent or Persistent Cervical Cancer in Mexico. Value Health: J Int Soc Pharmacoeconom Outcomes Res (2014) 17(7):A639. doi: 10.1016/j.jval.2014.08.2300

CrossRef Full Text | Google Scholar

161. Cervera E, Candelaria M, López-Navarro O, Labardini J, Gonzalez-Fierro A, Taja-Chayeb L, et al. Epigenetic Therapy With Hydralazine and Magnesium Valproate Reverses Imatinib Resistance in Patients With Chronic Myeloid Leukemia. Clin Lymphoma Myeloma Leukemia (2012) 12(3):207–12. doi: 10.1016/j.clml.2012.01.005

CrossRef Full Text | Google Scholar

162. Méndez-Lucio O, Tran J, Medina-Franco JL, Meurice N, Muller M. Toward Drug Repurposing in Epigenetics: Olsalazine as a Hypomethylating Compound Active in a Cellular Context. ChemMedChem (2014) 9(3):560–5. doi: 10.1002/cmdc.201300555

PubMed Abstract | CrossRef Full Text | Google Scholar

163. Lin R-K, Hsu C-H, Wang Y-C. Mithramycin A inhibits DNA methyltransferase and metastasis potential of lung cancer cells. Anti Cancer Drugs (2007) 18(10):1157–64. doi: 10.1097/CAD.0b013e3282a215e9

PubMed Abstract | CrossRef Full Text | Google Scholar

164. Kuck D, Caulfield T, Lyko F, Medina-Franco JL. Nanaomycin A selectively inhibits DNMT3B and reactivates silenced tumor suppressor genes in human cancer cells. Mol Cancer Ther (2010) 9(11):3015–23. doi: 10.1158/1535-7163.MCT-10-0609

PubMed Abstract | CrossRef Full Text | Google Scholar

165. Kalra G, De Sousa A, Shrivastava A. Disulfiram in the management of alcohol dependence: A comprehensive clinical review. Open J Psychiatry (2014) 4(1):720–6. doi: 10.4236/ojpsych.2014.41007

CrossRef Full Text | Google Scholar

166. Arce C, Segura-Pacheco B, Perez-Cardenas E, Taja-Chayeb L, Candelaria M, Dueñnas-Gonzalez A. Hydralazine target: From blood vessels to the epigenome. J Trans Med (2006) 4:10. doi: 10.1186/1479-5876-4-10

CrossRef Full Text | Google Scholar

167. Segura-Pacheco B, Trejo-Becerril C, Perez-Cardenas E, Taja-Chayeb L, Mariscal I, Chavez A, et al. Reactivation of tumor suppressor genes by the cardiovascular drugs hydralazine and procainamide and their potential use in cancer therapy. Clin Cancer Res: Off J Am Assoc Cancer Res (2003) 9(5):1596–603.

Google Scholar

168. Lin X, Asgari K, Putzi MJ, Gage WR, Yu X, Cornblatt BS, et al. Reversal of GSTP1 CpG island hypermethylation and reactivation of pi-class glutathione S-transferase (GSTP1) expression in human prostate cancer cells by treatment with procainamide. Cancer Res (2001) 61(24):8611–6.

PubMed Abstract | Google Scholar

169. Gao Z, Xu Z, Hung M-S, Lin Y-C, Wang T, Gong M, et al. Procaine and procainamide inhibit the Wnt canonical pathway by promoter demethylation of WIF-1 in lung cancer cells. Oncol Rep (2009) 22(6):1479–84. doi: 10.3892/or_00000590

PubMed Abstract | CrossRef Full Text | Google Scholar

170. Tada M, Imazeki F, Fukai K, Sakamoto A, Arai M, Mikata R, et al. Procaine inhibits the proliferation and DNA methylation in human hepatoma cells. Hepatol Int (2007) 1(3):355–64. doi: 10.1007/s12072-007-9014-5

PubMed Abstract | CrossRef Full Text | Google Scholar

171. Jagadeesh S, Sinha S, Pal BC, Bhattacharya S, Banerjee PP. Mahanine reverses an epigenetically silenced tumor suppressor gene RASSF1A in human prostate cancer cells. Biochem Biophys Res Commun (2007) 362(1):212–7. doi: 10.1016/j.bbrc.2007.08.005

PubMed Abstract | CrossRef Full Text | Google Scholar

172. Fang MZ, Chen D, Sun Y, Jin Z, Christman JK, Yang CS. Reversal of hypermethylation and reactivation of p16INK4a, RARbeta, and MGMT genes by genistein and other isoflavones from soy. Clin Cancer Res: Off J Am Assoc Cancer Res (2005) 11(19 Pt 1):7033–41. doi: 10.1158/1078-0432.CCR-05-0406

CrossRef Full Text | Google Scholar

173. Majid S, Dar AA, Ahmad AE, Hirata H, Kawakami K, Shahryari V, et al. BTG3 tumor suppressor gene promoter demethylation, histone modification and cell cycle arrest by genistein in renal cancer. Carcinogenesis (2009) 30(4):662–70. doi: 10.1093/carcin/bgp042

PubMed Abstract | CrossRef Full Text | Google Scholar

174. Hodgson N, Trivedi M, Muratore C, Li S, Deth R. Soluble oligomers of amyloid-β cause changes in redox state, DNA methylation, and gene transcription by inhibiting EAAT3 mediated cysteine uptake. J Alzheimer’s Disease: JAD (2013) 36(1):197–209. doi: 10.3233/JAD-130101

CrossRef Full Text | Google Scholar

175. Trivedi MS, Shah JS, Al-Mughairy S, Hodgson NW, Simms B, Trooskens GA, et al. Food-derived opioid peptides inhibit cysteine uptake with redox and epigenetic consequences. J Nutr Biochem (2014) 25(10):1011–8. doi: 10.1016/j.jnutbio.2014.05.004

PubMed Abstract | CrossRef Full Text | Google Scholar

176. Trivedi MS, Hodgson NW, Walker SJ, Trooskens G, Nair V, Deth RC. Epigenetic effects of casein-derived opioid peptides in SH-SY5Y human neuroblastoma cells. Nutr Metab (2015) 12(1):54. doi: 10.1186/s12986-015-0050-1

CrossRef Full Text | Google Scholar

177. Gopal YNV, Arora TS, Van Dyke MW. Parthenolide specifically depletes histone deacetylase 1 protein and induces cell death through ataxia telangiectasia mutated. Chem Biol (2007) 14(7):813–23. doi: 10.1016/j.chembiol.2007.06.007

PubMed Abstract | CrossRef Full Text | Google Scholar

178. Dawood M, Ooko E, Efferth T. Collateral Sensitivity of Parthenolide via NF-κB and HIF-α Inhibition and Epigenetic Changes in Drug-Resistant Cancer Cell Lines. Front Pharmacol (2019) 10:542. doi: 10.3389/fphar.2019.00542

PubMed Abstract | CrossRef Full Text | Google Scholar

179. Hartman ML, Talar B, Sztiller-Sikorska M, Nejc D, Czyz M. Parthenolide induces MITF-M downregulation and senescence in patient-derived MITF-Mhigh melanoma cell populations. Oncotarget (2016) 7(8):9026–40. doi: 10.18632/oncotarget.7030

PubMed Abstract | CrossRef Full Text | Google Scholar

180. Koprowska K, Czyz M. [Molecular mechanisms of parthenolide’s action: Old drug with a new face]. Postepy Higieny I Medycyny Doswiadczalnej (2010) 64:100–14.

PubMed Abstract | Google Scholar

181. Liu Z, Liu S, Xie Z, Pavlovicz RE, Wu J, Chen P, et al. Modulation of DNA Methylation by a Sesquiterpene Lactone Parthenolide. J Pharmacol Exp Ther (2009) 329(2):505–14. doi: 10.1124/jpet.108.147934

PubMed Abstract | CrossRef Full Text | Google Scholar

182. Izquierdo-Torres E, Hernández-Oliveras A, Meneses-Morales I, Rodríguez G, Fuentes-García G, Zarain-Herzberg Á. Resveratrol up-regulates ATP2A3 gene expression in breast cancer cell lines through epigenetic mechanisms. Int J Biochem Cell Biol (2019) 113:37–47. doi: 10.1016/j.biocel.2019.05.020

PubMed Abstract | CrossRef Full Text | Google Scholar

183. Chatterjee B, Ghosh K, Kanade SR. Resveratrol modulates epigenetic regulators of promoter histone methylation and acetylation that restores BRCA1, p53, p21CIP1 in human breast cancer cell lines. BioFactors (Oxford England) (2019) 45(5):818–29. doi: 10.1002/biof.1544

CrossRef Full Text | Google Scholar

184. Venturelli S, Berger A, Böcker A, Busch C, Weiland T, Noor S, et al. Resveratrol as a pan-HDAC inhibitor alters the acetylation status of histone [corrected] proteins in human-derived hepatoblastoma cells. PloS One (2013) 8(8):e73097. doi: 10.1371/journal.pone.0073097

PubMed Abstract | CrossRef Full Text | Google Scholar

185. Liu X, Li H, Wu M-L, Wu J, Sun Y, Zhang K-L, et al. Resveratrol Reverses Retinoic Acid Resistance of Anaplastic Thyroid Cancer Cells via Demethylating CRABP2 Gene. Front Endocrinol (2019) 10:734. doi: 10.3389/fendo.2019.00734

CrossRef Full Text | Google Scholar

186. Lee WJ, Shim J-Y, Zhu BT. Mechanisms for the inhibition of DNA methyltransferases by tea catechins and bioflavonoids. Mol Pharmacol (2005) 68(4):1018–30. doi: 10.1124/mol.104.008367

PubMed Abstract | CrossRef Full Text | Google Scholar

187. Fang MZ, Wang Y, Ai N, Hou Z, Sun Y, Lu H, et al. Tea polyphenol (-)-epigallocatechin-3-gallate inhibits DNA methyltransferase and reactivates methylation-silenced genes in cancer cell lines. Cancer Res (2003) 63(22):7563–70.

PubMed Abstract | Google Scholar

188. Nandakumar V, Vaid M, Katiyar SK. (-)-Epigallocatechin-3-gallate reactivates silenced tumor suppressor genes, Cip1/p21 and p16INK4a, by reducing DNA methylation and increasing histones acetylation in human skin cancer cells. Carcinogenesis (2011) 32(4):537–44. doi: 10.1093/carcin/bgq285

PubMed Abstract | CrossRef Full Text | Google Scholar

189. Chen K-L, Wang SS-S, Yang Y-Y, Yuan R-Y, Chen R-M, Hu C-J. The epigenetic effects of amyloid-beta(1-40) on global DNA and neprilysin genes in murine cerebral endothelial cells. Biochem Biophys Res Commun (2009) 378(1):57–61. doi: 10.1016/j.bbrc.2008.10.173

PubMed Abstract | CrossRef Full Text | Google Scholar

190. Yuan Z, Chen S, Gao C, Dai Q, Zhang C, Sun Q, et al. Development of a versatile DNMT and HDAC inhibitor C02S modulating multiple cancer hallmarks for breast cancer therapy. Bioorg Chem (2019) 87:200–8. doi: 10.1016/j.bioorg.2019.03.027

PubMed Abstract | CrossRef Full Text | Google Scholar

191. Cicero AFG, Baggioni A. Berberine and Its Role in Chronic Disease. Adv Exp Med Biol (2016) 928:27–45. doi: 10.1007/978-3-319-41334-1_2

PubMed Abstract | CrossRef Full Text | Google Scholar

192. Wang Z, Liu Y, Xue Y, Hu H, Ye J, Li X, et al. Berberine acts as a putative epigenetic modulator by affecting the histone code. Toxicol Vitro (2016) 36:10–7. doi: 10.1016/j.tiv.2016.06.004

CrossRef Full Text | Google Scholar

193. Qing Y, Hu H, Liu Y, Feng T, Meng W, Jiang L, et al. Berberine induces apoptosis in human multiple myeloma cell line U266 through hypomethylation of p53 promoter. Cell Biol Int (2014) 38(5):563–70. doi: 10.1002/cbin.10206

PubMed Abstract | CrossRef Full Text | Google Scholar

194. Kalaiarasi A, Anusha C, Sankar R, Rajasekaran S, John Marshal J, Muthusamy K, et al. Plant Isoquinoline Alkaloid Berberine Exhibits Chromatin Remodeling by Modulation of Histone Deacetylase To Induce Growth Arrest and Apoptosis in the A549 Cell Line. J Agric Food Chem (2016) 64(50):9542–50. doi: 10.1021/acs.jafc.6b04453

PubMed Abstract | CrossRef Full Text | Google Scholar

195. Tipoe GL, Leung T-M, Hung M-W, Fung M-L. Green tea polyphenols as an anti-oxidant and anti-inflammatory agent for cardiovascular protection. Cardiovasc Hematol Disord Drug Targets (2007) 7(2):135–44. doi: 10.2174/187152907780830905

PubMed Abstract | CrossRef Full Text | Google Scholar

196. Fernandes GFS, Silva GDB, Pavan AR, Chiba DE, Chin CM, Dos Santos JL. Epigenetic Regulatory Mechanisms Induced by Resveratrol. Nutrients (2017) 9(11):1–2. doi: 10.3390/nu9111201

CrossRef Full Text | Google Scholar

197. Daud AI, Dawson J, DeConti RC, Bicaku E, Marchion D, Bastien S, et al. Potentiation of a Topoisomerase I Inhibitor, Karenitecin, by the Histone Deacetylase Inhibitor Valproic Acid in Melanoma: Translational and Phase I/II Clinical Trial. Clin Cancer Res (2009) 15(7):2479–87. doi: 10.1158/1078-0432.CCR-08-1931

PubMed Abstract | CrossRef Full Text | Google Scholar

198. Rocca A, Minucci S, Tosti G, Croci D, Contegno F, Ballarini M, et al. A phase I–II study of the histone deacetylase inhibitor valproic acid plus chemoimmunotherapy in patients with advanced melanoma. Br J Cancer (2009) 100(1):28–36. doi: 10.1038/sj.bjc.6604817

PubMed Abstract | CrossRef Full Text | Google Scholar

199. Patel MM, Patel BM. Repurposing of sodium valproate in colon cancer associated with diabetes mellitus: Role of HDAC inhibition. Eur J Pharm Sci: Off J Eur Fed Pharm Sci (2018) 121:188–99. doi: 10.1016/j.ejps.2018.05.026

CrossRef Full Text | Google Scholar

200. Du X, Li Q, Du F, He Z, Wang J. Sodium Valproate Sensitizes Non-Small Lung Cancer A549 Cells to γδ T-Cell-Mediated Killing through Upregulating the Expression of MICA. J Biochem Mol Toxicol (2013) 27(11):492–8. doi: 10.1002/jbt.21513

PubMed Abstract | CrossRef Full Text | Google Scholar

201. Friedmann I, Atmaca A, Chow KU, Jäger E, Weidmann E. Synergistic Effects of Valproic Acid and Mitomycin C in Adenocarcinoma Cell Lines and Fresh Tumor Cells of Patients with Colon Cancer. J Chemother (2006) 18(4):415–20. doi: 10.1179/joc.2006.18.4.415

PubMed Abstract | CrossRef Full Text | Google Scholar

202. Yan H-C, Zhang J. Effects of sodium valproate on the growth of human ovarian cancer cell line HO8910. Asian Pacific J Cancer Prevent: APJCP (2012) 13(12):6429–33. doi: 10.7314/apjcp.2012.13.12.6429

CrossRef Full Text | Google Scholar

203. Kumari K, Keshari S, Sengupta D, Sabat SC, Mishra SK. Transcriptome analysis of genes associated with breast cancer cell motility in response to Artemisinin treatment. BMC Cancer (2017) 17(1):858. doi: 10.1186/s12885-017-3863-7

PubMed Abstract | CrossRef Full Text | Google Scholar

204. Byun MR, Lee DH, Jang YP, Lee HS, Choi JW, Lee SK. Repurposing natural products as novel HDAC inhibitors by comparative analysis of gene expression profiles. Phytomedicine (2019) 59:152900. doi: 10.1016/j.phymed.2019.152900

PubMed Abstract | CrossRef Full Text | Google Scholar

205. Joung KE, Kim D-K, Sheen YY. Antiproliferative effect of trichostatin a and hc-toxin in T47D Human breast cancer cells. Arch Pharmacal Res (2004) 27(6):640–5. doi: 10.1007/BF02980164

CrossRef Full Text | Google Scholar

206. Deubzer HE, Ehemann V, Westermann F, Heinrich R, Mechtersheimer G, Kulozik AE, et al. Histone deacetylase inhibitor Helminthosporium carbonum (HC)-toxin suppresses the malignant phenotype of neuroblastoma cells. Int J Cancer (2008) 122(8):1891–900. doi: 10.1002/ijc.23295

PubMed Abstract | CrossRef Full Text | Google Scholar

207. Park Y, Liu Y, Hong J, Lee C-O, Cho H, Kim D-K, et al. New bromotyrosine derivatives from an association of two sponges, Jaspis wondoensis and Poecillastra wondoensis. J Natural Products (2003) 66(11):1495–8. doi: 10.1021/np030162j

CrossRef Full Text | Google Scholar

208. Kim DH, Shin J, Kwon HJ. Psammaplin A is a natural prodrug that inhibits class I histone deacetylase. Exp Mol Med (2007) 39(1):47–55. doi: 10.1038/emm.2007.6

PubMed Abstract | CrossRef Full Text | Google Scholar

209. Ahn MY, Jung JH, Na YJ, Kim HS. A natural histone deacetylase inhibitor, Psammaplin A, induces cell cycle arrest and apoptosis in human endometrial cancer cells. Gynecol Oncol (2008) 108(1):27–33. doi: 10.1016/j.ygyno.2007.08.098

PubMed Abstract | CrossRef Full Text | Google Scholar

210. Kim D, Lee IS, Jung JH, Lee CO, Choi SU. Psammaplin A, a natural phenolic compound, has inhibitory effect on human topoisomerase II and is cytotoxic to cancer cells. Anticancer Res (1999) 19(5B):4085–90.

PubMed Abstract | Google Scholar

211. Shin H, Lee YS, Lee YC. Sodium butyrate-induced DAPK-mediated apoptosis in human gastric cancer cells. Oncol Rep (2012) 27(4):1111–5. doi: 10.3892/or.2011.1585

PubMed Abstract | CrossRef Full Text | Google Scholar

212. Li L, Sun Y, Liu J, Wu X, Chen L, Ma L, et al. Histone deacetylase inhibitor sodium butyrate suppresses DNA double strand break repair induced by etoposide more effectively in MCF-7 cells than in HEK293 cells. BMC Biochem (2015) 16:3–8. doi: 10.1186/s12858-014-0030-5

PubMed Abstract | CrossRef Full Text | Google Scholar

213. Cang S, Xu X, Ma Y, Liu D, Chiao JW. Hypoacetylation, hypomethylation, and dephosphorylation of H2B histones and excessive histone deacetylase activity in DU-145 prostate cancer cells. J Hematol Oncol (2016) 9:2–3. doi: 10.1186/s13045-016-0233-x

PubMed Abstract | CrossRef Full Text | Google Scholar

214. Vigushin DM, Ali S, Pace PE, Mirsaidi N, Ito K, Adcock I, et al. Trichostatin A is a histone deacetylase inhibitor with potent antitumor activity against breast cancer in vivo. Clin Cancer Res: Off J Am Assoc Cancer Res (2001) 7(4):971–6.

Google Scholar

215. Chambers AE, Banerjee S, Chaplin T, Dunne J, Debernardi S, Joel SP, et al. Histone acetylation-mediated regulation of genes in leukaemic cells. Eur J Cancer (Oxford England: 1990) (2003) 39(8):1165–75. doi: 10.1016/s0959-8049(03)00072-8

CrossRef Full Text | Google Scholar

216. Ma J, Guo X, Zhang S, Liu H, Lu J, Dong Z, et al. Trichostatin A, a histone deacetylase inhibitor, suppresses proliferation and promotes apoptosis of esophageal squamous cell lines. Mol Med Rep (2015) 11(6):4525–31. doi: 10.3892/mmr.2015.3268

PubMed Abstract | CrossRef Full Text | Google Scholar

217. Fortson WS, Kayarthodi S, Fujimura Y, Xu H, Matthews R, Grizzle WE, et al. Histone deacetylase inhibitors, valproic acid and trichostatin-A induce apoptosis and affect acetylation status of p53 in ERG-positive prostate cancer cells. Int J Oncol (2011) 39(1):111–9. doi: 10.3892/ijo.2011.1014

PubMed Abstract | CrossRef Full Text | Google Scholar

218. Zhang H, Zhao X, Liu H, Jin H, Ji Y. Trichostatin A inhibits proliferation of PC3 prostate cancer cells by disrupting the EGFR pathway. Oncol Lett (2019) 18(1):687–93. doi: 10.3892/ol.2019.10384

PubMed Abstract | CrossRef Full Text | Google Scholar

219. Tiffon C. Histone Deacetylase Inhibition Restores Expression of Hypoxia-Inducible Protein NDRG1 in Pancreatic Cancer. Pancreas (2018) 47(2):200–7. doi: 10.1097/MPA.0000000000000982

PubMed Abstract | CrossRef Full Text | Google Scholar

220. Sanaei M, Kavoosi F. Effect of 5-Aza-2’-Deoxycytidine in Comparison to Valproic Acid and Trichostatin A on Histone Deacetylase 1, DNA Methyltransferase 1, and CIP/KIP Family (p21, p27, and p57) Genes Expression, Cell Growth Inhibition, and Apoptosis Induction in Colon Cancer SW480 Cell Line. Adv Biomed Res (2019) 8:9–15. doi: 10.4103/abr.abr_91_19

PubMed Abstract | CrossRef Full Text | Google Scholar

221. Hernández-Oliveras A, Izquierdo-Torres E, Meneses-Morales I, Rodríguez G, Zarain-Herzberg Á, Santiago-García J. Histone deacetylase inhibitors promote ATP2A3 gene expression in hepatocellular carcinoma cells: p300 as a transcriptional regulator. Int J Biochem Cell Biol (2019) 113:8–16. doi: 10.1016/j.biocel.2019.05.014

PubMed Abstract | CrossRef Full Text | Google Scholar

222. San-Miguel JF, Hungria VTM, Yoon S-S, Beksac M, Dimopoulos MA, Elghandour A, et al. Panobinostat plus bortezomib and dexamethasone versus placebo plus bortezomib and dexamethasone in patients with relapsed or relapsed and refractory multiple myeloma: a multicentre, randomised, double-blind phase 3 trial. Lancet Oncol (2014) 15(11):1195–206. doi: 10.1016/S1470-2045(14)70440-1

PubMed Abstract | CrossRef Full Text | Google Scholar

223. Bradley D, Rathkopf D, Dunn R, Stadler WM, Liu G, Smith DC, et al. Vorinostat in advanced prostate cancer patients progressing on prior chemotherapy (National Cancer Institute Trial 6862): trial results and interleukin-6 analysis: a study by the Department of Defense Prostate Cancer Clinical Trial Consortium and University of Chicago Phase 2 Consortium. Cancer (2009) 115(23):5541–9. doi: 10.1002/cncr.24597

PubMed Abstract | CrossRef Full Text | Google Scholar

224. Watanabe T, Kato H, Kobayashi Y, Yamasaki S, Morita-Hoshi Y, Yokoyama H, et al. Potential efficacy of the oral histone deacetylase inhibitor vorinostat in a phase I trial in follicular and mantle cell lymphoma. Cancer Sci (2010) 101(1):196–200. doi: 10.1111/j.1349-7006.2009.01360.x

PubMed Abstract | CrossRef Full Text | Google Scholar

225. Raedler LA. Farydak (Panobinostat): First HDAC Inhibitor Approved for Patients with Relapsed Multiple Myeloma. Am Health Drug Benefits (2016) 9(Spec Feature):84–7.

PubMed Abstract | Google Scholar

226. Fukui Y, Narita K, Dan S, Yamori T, Ito A, Yoshida M, et al. Total synthesis of burkholdacs A and B and 5,6,20-tri-epi-burkholdac A: HDAC inhibition and antiproliferative activity. Eur J Medicinal Chem (2014) 76:301–13. doi: 10.1016/j.ejmech.2014.02.044

CrossRef Full Text | Google Scholar

227. Benelkebir H, Donlevy AM, Packham G, Ganesan A. Total Synthesis and Stereochemical Assignment of Burkholdac B, a Depsipeptide HDAC Inhibitor. Organic Lett (2011) 13(24):6334–7. doi: 10.1021/ol202197q

CrossRef Full Text | Google Scholar

228. Crabb SJ, Howell M, Rogers H, Ishfaq M, Yurek-George A, Carey K, et al. Characterisation of the in vitro activity of the depsipeptide histone deacetylase inhibitor spiruchostatin A. Biochem Pharmacol (2008) 76(4):463–75. doi: 10.1016/j.bcp.2008.06.004

PubMed Abstract | CrossRef Full Text | Google Scholar

229. Narita K, Fukui Y, Sano Y, Yamori T, Ito A, Yoshida M, et al. Total synthesis of bicyclic depsipeptides spiruchostatins C and D and investigation of their histone deacetylase inhibitory and antiproliferative activities. Eur J Medicinal Chem (2012) 60:295–304. doi: 10.1016/j.ejmech.2012.12.023

CrossRef Full Text | Google Scholar

230. Hong J. Apicidin, a histone deacetylase inhibitor, induces differentiation of HL-60 cells. Cancer Lett (2003) 189(2):197–206. doi: 10.1016/S0304-3835(02)00500-1

PubMed Abstract | CrossRef Full Text | Google Scholar

231. Wu L-P, Wang X, Li L, Zhao Y, Lu S, Yu Y, et al. Histone deacetylase inhibitor depsipeptide activates silenced genes through decreasing both CpG and H3K9 methylation on the promoter. Mol Cell Biol (2008) 28(10):3219–35. doi: 10.1128/MCB.01516-07

PubMed Abstract | CrossRef Full Text | Google Scholar

232. Durczak M, Jagodzinski P. Apicidin upregulates PHD2 prolyl hydroxylase gene expression in cervical cancer cells. Anti Cancer Drugs (2010) 21(6):619–24. doi: 10.1097/CAD.0b013e328339848b

PubMed Abstract | CrossRef Full Text | Google Scholar

233. Im JY, Park H, Kang KW, Choi WS, Kim HS. Modulation of cell cycles and apoptosis by apicidin in estrogen receptor (ER)-positive and-negative human breast cancer cells. Chemico Biol Interact (2008) 172(3):235–44. doi: 10.1016/j.cbi.2008.01.007

CrossRef Full Text | Google Scholar

234. Ahn MY, Kang DO, Na YJ, Yoon S, Choi WS, Kang KW, et al. Histone deacetylase inhibitor, apicidin, inhibits human ovarian cancer cell migration via class II histone deacetylase 4 silencing. Cancer Lett (2012) 325(2):189–99. doi: 10.1016/j.canlet.2012.06.017

PubMed Abstract | CrossRef Full Text | Google Scholar

235. Ahn M-Y. HDAC inhibitor apicidin suppresses murine oral squamous cell carcinoma cell growth in vitro and in vivo via inhibiting HDAC8 expression. Oncol Lett (2018) 16(5):6552–60. doi: 10.3892/ol.2018.9468

PubMed Abstract | CrossRef Full Text | Google Scholar

236. Wagner JM, Hackanson B, Lübbert M, Jung M. Histone deacetylase (HDAC) inhibitors in recent clinical trials for cancer therapy. Clin Epigenet (2010) 1(3–4):117–36. doi: 10.1007/s13148-010-0012-4

CrossRef Full Text | Google Scholar

237. Zhao HL, Harding SV, Marinangeli CPF, Kim YS, Jones PJH. Hypocholesterolemic and anti-obesity effects of saponins from Platycodon grandiflorum in hamsters fed atherogenic diets. J Food Sci (2008) 73(8):H195–200. doi: 10.1111/j.1750-3841.2008.00915.x

PubMed Abstract | CrossRef Full Text | Google Scholar

238. Brosch G, Ransom R, Lechner T, Walton JD, Loidl P. Inhibition of maize histone deacetylases by HC toxin, the host-selective toxin of Cochliobolus carbonum. Plant Cell (1995) 7(11):1941–50. doi: 10.1105/tpc.7.11.1941

PubMed Abstract | CrossRef Full Text | Google Scholar

239. Darkin-Rattray SJ, Gurnett AM, Myers RW, Dulski PM, Crumley TM, Allocco JJ, et al. Apicidin: a novel antiprotozoal agent that inhibits parasite histone deacetylase. Proc Natl Acad Sci U S A (1996) 93(23):13143–7. doi: 10.1073/pnas.93.23.13143

PubMed Abstract | CrossRef Full Text | Google Scholar

240. Liu S, Fu X, Schmitz FJ, Kelly-Borges M. Psammaplysin F, a new bromotyrosine derivative from a sponge, Aplysinella sp. J Natural Products (1997) 60(6):614–5. doi: 10.1021/np970070s

CrossRef Full Text | Google Scholar

241. Piña IC, Gautschi JT, Wang G-Y-S, Sanders ML, Schmitz FJ, France D, et al. Psammaplins from the sponge Pseudoceratina purpurea: inhibition of both histone deacetylase and DNA methyltransferase. J Organic Chem (2003) 68(10):3866–73. doi: 10.1021/jo034248t

CrossRef Full Text | Google Scholar

242. Rinehart KL, Holt TG, Fregeau NL, Keifer PA, Wilson GR, Perun TJ, et al. Bioactive compounds from aquatic and terrestrial sources. J Natural Products (1990) 53(4):771–92. doi: 10.1021/np50070a001

CrossRef Full Text | Google Scholar

243. Mateos MV, Cibeira M, Richardson PG, Prosper F, Oriol A, de la Rubia J, et al. Phase II Clinical and Pharmacokinetic Study of Plitidepsin 3-Hour Infusion Every Two Weeks Alone or with Dexamethasone in Relapsed and Refractory Multiple Myeloma. Clin Cancer Res: Official J Am Assoc Cancer Res (2010) 16(12):3260–9. doi: 10.1158/1078-0432.CCR-10-0469

CrossRef Full Text | Google Scholar

244. Spicka I, Ocio EM, Oakervee HE, Greil R, Banh RH, Huang S-Y, et al. Randomized phase III study (ADMYRE) of plitidepsin in combination with dexamethasone vs. dexamethasone alone in patients with relapsed/refractory multiple myeloma. Ann Hematol (2019) 98(9):2139–50. doi: 10.1007/s00277-019-03739-2

PubMed Abstract | CrossRef Full Text | Google Scholar

245. Alonso-Álvarez S, Pardal E, Sánchez-Nieto D, Navarro M, Caballero MD, Mateos MV, et al. Plitidepsin: design, development, and potential place in therapy. Drug Design Dev Ther (2017) 11:253–64. doi: 10.2147/DDDT.S94165

CrossRef Full Text | Google Scholar

246. Balasubramanyam K, Swaminathan V, Ranganathan A, Kundu TK. Small Molecule Modulators of Histone Acetyltransferase p300. J Biol Chem (2003) 278(21):19134–40. doi: 10.1074/jbc.M301580200

PubMed Abstract | CrossRef Full Text | Google Scholar

247. Collins HM, Abdelghany MK, Messmer M, Yue B, Deeves SE, Kindle KB, et al. Differential effects of garcinol and curcumin on histone and p53 modifications in tumour cells. BMC Cancer (2013) 13(1):37. doi: 10.1186/1471-2407-13-37

PubMed Abstract | CrossRef Full Text | Google Scholar

248. Sung B, Pandey MK, Ahn KS, Yi T, Chaturvedi MM, Liu M, et al. Anacardic acid (6-nonadecyl salicylic acid), an inhibitor of histone acetyltransferase, suppresses expression of nuclear factor-kappaB-regulated gene products involved in cell survival, proliferation, invasion, and inflammation through inhibition of the inhibitory subunit of nuclear factor-kappaBalpha kinase, leading to potentiation of apoptosis. Blood (2008) 111(10):4880–91. doi: 10.1182/blood-2007-10-117994

PubMed Abstract | CrossRef Full Text | Google Scholar

249. Balasubramanyam K, Altaf M, Varier RA, Swaminathan V, Ravindran A, Sadhale PP, et al. Polyisoprenylated Benzophenone, Garcinol, a Natural Histone Acetyltransferase Inhibitor, Represses Chromatin Transcription and Alters Global Gene Expression. J Biol Chem (2004) 279(32):33716–26. doi: 10.1074/jbc.M402839200

PubMed Abstract | CrossRef Full Text | Google Scholar

250. Sethi G, Chatterjee S, Rajendran P, Li F, Shanmugam MK, Wong KF, et al. Inhibition of STAT3 dimerization and acetylation by garcinol suppresses the growth of human hepatocellular carcinoma in vitro and in vivo. Mol Cancer (2014) 13:66. doi: 10.1186/1476-4598-13-66

PubMed Abstract | CrossRef Full Text | Google Scholar

251. Wang J, Wu M, Zheng D, Zhang H, Lv Y, Zhang L, et al. Garcinol inhibits esophageal cancer metastasis by suppressing the p300 and TGF-β1 signaling pathways. Acta Pharmacol Sin (2020) 41(1):82–92. doi: 10.1038/s41401-019-0271-3

PubMed Abstract | CrossRef Full Text | Google Scholar

252. Galvez AF, Chen N, Macasieb J, de Lumen BO. Chemopreventive property of a soybean peptide (lunasin) that binds to deacetylated histones and inhibits acetylation. Cancer Res (2001) 61(20):7473–8.

PubMed Abstract | Google Scholar

253. Ravindra KC, Selvi BR, Arif M, Reddy BAA, Thanuja GR, Agrawal S, et al. Inhibition of Lysine Acetyltransferase KAT3B/p300 Activity by a Naturally Occurring Hydroxynaphthoquinone, Plumbagin. J Biol Chem (2009) 284(36):24453–64. doi: 10.1074/jbc.M109.023861

PubMed Abstract | CrossRef Full Text | Google Scholar

254. Casaos J, Huq S, Lott T, Felder R, Choi J, Gorelick N, et al. Ribavirin as a potential therapeutic for atypical teratoid/rhabdoid tumors. Oncotarget (2018) 9(8):8054–67. doi: 10.18632/oncotarget.23883

PubMed Abstract | CrossRef Full Text | Google Scholar

255. Hu H, Qian K, Ho M-C, Zheng YG. Small Molecule Inhibitors of Protein Arginine Methyltransferases. Expert Opin Invest Drugs (2016) 25(3):335–58. doi: 10.1517/13543784.2016.1144747

CrossRef Full Text | Google Scholar

256. Chen J, Xu X, Chen J. Clinically relevant concentration of anti-viral drug ribavirin selectively targets pediatric osteosarcoma and increases chemosensitivity. Biochem Biophys Res Commun (2018) 506(3):604–10. doi: 10.1016/j.bbrc.2018.10.124

PubMed Abstract | CrossRef Full Text | Google Scholar

257. De la Cruz-Hernandez E, Medina-Franco JL, Trujillo J, Chavez-Blanco A, Dominguez-Gomez G, Perez-Cardenas E, et al. Ribavirin as a tri-targeted antitumor repositioned drug. Oncol Rep (2015) 33(5):2384–92. doi: 10.3892/or.2015.3816

PubMed Abstract | CrossRef Full Text | Google Scholar

258. Catalano R, Rocca R, Juli G, Costa G, Maruca A, Artese A, et al. A drug repurposing screening reveals a novel epigenetic activity of hydroxychloroquine. Eur J Medicinal Chem (2019) 183:111715. doi: 10.1016/j.ejmech.2019.111715

CrossRef Full Text | Google Scholar

259. Han H, Yang X, Pandiyan K, Liang G. Synergistic Re-Activation of Epigenetically Silenced Genes by Combinatorial Inhibition of DNMTs and LSD1 in Cancer Cells. PloS One (2013) 8(9):e75136. doi: 10.1371/journal.pone.0075136

PubMed Abstract | CrossRef Full Text | Google Scholar

260. Sakane C, Okitsu T, Wada A, Sagami H, Shidoji Y. Inhibition of lysine-specific demethylase 1 by the acyclic diterpenoid geranylgeranoic acid and its derivatives. Biochem Biophys Res Commun (2014) 444(1):24–9. doi: 10.1016/j.bbrc.2013.12.144

PubMed Abstract | CrossRef Full Text | Google Scholar

261. Wang M, Liu X, Guo J, Weng X, Jiang G, Wang Z, et al. Inhibition of LSD1 by Pargyline inhibited process of EMT and delayed progression of prostate cancer in vivo. Biochem Biophys Res Commun (2015) 467(2):310–5. doi: 10.1016/j.bbrc.2015.09.164

PubMed Abstract | CrossRef Full Text | Google Scholar

262. Lee MG, Wynder C, Schmidt DM, McCafferty DG, Shiekhattar R. Histone H3 Lysine 4 Demethylation Is a Target of Nonselective Antidepressive Medications. Chem Biol (2006) 13(6):563–7. doi: 10.1016/j.chembiol.2006.05.004

PubMed Abstract | CrossRef Full Text | Google Scholar

263. Singh MM, Manton CA, Bhat KP, Tsai W-W, Aldape K, Barton MC, et al. Inhibition of LSD1 sensitizes glioblastoma cells to histone deacetylase inhibitors. Neuro-Oncology (2011) 13(8):894–903. doi: 10.1093/neuonc/nor049

PubMed Abstract | CrossRef Full Text | Google Scholar

264. Speranzini V, Rotili D, Ciossani G, Pilotto S, Marrocco B, Forgione M, et al. Polymyxins and quinazolines are LSD1/KDM1A inhibitors with unusual structural features. Sci Adv (2016) 2(9):1–3. doi: 10.1126/sciadv.1601017

CrossRef Full Text | Google Scholar

265. Wakchaure P, Velayutham R, Roy KK. Structure investigation, enrichment analysis and structure-based repurposing of FDA-approved drugs as inhibitors of BET-BRD4. J Biomol Struct Dynamics (2019) 37(12):3048–57. doi: 10.1080/07391102.2018.1507838

CrossRef Full Text | Google Scholar

266. Jiang H, Xing J, Wang C, Zhang H, Yue L, Wan X, et al. Discovery of novel BET inhibitors by drug repurposing of nitroxoline and its analogues. Organic Biomol Chem (2017) 15(44):9352–61. doi: 10.1039/C7OB02369C

CrossRef Full Text | Google Scholar

267. Wan X, Liu H, Sun Y, Zhang J, Chen X, Chen N. Lunasin: A promising polypeptide for the prevention and treatment of cancer. Oncol Lett (2017) 13(6):3997–4001. doi: 10.3892/ol.2017.6017

PubMed Abstract | CrossRef Full Text | Google Scholar

268. Bortolato M, Chen K, Shih JC. Monoamine oxidase inactivation: from pathophysiology to therapeutics. Adv Drug Delivery Rev (2008) 60(13–14):1527–33. doi: 10.1016/j.addr.2008.06.002

CrossRef Full Text | Google Scholar

269. Jose A, Shenoy GG, Sunil Rodrigues G, Kumar NAN, Munisamy M, Thomas L, et al. Histone Demethylase KDM5B as a Therapeutic Target for Cancer Therapy. Cancers (2020) 12(8):4–7. doi: 10.3390/cancers12082121

CrossRef Full Text | Google Scholar

270. Hanahan D, Weinberg RA. Hallmarks of Cancer: The Next Generation. Cell (2011) 144(5):646–74. doi: 10.1016/j.cell.2011.02.013

PubMed Abstract | CrossRef Full Text | Google Scholar

271. Esteller M, Garcia-Foncillas J, Andion E, Goodman SN, Hidalgo OF, Vanaclocha V, et al. Inactivation of the DNA-repair gene MGMT and the clinical response of gliomas to alkylating agents. New Engl J Med (2000) 343(19):1350–4. doi: 10.1056/NEJM200011093431901

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: epidrugs, drug repurposing, cancer therapy, cancer, epigenetic inhibitors, epigenetics

Citation: Montalvo-Casimiro M, González-Barrios R, Meraz-Rodriguez MA, Juárez-González VT, Arriaga-Canon C and Herrera LA (2020) Epidrug Repurposing: Discovering New Faces of Old Acquaintances in Cancer Therapy. Front. Oncol. 10:605386. doi: 10.3389/fonc.2020.605386

Received: 12 September 2020; Accepted: 15 October 2020;
Published: 18 November 2020.

Edited by:

Teresita Padilla-Benavides, Wesleyan University, United States

Reviewed by:

Shameer Khader, AstraZeneca, United States
Francisco Cuellar-Perez, University of Texas Southwestern Medical Center, United States

Copyright © 2020 Montalvo-Casimiro, González-Barrios, Meraz-Rodriguez, Juárez-González, Arriaga-Canon and Herrera. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Luis A. Herrera, metil@hotmail.com

These authors have contributed equally to this work

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.