Innate and acquired resistance towards the conventional therapeutic regimen imposes a significant challenge for the successful management of cancer for decades. In patients with advanced carcinomas, acquisition of drug resistance often leads to tumor recurrence and poor prognosis after the first therapeutic cycle. In this context, cancer stem cells (CSCs) are considered as the prime drivers of therapy resistance in cancer due to their ‘non-targetable’ nature. Drug resistance in cancer is immensely influenced by different properties of CSCs such as epithelial-to-mesenchymal transition (EMT), a profound expression of drug efflux pump genes, detoxification genes, quiescence, and evasion of apoptosis, has been highlighted in this review article. The crucial epigenetic alterations that are intricately associated with regulating different mechanisms of drug resistance, have been discussed thoroughly. Additionally, special attention is drawn towards the epigenetic mechanisms behind the interaction between the cancer cells and their microenvironment which assists in tumor progression and therapy resistance. Finally, we have provided a cumulative overview of the alternative treatment strategies and epigenome-modifying therapies that show the potential of sensitizing the resistant cells towards the conventional treatment strategies. Thus, this review summarizes the epigenetic and molecular background behind therapy resistance, the prime hindrance of present day anti-cancer therapies, and provides an account of the novel complementary epi-drug-based therapeutic strategies to combat drug resistance.

In 1971, ‘National Cancer Act’ of the U.S. Congress declared a ‘War on cancer.’ Since then, billions of dollars have been invested in cancer research. Yet, despite decade-long studies and advancement in anti-cancer therapeutic regimens, cancer still poses as one of the major causes of global morbidity and mortality, with the diagnosis of millions of new cases every year. In addition, the development of drug resistance, consequent ineffectiveness of therapy, and successive tumor relapse lead to poor predictive outcomes in the majority of the cases and serve as the major limiting factors of the present day anti-cancer therapies. Studies have shown that despite substantial development in the early detection of breast cancer, more than one-third of patients do not respond to the primary chemotherapeutic regimens and develop therapy resistance [1,2].

Moreover, the fact that chemotherapy is used as the first-line treatment strategy against the majority of the advanced carcinomas only worsens the scenario [3]. Despite the effective reduction in tumor size at the primary site, several patients display the development of distant metastasis [4], which is frequently associated with the development of chemoresistance. It is reported that the selection pressure induced by chemotherapy favors inherently drug-resistant cells (for instance, cancer stem cells (CSCs)) to self-renew, proliferate, and even undergo epithelial-to-mesenchymal transition (EMT) that promotes their metastatic dissemination [5–7]. The tumor cells can also employ intricate signaling cascades to acquire chemoresistant properties during therapy [8,9]. These inherent and acquired chemoresistant cells serve as a repertoire for future tumor recurrence that can give rise to even more aggressive forms of the disease.

The concept of drug resistance was first reported in bacteria, based on antibiotic resistance. Studies on cancer and several other diseases showed similar, evolutionarily conserved mechanisms that impart resistance towards treatment stratagems [10]. For example, therapy resistance in cancer was first noted in the 1940s [11], and over the years, it has become one of the significant concerns of present day cancer management. In addition, a thorough understanding of critical molecular and epigenetic pathways behind this drug-resistant property of cancer cells could unveil a magnitude of possibilities that can be employed to sensitize these resistant cells towards conventional treatment regimens successfully. Finally, a comprehensive overview of the current understanding of the epigenetic mechanisms that render therapy resistance in cancer cells is also elucidated through this review.

The recent era of cancer research has shifted a significant focus towards CSCs as they are being reported as one of the major culprits behind maintaining all the hallmarks of cancer. CSCs, a distinct subset of cells within the heterogeneous tumor mass, are adorned with the ability to self-renew and differentiate into multiple lineages [12]. Over the past few years, many studies have established that CSCs are primary therapy-resistant cells within a tumor, which show numerous resistance mechanisms towards radiotherapy and conventional chemotherapeutic regimen [13,14]. The CSCs are usually resistant to chemotherapy/radiotherapy, which drives as a significant cause of cancer recurrence. They survive after the therapeutic intervention, serving as the seed to promote future tumor relapse. Neoadjuvant chemotherapy enriches the CSC population, suggesting a higher degree of resistance to therapy than the rest of the bulk tumor [15]. Interestingly, CSCs are reported to display different mechanisms of therapy resistance, including induction of EMT, high expression of ATP-binding cassette transporters (ABC transporters) and detoxification genes, quiescence, and evasion of apoptosis [16,17].

EMT

The EMT is the preliminary step in cancer metastasis whose induction is governed by epigenetic modifiers, like in the case of one of the transcription factors, Snail, which recruits multiple chromatin modifiers such as PRC2, histone deacetylase (HDAC)1/2, G9a, LSD1, to the promoter of E-cadherin, thereby promoting Snail-mediated E-cadherin repression, a hallmark of EMT [18].

It has been reported that the induction of EMT promotes the acquisition of stem cell-like features in cancer cells. For instance, Twist, Slug, and Snail, the critical regulators of EMT, induce mesenchymal properties in breast cancer cells, triggering stem cell-like properties and mammosphere-forming ability [19]. It has been suggested that EMT is associated with the induction of drug resistance [7,20]. Previous studies have indicated that the distinct phenotypic differences between CSCs and non-stem cancer cells (NSCCs) are majorly due to the induction of EMT and mesenchymal properties in the CSCs [6]. However, the conventional therapeutic regimens fail to eliminate the cancer cells with CSC-like and mesenchymal properties, promoting CSC-mediated tumor recurrence [7].

Interestingly, EMT and stemness share common signaling pathways like Wnt, Hedgehog, and Notch [20], which suggest their standard mode of action in the induction of drug resistance [21,22]. For example, TGF-β-induced EMT promotes drug resistance [23], while inhibition of TGF-β sensitizes the cells towards chemotherapy [24,25]. Notch 1 signaling-induced EMT activation triggers gefitinib resistance in lung cancer cells [26,27]. The transcription factor ZEB1, which contributes to the regulation of stemness, also plays a significant role in inducing chemoresistance [28]. In pancreatic ductal adenocarcinoma cells, loss of Twist and Snail sensitizes them towards gemcitabine, improving the prognostic outcome [29]. Finally, mechanistically stemness, EMT, and drug resistance are linked, and understanding this network is crucial for developing more potent therapeutic strategies.

Increased expression of ABC transporter and detoxification genes

ABC transporters like ABCB1 (MDR1), ABCC1 (MRP1), ABCG2 are transmembrane proteins that pump toxins out of the cells using energy from ATP hydrolysis. CSCs of various solid tumors display significantly high expression of these drug efflux pumps, rendering them intrinsically resistant to most conventional therapeutic interventions [30,31]. ABCG2 has been reported to efflux the chemotherapeutic drugs doxorubicin and methotrexate [32] and its down-regulation enhances the chemosensitivity of breast CSCs [33]. Importantly, in glioma tumor stem-like cells, PI3K/Akt signaling promotes localization of ABCG2 to the plasma membrane [34]. Moreover, CSC marker aldehyde dehydrogenase 1 (ALDH1), which catalyzes the oxidation of aldehydes, protects them from reactive oxygen species (ROS)-mediated damages [35]. It is also reported that lung carcinoma stem cells showing elevated expression of ALDH1 confer resistance towards gefitinib, an epidermal growth factor receptor tyrosine kinase inhibitor (EGFR-TKI) and other chemotherapeutic drugs [36].

Dormancy or quiescence

Chemotherapy and radiation therapy effectively damages the proliferating tumor cells. However, they have little effect on the cells in dormant/quiescent state [37]. The quiescent state is a state of reversible cell cycle arrest, where the cells remain in the G0 phase and require a mitogen stimulus to enter into the division phase [38]. In the core tumor, CSCs often remain in the state of dormancy, which helps them avoid being targeted by drugs and may give them time for efficiently repairing the damaged DNA. Interestingly, in a study in the human bladder cancer xenograft model, between the gap of chemotherapy cycles, quiescent-label retaining CSCs got recruited into cell division in response to the drug-induced damage. This shows similarity to the recruitment of normal stem cells during the wound-healing process and indicates the role of CSCs in the development of therapy resistance [39].

Evasion of apoptosis

The ability of CSCs to evade apoptosis, a major hallmark of cancer, is well established. Different signaling pathways such as Hedgehog and Notch signaling are activated in CSCs that reportedly promote anti-apoptosis, linked to docetaxel resistance in prostate cancer [40]. Additionally, CSCs can survive DNA damage-mediated apoptosis in various ways. For example, the guardian of genome p53 gets activated in response to DNA damage, and depending on the extent of damage, it positively regulates apoptosis. But in CSCs, p53 is either down-regulated or mutated, which is positively associated with survival of CSCs through dysregulation of apoptosis [41]. Besides, CSCs have a well-developed free radical scavenging system similar to normal stem cells to minimize the level of intracellular ROS, which are considered mediators of ionizing radiation-driven cell death, leading to resistance towards radiotherapy [42].

Several studies have established the significance of epigenetic changes in cancer drug-tolerant persister (DTP) cells which survive within the heterogeneous population and evolve to be more tolerant to increased drug pressure [43–45]. Furthermore, it has been indicated that modulation of epigenetic landscape contributes to the survival and maintenance of CSCs [46]. In addition, cancer cells reorient their epigenomic landscape through DNA methylation and modifications of histone and non-histone proteins to develop different drug resistance mechanisms to avert anti-cancer therapies. Other epigenetic regulators such as DNA methyltransferases (DNMTs), chromatin readers, writers and erasers, various histone modifiers and non-coding RNAs fine-tune the regulation of genes involved in multiple modes of therapy resistance.

Elevated expression of drug efflux pumps, the role of epigenetics

High efflux of therapeutic agents leading to reduced intracellular drug accumulation is established to be the critical reason behind therapy resistance. The ability to efflux drugs at a significantly high rate, to render resistance may be either intrinsic to the tumor or acquired following an initial cycle of therapy. The ABC transporter superfamily, which includes ABCG2, MRP1, and MDR1, are the primary transmembrane transporters responsible for drug efflux [47]. These drug efflux pumps can pump out various therapeutic drugs such as anthracyclines, alkylating agents, vinca alkaloids etc.

CSCs are reported to display the highest expression of these drug efflux pumps within the tumor mass, which is attributable to their resistant phenotype [48]. A recent study has reported that Oct4 and Sox2, the two major stemness factors, indirectly up-regulate ABCG2 expression in CSCs by repressing SMAR1 via recruitment of HDAC1 which leads to deacetylation of SMAR1 promoter (Figure 1). In the absence of Oct4 and Sox2, SMAR1 represses ABCG2 by directly recruiting HDAC2 on its promoter [49]. Also, the high expression of ABC transporters in CSCs is associated with promoter hypomethylation [50]. Thus, promoting DNMT activity in CSCs might be a potential tool to chemosensitize them. For example, a study by Wang et al. has demonstrated that increasing DNMT activity by afatinib results in hypermethylation of the ABCG2 promoter and lowers ABCG2 expression in drug-resistant breast cancer cells [51]. Another study by Martin et al. reported that inducing DNMT activity in brain CSCs increases ABCG2 promoter methylation and represses its expression [52].

The epigenetic regulations behind the different mechanisms of drug resistance

Figure 1
The epigenetic regulations behind the different mechanisms of drug resistance

Schematic diagram representing the different epigenetic alterations that render increased drug efflux, elevated drug detoxification, augmented DNA repair, and apoptosis resistance properties to the cancer cells, making them resistant towards conventional therapeutic regimens. The interaction between tumor microenvironment and cancer cells, which shapes tumor survival and progression, is also represented schematically.

Figure 1
The epigenetic regulations behind the different mechanisms of drug resistance

Schematic diagram representing the different epigenetic alterations that render increased drug efflux, elevated drug detoxification, augmented DNA repair, and apoptosis resistance properties to the cancer cells, making them resistant towards conventional therapeutic regimens. The interaction between tumor microenvironment and cancer cells, which shapes tumor survival and progression, is also represented schematically.

Close modal

Moreover, the increased expression of ABCG2 and the reduced folate carrier SLC19A1 is reported to be due to promoter hypomethylation, which positively correlates with induction of drug resistance [53,54]. Furthermore, the expression of Phase-III transporters such as ABCC6 and SLC22A3 is also modulated by alterations in DNA methylations [55]. Interestingly, in triple-negative breast cancer, the chromatin reader and tumor suppressor protein ZMYND8 promote the formation of a transcription repressor complex with KDM5C and EZH2 that increases the H3K27me3 mark on promoters of ABCB1, ABCC1, and ABCC2 drug efflux pumps, thereby repressing their expression [56]. These pieces of evidence indicate the ability of CSCs to promote drug resistance by epigenetic modifications of DNA and chromatin landscape for up-regulating drug efflux pump expression.

Epigenetic regulation in drug detoxification

Drug-metabolizing enzymes (DMEs), including phase-I and phase-II DMEs, are responsible for biotransformation and detoxification. In phase-I reaction, the drugs are hydrolyzed and detoxified by DMEs like cytochrome P450 enzymes, aldehyde dehydrogenases etc. [57]. While phase-II DMEs catalyze the conjugation of the hydrophilic compound with the products of phase-I reaction making the latter water-soluble compound for easy excretion. Phase-II DMEs consist of transferases such as glutathione S-transferase (GST) [57]. Epigenetic interplay differentially regulates the DME genes in cancer, mediating their drug resistance potential. In addition, CSCs display a significant increase in drug detoxification machineries, promoting their ‘non-targetable’ nature. CSCs are associated with high expression of detoxifying enzymes like ALDH1, which is used to identify CSCs [58]. A report by Honoki et al. demonstrated that the CSCs residing in tumorospheres possess high ALDH1 activity and show the robust capability of chemoresistance and drug detoxification [59] (Figure 1). Another recent study by Peitzsch et al. demonstrated that chemotherapy and irradiation lead to an enrichment of H3K36me3 transcription activation mark on ALDH1 promoter, which up-regulates ALDH1 expression and promotes resistant CSC-like phenotype in the treated cells [60].

P450 is an abundant DME family, subdivided into two subfamilies CYP1-4 and CYP7-51 [57]. Genome-wide integrative analysis established that the DME genes such as CYP2D6, SULT1A1, GSTM5, CYP2C19, CYP1A2, GSTT1, and GSTA4 are differentially modulated by DNA methylation in cancer which leads to inter- and intra-individual differences in drug metabolism, ultimately correlating to poor prognostic outcome [61]. Interestingly, in prostate cancer, CYP24A1, another cytochrome P450 family enzyme, is repressed by promoter methylation and histone repressor mark H3K9me2 [62]. The methylation landscape of the promoters of DMEs predicts prognostic outcome and treatment efficacy of hormone-responsive breast cancers. For instance, the promoter methylation status of estradiol and tamoxifen-metabolizing enzyme CYP1B1 from the P450 family has been reported to predict survival and prognosis in tamoxifen-treated and untreated patients [63]. Another example in this context is the hypermethylation of a phase-II DME, N-acetyl transferase 1 (NAT1) gene, which regulates tamoxifen resistance in breast cancer [64].

The epigenetic mechanisms behind heightened DNA repair

The DNA repair pathway responds to DNA damage and maintains genomic integrity. The DNA repair machinery comprises an intricate system of sensors, transducers, and effectors that synchronize the repair of damaged DNA, thus ensuring cellular survival [65]. The compaction of chromatin regulates the accessibility of DNA repair and other cellular machineries [66,67]. So, by modulating the chromatin compaction, cancer cells modulate the impact of damaging agents over the exposed DNA and subsequently regulate the DNA repair mechanism [68]. It is indicated that the level of radiation-induced double-strand break (DSB) was 5–50 folds higher in the euchromatin region as compared with the heterochromatin region, and also the euchromatin is more prone towards the regional mutation rate variation [69,70]. Accumulating pieces of evidence of the past few years have indicated cancer cells display altered DNA repair pathways that play a crucial role in their resistance towards genotoxic drugs [71]. A study by Kandoth et al. using next-generation sequencing-based large-scale mutation mapping in cancer has unravelled that cancer cells show an elevated rate of gene mutations that play a role in DNA and histone alterations and chromatin remodeling [72], thus leading to possible epigenomic alteration in cancer cells. This altered epigenetic landscape in cancer cells leads to global changes in histone modification patterns, DNA methylation, and nucleosome positions, affecting their chromatin architecture [73]. These, in turn, alter the DNA damage response and repair in cancer cells, which can either allow accumulation of heightened DNA damage, causing instability in the genome (a hallmark of cancer), or can promote enhanced DNA repair, rendering resistance to genotoxic drugs.

The DNA repair protein Ataxia Telangiectasia Mutation (ATM) functions to sense DNA damage and simultaneously activate the repair mechanism associated with it, including homologous recombination and non-homologous end-joining methods [74]. In addition, ATM is a critical player in DNA DSB repair induced by chemotherapeutic treatment, ionizing radiation etc. In addition, ATM gene aberration and epigenetic alteration of ATM expression in cancer are routinely associated with chemotherapy resistance and poor prognostic outcome [75]. Hypermethylation in the promoter of hMLH1, a protein involved in mismatch repair, is commonly seen in colorectal cancer [76] (Figure 1). 5-Aza-2′-deoxycytidine (decitabine)-mediated demethylation of hMLH1 gene promoter leads to sensitization of colorectal cancer cells towards chemotherapeutic drug, 5-fluorouracil [77].

The CSCs show an even more improved rate of DNA repair, which allows them to be the most drug-resistant cells within the heterogeneous tumor mass [78]. The high expressions of O(6)-methylguanine-DNMT, NBS1, Chk1, and Chk2 in CSCs in comparison to NSCCs renders them more resistant towards DNA-damaging therapies, and the altered expression of these genes in CSCs can be attributed to their differential epigenetic landscape [37,79].

The epigenetic allies in inhibiting apoptosis

Tumor development, progression, and response to therapeutic interventions are tightly regulated by programmed cell death or apoptosis [80]. Evading apoptosis is one of the main hallmarks of cancer, ensuring cancer cell survival upon exposure to cytotoxic and genotoxic stresses [81]. Since the mechanism of action of most of the current treatment regimens like chemotherapy is to activate apoptotic pathways, resistance towards apoptosis shapes therapeutic resistance to a massive extent [82]. The acquisition of apoptosis resistance depends on various factors, including equilibrium of pro- and anti-apoptotic signals and expression of the apoptosis-related genes [83]. The different pro- and anti-apoptotic signaling factors are regulated epigenetically at transcription and post-translation levels in the cancer cells. For instance, enhanced expression or activity of anti-apoptotic factors like Bcl-xl, Bcl-2, IAPs etc. leads to the development of apoptosis resistance. Similarly, repression of pro-apoptotic factors like Bax, Bid, Puma, Noxa, Bim etc. promotes resistance towards apoptosis. Moreover, Akt and their transcriptional regulators, NF-κB and STAT, remain highly overexpressed in various advanced cancers, indicating their possible involvement in therapy resistance [84].

Different histone variants and their modifications are associated with carcinogenesis and therapy resistance. For example, the histone variant H2AZ is reported to be overexpressed in various solid cancers, contributing to tumor progression [85]. Knockdown of H2AZ leads to repression of Bcl-2 and elevation of pro-apoptotic marks Caspase 3, Caspase 9, and Bak, thus stimulating apoptotic cell death [86]. Aberrant histone alterations are often associated with cancer progression and therapy resistance. Previous studies have indicated the involvement of dephosphorylation of histone H1, phosphorylation of H2A, H2B, H3, and H4, and deubiquitylation of H2A in the cascade of apoptosis [87]. Tumor cells resist drug toxicity-mediated apoptosis by up-regulating the expressions of various DNA repair genes (like MGMT, BRCA 1/2 etc.) through activation of H3K4Me3 [83]. A study in human lung cancer cell line NCI-H460 revealed that repression of p16INK4a promoter by cumulative action of ZBP-89 and HDAC3 regulates senescence and apoptosis [88] (Figure 1). In breast carcinoma, CDKN1C (the gene encoding tumor suppressor p57KIP2) is repressed by EZH2-mediated histone H3K27 trimethylation (H3K27me3), leading to the acquisition of a more resistant phenotype [89] (Figure 1). In bladder carcinoma cells, the repression of the p21WAF1 promoter is mediated by histone deacetylation through the activity of HDACs [90]. This leads to the development of apoptotic resistance, which can be effectively reversed by targeting the cells with HDAC inhibitors (HDACis) [90]. In addition, specific histone modifications can lead to the recruitment of protein complexes governing the balance between cell death and survival. For example, phosphorylation of H2A.X-Y142 represses MDC1-induced recruitment of DNA repair proteins to DNA damage site (yH2AX) [91]. This, in turn, promotes the recruitment of pro-apoptotic complexes to the site, pushing the cell towards apoptotic fate [91,92]. During the acquisition of therapy resistance, this pathway alters in the cancer cells, thus resisting cell death even after being treated with DNA damaging agents like chemotherapy.

DNA methylation status also plays a pivotal role in modulating apoptotic cell fate in various cancers. The CpG islands on the promoters of tumor suppressor and pro-apoptotic genes remain hypermethylated by the activity of DNMT in cancer cells [93] (Figure 1). Hypermethylation of these genes inhibits the programmed cell death pathway, rendering the tumor cells a resistant phenotype. For instance, the promoter of pro-apoptotic markers Bcl-2 and Bik reportedly remain hypermethylated in prostate cancer [94,95]. Similarly, in multiple myeloma, the promoters of Bad, Bax, Bak, and Puma display hypermethylation [96]. The promoter of Bim reportedly shows hypermethylation and repression in chronic myeloid leukemia [97]. Promoter methylation and repression of Caspase 8 and 10, two significant contributors of the apoptotic pathway, render resistant phenotype in bladder cancer, hepatocellular carcinoma, glioblastoma, small-cell lung cancer, retinoblastoma, and neuroblastoma [98–102]. In gastric cancer, the expression of Bcl-2l10, BNIP3, and HRK remains inhibited by promoter hypermethylation [103–105]. p53, the master tumor suppressor, plays a pivotal role in initiating the apoptotic cascade. In patients with acute lymphatic leukemia, the p53 gene remains hypermethylated and inhibited, leading to the ineffectiveness of treatment regimens [106]. Also, repression of hypermethylated in cancer 1 (HIC1) in cancer cells leads to inactivation of p53, thus inhibiting DNA damage-mediated apoptosis pathway [107]. MGMT, MLH1, BRCA1, APC, and APAF1 expressions are also reported to be diminished in cancer cells by hypermethylation of their promoters [108,109]. Repression of FAS expression by hypermethylation in its promoter is reported to be involved in cutaneous T-cell lymphoma carcinogenesis and apoptosis resistance [110].

Apart from DNA methylation and histone post-translational modifications, miRNAs also play a critical role in the modulation of apoptosis resistance. For instance, p53 induces the expression of the miR-34 family, which reportedly can inhibit the expression of MYCN, CDK4/6, Notch, Cyclin E2, and Bcl-2 [111]. The expression of miR-34 is repressed in various solid and liquid cancers, including breast/colon/gastric/kidney/pancreatic carcinoma, Burkitt’s lymphoma, and chronic lymphocytic leukemia indicating the ingenious strategy of cancer cells to resist apoptosis [112,113]. Similarly, the expressions of miR-15, miR-16, miR-193a-3p, miR-29b, miR-133b, miR-512-5p remain repressed in different types of cancers due to their involvement in promoting apoptosis [114–119]. In contrast, miR-221 and miR-222 remains overexpressed in various solid cancers, leading to repression of TIMP3, PTEN, FOXO3A, PUMA, Caspase 3, etc. [120–123]. miR-BART5 is reported to be involved in the apoptosis-resistance of gastric cancer cells due to its ability to target PUMA [124]. miR-135a, which reportedly represses JAK2 and Bcl-xl, displays down-regulation in various cancers like ovarian cancer, AML etc. [125,126].

All these reports indicate the intricate involvement of epigenetic machinery in modulating apoptosis in cancer cells, indicating that epigenetic reprogramming plays a pivotal role in rendering apoptosis resistance to the cancer cells.

The tumor microenvironment (TME) is an intricate and dynamic landscape that consists of proliferating tumor cells, tumor-associated immune cells, endothelial cells, and extracellular matrix. The TME plays a crucial part in tumor progression, invasiveness, metastatic insemination and impacts the clinical outcome of the tumor. The interaction between the cell and extracellular matrix of the TME is critical for cancer progression, and it plays a vital role in the development of therapy resistance. The different components of TME are reported to be regulated and reprogrammed by epigenetic alterations, including DNA and histone modifications.

Role of epigenetics in reprogramming tumor-associated immune landscape

It is well established that epigenetic machinery plays a critical role in reprogramming and recruiting the intrinsic and adaptive immune cells into the TME, supporting tumor progression. It is reported that HDAC5-mediated inhibition of Socs3 and up-regulation of CCL2 leads to recruitment of tumor-associated macrophages (TAMs) [127]. In the TAMs, ERK and JNK facilitated histone deacetylation of CIITA promoter leads to decoy receptor (DcR3)-induced repression of MHC-II expression [128]. This impairs antigen presentation in TAMs, resulting in immunosuppression in the TME [128]. In addition, the myeloid-derived suppressor cells (MDSCs) are a subset of immune cells that expands during cancer and plays a critical role in tumor development, progression, and therapy resistance [129]. It is reported that, in hepatocellular carcinoma, the interaction between EZH2 and p65-NF-kB and their binding on IL-6 promoter augments its expression and induces MDSC recruitment in the TME, which is correlated with a poor therapeutic response [130] (Figure 1).

Furthermore, the CD8+ cytotoxic T cells recognize and kill target antigens through MHC-I, which is epigenetically repressed in the TME as a strategy for immune evasion. Finally, a study has reported that PRC2 suppresses MHC-I through bivalent H3K4me3 and H3K27me3 modifications [131] (Figure 1). Moreover, the expression of DNMT1 and EZH2 shows a negative correlation with CD8+ T-cell infiltration within the TME and consequently the patient’s clinical outcome [132].

MicroRNAs also play a distinct role in shaping the immunosuppressed environment within the TMEs, thereby promoting tumor progression. For instance, the repression of miR-29 in cancer promotes augmentation of B7-H3, resulting in dysfunction of Natural Killer (NK) cells and immune evasion [133]. Also, the inhibition of perforin and granzyme expression by miR-27a* reduces anti-cancer cytotoxic functions of NK cells and cytotoxic T cells within the TME [134].

The role of TME components in tumor progression

A study by Hanson et al. in prostate cancer revealed differential promoter methylation status of various genes in epithelial and stromal cells in benign and malignant tumors [135]. It was shown that a distinct promoter methylation pattern in stromal cells is correlated with tumor development. The TME also plays a critical role in inducing intrinsic resistance to chemotherapy via ‘reversed pH gradient,’ pumping out protons through proton transporters compared with normal cells, which have low intracellular pH compared with extracellular pH [136]. The TME also adapts to dynamic changes as therapy continues, which leads to a gain in resistance to both chemo and targeted drugs. For instance, the CSC component of the TME plays a critical role in developing therapy resistance. The CSCs survive after radio or chemotherapy and shows resistance even towards one of the most novel and selective therapeutic system, immunotherapy [137]. The NSCCs, CSCs, and immune components of the TME, through their secretions, reciprocate feedback to each other to overcome therapy-induced stress and increase survival [138].

Intricate epigenetic regulatory pathways within the different components of TME play a crucial role in ensuring tumor progression. For example, in TAMs, activating epigenetic marks like histone phosphorylation by ERK-1/2 at the IL-10 promoter results in elevated production of IL-10, which stimulates immunosuppression [139]. Moreover, in TAMs, Tet2 reportedly sustains its immunosuppressive and tumor-promoting activity through DNA methylation [140]. In triple-negative breast cancer, LSD1 demethylase is allegedly involved in M2 polarization by activating the expression of IL-1b, IL-12b, IL-8, NOS2, CCR7, Gpr18 etc. [141]. Also, exosomal miR-21 and miR-1246 can enhance the production of IL-10 and TGF-β in primary human macrophages, leading to M2 activation [142] (Figure 1). M2 macrophages are established to have pro-tumor functions and have previously been associated with the acquisition of drug resistance [143]. In chemoresistant acute lymphoblastic leukemia, the interaction of cells within the bone marrow microenvironment is reported to be facilitated by epigenetic modifiers. Inhibition of these interactions using DNMT inhibitor azacitidine and HDACi panobinostat reportedly chemosensitizes these cells [144].

The components of core TME undergo hypoxic stress that causes elevated expression of hypoxia-inducible factor 1 α (HIF-1α). In hypoxic TME, reduced hydroxylation of HIF-1α favors its stabilization. Also, post-translational modifications of HIF-1α, including phosphorylation and SUMOylation have been reported to promote its stabilization [145,146]. Overexpression of HIF-1α orchestrates the transcription of various genes leading to the alteration in metabolism, activation of angiogenesis and metastasis, and the development of chemoresistance [147–149]. In addition, several microRNAs are reported to play an intricate part in the hypoxia-mediated acquisition of drug resistance. For instance, down-regulation of miR-15 and miR-16 due to hypoxia is associated with chemoresistance [150,151]. HIF-1α transcriptionally activates the expression of miR-20a, an elevated level of which has been found to induce therapy resistance [152,153]. Also, hypoxia-regulated expression of miR-200b reportedly promotes chemoresistance in cholangiocarcinoma [154].

All these studies cumulatively point towards the intricate involvement of epigenetic machinery in reprogramming and utilizing TME in a pro-tumor manner.

For advanced stages of cancer, a combination of chemotherapeutic agents is often favored over single agents to achieve a higher and quicker therapeutic response. Combination chemotherapy regimens are reported to show a better response in the case of tumor regression [155]. For instance, FAC (5-fluorouracil+adrenamycin+cyclophosphamide) and FEC (5-fluorouracil+adrenamycin+cyclophosphamide) are routinely used combinations for breast cancer. In metastatic breast carcinomas, Gemcitabine has also been utilized with paclitaxel for clinical trials [156]. CHOP (cyclophosphamide, doxorubicin hydrochloride, vincristine, and prednisone) is one of the most routinely used therapy in diffuse-large B-cell lymphoma [157]. However, combining two or more chemotherapeutic agents increases the amount of systemic toxicity and increases the chance of chemotherapy-related complications. To address this issue, several studies have utilized other drugs, small molecules, or pathway inhibitors in complement with chemotherapy to sensitize the drug-resistant cells. In recurrent ovarian cancers, VEGF and PARP inhibitors are combined with chemotherapeutic drugs [158]. The anti-diabetic drugs thiazolidinedione and metformin have been repurposed to target the chemoresistant CSCs by several studies, in which these drugs significantly sensitized the CSCs towards chemotherapy [159,160]. Curcumin, an active component obtained from turmeric (Curcuma longa) rhizome, has been well established by several studies as a potent chemosensitizing agent for various cancers [161–163]. The role of NFκB in the acquisition of drug resistance is well recognized by several studies [164,165]. The commonly used NSAID aspirin, due to its ability to inhibit nuclear transport of NFκB, has been extensively utilized by several studies to target cancer cells and CSCs [49,166,167]. Aspirin pre-treatment has been shown to suppress the acquisition of chemoresistance in breast CSCs [168]. Quite a handful of studies have pointed to the efficacy of using various miRNAs combined with chemotherapy to sensitize chemoresistance in cancer [169–172]. However, cancer cells’ extensive epigenetic reprogramming ability poses a major challenge to all the combinatorial approaches and even increases the possibility of acquiring resistance towards these therapies.

Various studies so far have established that in addition to genetic mutations, epigenetic modifications play a pivotal role in acquiring chemoresistance in cancer cells. However, the epigenetic alternations are reversible and require to be actively maintained by epigenetic modifiers, making them an attractive target for therapeutic intervention. The current understanding of CSC epigenome provides new insights into targeted therapy by epigenetic drugs to overcome CSC drug resistance [173]. Several studies have utilized epigenome modifying drugs alone or combined with conventional treatments to modulate the epigenetic changes to attenuate drug resistance [174] (Figure 2). Although these epigenome-altering drugs are ‘non-specific’ in nature and may affect global gene expression, it is also reported that they can cause local changes in gene expression depending on the gene chromatin environment [175]. Furthermore, it is reported that the sensitivity towards epigenetic alternators can be genomic loci-specific depending on the three-dimensional structure of chromatin [175].

Epidrugs are used to target and sensitize CSCs

Figure 2
Epidrugs are used to target and sensitize CSCs

Several studies have utilized different epigenetic modulators including DNMT inhibitors, HMT inhibitors, BET inhibitors, and HDACis, to target therapy resistance, self-renewal, proliferation, and migratory potential of CSCs and to sensitize them towards chemotherapy. Abbreviation: AZA, azacytidine; DAC, decitabine; E-Cad, E-cadherin; HMT, histone methyltransferase; N-Cad, N-cadherin; OAA, oxaloacetic acid; SAHA, suberoylanilide hydroxamic acid; TSA, trichostatin A; Vim, vimentin; VPA, valproic acid; α-KG, α-ketoglutarate.

Figure 2
Epidrugs are used to target and sensitize CSCs

Several studies have utilized different epigenetic modulators including DNMT inhibitors, HMT inhibitors, BET inhibitors, and HDACis, to target therapy resistance, self-renewal, proliferation, and migratory potential of CSCs and to sensitize them towards chemotherapy. Abbreviation: AZA, azacytidine; DAC, decitabine; E-Cad, E-cadherin; HMT, histone methyltransferase; N-Cad, N-cadherin; OAA, oxaloacetic acid; SAHA, suberoylanilide hydroxamic acid; TSA, trichostatin A; Vim, vimentin; VPA, valproic acid; α-KG, α-ketoglutarate.

Close modal

DNMTs, HDACs, and histone methyltransferases (HMTs) are among the most widely targeted epigenetic modifiers in contemporary anti-cancer therapies. Inhibitors of DNMTs, HDACs, histone demethylases (HDMs), HMTs and bromodomain proteins are used individually or in combinations to various clinical trials with or without chemotherapy [176–178]. Table 1 enlists different types of epigenome-modifying drugs and their utilization, targeting multiple types of solid cancers and hematological malignancies.

Table 1
Various epigenetic modulators in anti-cancer studies and clinical trials
TypeCompoundActivityStudies done in cancersStatus of clinical trial
DNMT inhibitors Azacitidine Inhibits DNMT activity AML [179], prostate [180], multiple myeloma [181], non-small cell lung carcinoma (NSCLC) [182• AML: Phase III (completed- NCT00887068; active- NCT01757535)
• Prostate cancer: Phase II (Completed- NCT00384839)
• MDS: Phase III and IV (completed- NCT00071799, NCT01201811)
• Pancreatic cancer: Phase II (Recruiting- NCT01845805)
• CML: Phase II (completed-NCT01350947)
• Refractory T-cell lymphoma: Phase III (recruiting- NCT03703375)
• Head and neck squamous cell carcinoma: Phase II (recruiting- NCT02178072) 
 Decitabine Inhibits DNMT1 activity Bladder carcinoma [183], glioblastoma [184,185], hepatocellular carcinoma [186], renal cell carcinoma [187], acute leukemia [188• AML: Phase II (completed- NCT00416598)
• Refractory CML: Phase II (completed- NCT00042003)
• MDS: Phase II and III (completed- NCT00067808 and NCT00043381)
• Refractory diffuse-large B-cell lymphoma: Phase IV (recruiting- NCT03579082)
• Refractory T lymphoblastic lymphoma: Phase IV (recruiting- NCT03558412)
• Follicular thyroid cancer: Phase II (completed- NCT00085293) 
 Procaine Prevents DNMT1 and 3A from binding to DNA Gastric cancer [189], nasopharyngeal cancer [190• Nasopharyngeal neoplasms: Phase II (mixture of Procaine with dexamethasone, gentamicin and vitamin B12; unknown status-NCT02735317) 
 Zebularine Inhibits DNMT activity Ovarian cancer [191No clinical trial yet 
HDACi Vorinostat/SAHA Inhibits class I and II HDACs Leukemia [192], glioblastoma multiforme (GBM) [193], breast cancer [194,195], melanoma [196], pancreatic cancer [197], neuroblastoma [198], retinoblastoma [199], NSCLC [200• GBM: Phase II (vorinostat + temozolomide + radiation, active- NCT00731731)
• Advanced thyroid cancer: Phase II (completed- NCT00134043)
• Kidney cancer: Phase II (completed- NCT00278395)
• Advanced NSCLC: Phase II (completed- NCT00138203)
• Progressive GBM: Phase II (completed- NCT00238303)
• Breast cancer: Phase II (vorinostat + tamoxifen, completed- NCT00365599
• Progressive metastatic prostate cancer: Phase II (completed- NCT00330161)
• Advanced malignant pleural mesothelioma: Phase III (completed- NCT00128102) 
 Abexinostat Inhibits class I and II HDACs Nasopharyngeal carcinoma [201], breast cancer [202• Renal cell carcinoma: Phase III (active, recruiting- NCT03592472)
• Refractory follicular lymphoma: Phase II (active, not recruiting- NCT03600441)
• Non-Hodgkin’s lymphoma: Phase II (active, recruiting- NCT04024696) 
 Belinostat Pan HDACi NSCLC [203], Non-Hodgkin’s lymphoma [204], AML [205• Refractory peripheral T-cell lymphoma: Phase II (completed- NCT00865969)
• Liver cancer: Phase I/II (completed- NCT00321594)
• MDS- Phase II (completed NCT00357162)
• Advanced multiple myeloma: Phase II (completed-NCT00131261) 
 Panobinostat Pan HDACi inhibitor NSCLC [206], [207], diffuse large B-cell lymphoma [208], AML and MDS [209–211], ovarian cancer [212• High-risk MDS, AML: Phase I/II (active, not recruiting-NCT01451268)
• Metastatic thyroid cancer: Phase II (completed- NCT01013597)
• Refractory prostate cancer: Phase II (completed- NCT00667862)
• Refractory colorectal cancer: Phase II (completed-NCT00690677)
• Refractory CML: Phase II/III (completed- NCT00449761)
• HER2-negative locally recurrent or metastatic breast cancer: Phase II (completed- NCT00777049) 
 Valproic Acid Pan inhibitor, binding to the catalytic center of the HDACs Melanoma [213,214], colon cancer [215], Prostate cancer [216], Breast cancer [217,218], lung cancer [219], head and neck cancer [220], thyroid cancer [221,222• Melanoma: Phase II (active, recruiting- NCT0206858)
• Colorectal cancer: Phase I/II (valproic acid + radiation therapy, recruiting- NCT01898104)
• Advanced thyroid cancers: Phase II (completed- NCT01182285)
• High-grade gliomas, brain tumors: Phase II (Valproic acid + temozolomide + radiation, completed- NCT00302159)
• NSCLC- Phase I/II (valproic acid + chemoradiotherapy, status unknown- NCT01203735
• Pancreatic cancer: Phase II (valproic acid+ chemotherapy, status unknown- NCT01333631) 
 Phenylbutyrate Pan HDACi Renal cell carcinoma [223], NSCLC [224• Progressive or recurrent brain tumors: Phase II (completed-NCT00006450) 
 Entinostat Class 1 HDACi Breast cancer [225,226], Renal cell carcinoma and prostate cancer [227• Advanced breast cancer: Phase III (recruiting- NCT03538171), Phase II (completed- NCT00676663)
• MDS, AML, ALL: Phase II (completed- NCT00462605)\
• Metastatic melanoma: Phase II (completed- NCT00185302) 
 Givinostat Class 1 and 2 HDACi NSCLC [228• Chronic myeloproliferative neoplasms: Phase II (active, not recruiting- NCT01761968) 
 Romidepsin (FK228, Depsipeptide) Class 1 HDACi Clear cell renal cell carcinoma (ccRCC) and triple-negative breast cancer (TNBC)[229], lung cancer [230], breast cancer[231], bladder cancer[232], T-cell lymphoma [233• Progressive peripheral T-cell lymphoma: Phase II (active, not recruiting- NCT00426764, completed NCT00007345)
• Metastatic breast cancer: Phase II (completed- NCT00098397)
• Relapsed small-cell lung cancer: Phase II (completed-NCT00086827)
• Recurrent high-grade gliomas: Phase I/II (completed- NCT00085540)
• Relapsed or refractory multiple myeloma: Phase II (completed-NCT00066638) 
 Trichostatin A (TSA) Class 1/3/4/6/10 HDACs inhibitor NSCLC[234], prostate cancer [235], bladder cancer [236], CML[237], glioblastoma [238• Relapsed or refractory hematologic malignancies: Phase I (recruiting-NCT03838926) 
HDM inhibitors Pargyline Inhibits lysine-specific demethylase 1 (LSD1) Breast cancer [239], prostate cancer [240No clinical trial yet 
 HCI-2509 Inhibits LSD1 Lung adenocarcinoma [241No clinical trial yet 
 S2101 Inhibits LSD1 Ovarian cancer [242No clinical trial yet 
 MC3324 Inhibits LSD1 and LSD6A Breast cancer [243No clinical trial yet 
 Methylstat Inhibits LSD4B Breast cancer [244No clinical trial yet 
 JIB-04 Pan inhibitor of Jumanji-domain histone demethylases Ewing sarcoma [245No clinical trial yet 
HMT inhibitors Tazemetostat Inhibits EZH2 Solid tumors and hematological malignancies [246], refractory follicular lymphoma [247, 248], thyroid cancer [249], colorectal cancer [250], breast cancer [251• Malignant mesothelioma: Phase II (completed- NCT02860286)
• Metastatic prostate cancer: Phase I (recruiting, NCT04179864, NCT04846478)
• Solid tumors harboring an ARID1A mutation: Phase II (not yet recruiting- NCT05023655)
• Refractory INI1-Negative tumors or synovial sarcoma: Phase I (completed- NCT02601937)
• Peripheral nerve sheath tumor: Phase II (recruiting- NCT04917042)
• Metastatic melanoma: Phase II (recruiting- NCT04557956)
• Advanced colorectal carcinoma, advanced soft-tissue sarcoma, advanced pancreatic adenocarcinoma: Phase II (recruiting- NCT04705818) 
 CPI-1205 Inhibits EZH2 Small intestinal neuroendocrine tumors [252], B-cell lymphomas [253• B-cell lymphoma: Phase I (completed- NCT02395601)
• Metastatic castration-resistant prostate cancer: Phase I/II (active, not recruiting- NCT03480646) 
 PF-0682149 Inhibits EZH2 Brain cancers [254• Small-cell lung cancer, follicular lymphoma and castration-resistant prostate cancer: Phase I (active, recruiting- NCT03460977) 
 EPZ-5676 (pinometostat) Inhibits DOT1L Leukemia [255,256• Refractory leukemias bearing a rearrangement of the MLL gene: Phase I (completed- NCT02141828) 
BET inhibitors JQ1 Inhibits BET proteins BRD2, BRD3, BRD4, and BRDT Nuclear protein in testis (NUT)-midline carcinoma (NMC) [257], AML [258], medulloblastoma [259], breast cancer [260], and lung cancer [261• No clinical trial due to low oral bioavailability 
 JQ1 analog RO6870810 (TEN-010/JQ2) Inhibits BET proteins BRD2, BRD3, BRD4, and BRDT Advance multiple myeloma [262], AML and MDS [263• AML and MDS: Phase I (completed- NCT02308761)
• NUT carcinoma, diffuse large B cell lymphoma and other advanced solid tumors: Phase I (completed- NCT01987362) 
 OTX015 Inhibits BET proteins BRD2, BRD3, BRD4 Neuroblastoma [264], mesothelioma [265], multiple myeloma [266], and B-cell lymphoma [267• AML, diffuse large B-cell lymphoma, ALL, multiple myeloma: Phase I (completed- NCT01713582) 
 I-BET62 (Molibresib, GSK525762) Inhibits BET proteins BRD2, BRD3, BRD4 Multiple myeloma[268], pancreatic adenocarcinoma [269], and neuroblastoma [270• Metastatic breast cancer: Phase I (completed- NCT02964507)
• Castrate-resistant prostate cancer: Phase I (completed- NCT03150056)
• Refractory hematologic malignancies: Phase II (completed- NCT01943851) 
TypeCompoundActivityStudies done in cancersStatus of clinical trial
DNMT inhibitors Azacitidine Inhibits DNMT activity AML [179], prostate [180], multiple myeloma [181], non-small cell lung carcinoma (NSCLC) [182• AML: Phase III (completed- NCT00887068; active- NCT01757535)
• Prostate cancer: Phase II (Completed- NCT00384839)
• MDS: Phase III and IV (completed- NCT00071799, NCT01201811)
• Pancreatic cancer: Phase II (Recruiting- NCT01845805)
• CML: Phase II (completed-NCT01350947)
• Refractory T-cell lymphoma: Phase III (recruiting- NCT03703375)
• Head and neck squamous cell carcinoma: Phase II (recruiting- NCT02178072) 
 Decitabine Inhibits DNMT1 activity Bladder carcinoma [183], glioblastoma [184,185], hepatocellular carcinoma [186], renal cell carcinoma [187], acute leukemia [188• AML: Phase II (completed- NCT00416598)
• Refractory CML: Phase II (completed- NCT00042003)
• MDS: Phase II and III (completed- NCT00067808 and NCT00043381)
• Refractory diffuse-large B-cell lymphoma: Phase IV (recruiting- NCT03579082)
• Refractory T lymphoblastic lymphoma: Phase IV (recruiting- NCT03558412)
• Follicular thyroid cancer: Phase II (completed- NCT00085293) 
 Procaine Prevents DNMT1 and 3A from binding to DNA Gastric cancer [189], nasopharyngeal cancer [190• Nasopharyngeal neoplasms: Phase II (mixture of Procaine with dexamethasone, gentamicin and vitamin B12; unknown status-NCT02735317) 
 Zebularine Inhibits DNMT activity Ovarian cancer [191No clinical trial yet 
HDACi Vorinostat/SAHA Inhibits class I and II HDACs Leukemia [192], glioblastoma multiforme (GBM) [193], breast cancer [194,195], melanoma [196], pancreatic cancer [197], neuroblastoma [198], retinoblastoma [199], NSCLC [200• GBM: Phase II (vorinostat + temozolomide + radiation, active- NCT00731731)
• Advanced thyroid cancer: Phase II (completed- NCT00134043)
• Kidney cancer: Phase II (completed- NCT00278395)
• Advanced NSCLC: Phase II (completed- NCT00138203)
• Progressive GBM: Phase II (completed- NCT00238303)
• Breast cancer: Phase II (vorinostat + tamoxifen, completed- NCT00365599
• Progressive metastatic prostate cancer: Phase II (completed- NCT00330161)
• Advanced malignant pleural mesothelioma: Phase III (completed- NCT00128102) 
 Abexinostat Inhibits class I and II HDACs Nasopharyngeal carcinoma [201], breast cancer [202• Renal cell carcinoma: Phase III (active, recruiting- NCT03592472)
• Refractory follicular lymphoma: Phase II (active, not recruiting- NCT03600441)
• Non-Hodgkin’s lymphoma: Phase II (active, recruiting- NCT04024696) 
 Belinostat Pan HDACi NSCLC [203], Non-Hodgkin’s lymphoma [204], AML [205• Refractory peripheral T-cell lymphoma: Phase II (completed- NCT00865969)
• Liver cancer: Phase I/II (completed- NCT00321594)
• MDS- Phase II (completed NCT00357162)
• Advanced multiple myeloma: Phase II (completed-NCT00131261) 
 Panobinostat Pan HDACi inhibitor NSCLC [206], [207], diffuse large B-cell lymphoma [208], AML and MDS [209–211], ovarian cancer [212• High-risk MDS, AML: Phase I/II (active, not recruiting-NCT01451268)
• Metastatic thyroid cancer: Phase II (completed- NCT01013597)
• Refractory prostate cancer: Phase II (completed- NCT00667862)
• Refractory colorectal cancer: Phase II (completed-NCT00690677)
• Refractory CML: Phase II/III (completed- NCT00449761)
• HER2-negative locally recurrent or metastatic breast cancer: Phase II (completed- NCT00777049) 
 Valproic Acid Pan inhibitor, binding to the catalytic center of the HDACs Melanoma [213,214], colon cancer [215], Prostate cancer [216], Breast cancer [217,218], lung cancer [219], head and neck cancer [220], thyroid cancer [221,222• Melanoma: Phase II (active, recruiting- NCT0206858)
• Colorectal cancer: Phase I/II (valproic acid + radiation therapy, recruiting- NCT01898104)
• Advanced thyroid cancers: Phase II (completed- NCT01182285)
• High-grade gliomas, brain tumors: Phase II (Valproic acid + temozolomide + radiation, completed- NCT00302159)
• NSCLC- Phase I/II (valproic acid + chemoradiotherapy, status unknown- NCT01203735
• Pancreatic cancer: Phase II (valproic acid+ chemotherapy, status unknown- NCT01333631) 
 Phenylbutyrate Pan HDACi Renal cell carcinoma [223], NSCLC [224• Progressive or recurrent brain tumors: Phase II (completed-NCT00006450) 
 Entinostat Class 1 HDACi Breast cancer [225,226], Renal cell carcinoma and prostate cancer [227• Advanced breast cancer: Phase III (recruiting- NCT03538171), Phase II (completed- NCT00676663)
• MDS, AML, ALL: Phase II (completed- NCT00462605)\
• Metastatic melanoma: Phase II (completed- NCT00185302) 
 Givinostat Class 1 and 2 HDACi NSCLC [228• Chronic myeloproliferative neoplasms: Phase II (active, not recruiting- NCT01761968) 
 Romidepsin (FK228, Depsipeptide) Class 1 HDACi Clear cell renal cell carcinoma (ccRCC) and triple-negative breast cancer (TNBC)[229], lung cancer [230], breast cancer[231], bladder cancer[232], T-cell lymphoma [233• Progressive peripheral T-cell lymphoma: Phase II (active, not recruiting- NCT00426764, completed NCT00007345)
• Metastatic breast cancer: Phase II (completed- NCT00098397)
• Relapsed small-cell lung cancer: Phase II (completed-NCT00086827)
• Recurrent high-grade gliomas: Phase I/II (completed- NCT00085540)
• Relapsed or refractory multiple myeloma: Phase II (completed-NCT00066638) 
 Trichostatin A (TSA) Class 1/3/4/6/10 HDACs inhibitor NSCLC[234], prostate cancer [235], bladder cancer [236], CML[237], glioblastoma [238• Relapsed or refractory hematologic malignancies: Phase I (recruiting-NCT03838926) 
HDM inhibitors Pargyline Inhibits lysine-specific demethylase 1 (LSD1) Breast cancer [239], prostate cancer [240No clinical trial yet 
 HCI-2509 Inhibits LSD1 Lung adenocarcinoma [241No clinical trial yet 
 S2101 Inhibits LSD1 Ovarian cancer [242No clinical trial yet 
 MC3324 Inhibits LSD1 and LSD6A Breast cancer [243No clinical trial yet 
 Methylstat Inhibits LSD4B Breast cancer [244No clinical trial yet 
 JIB-04 Pan inhibitor of Jumanji-domain histone demethylases Ewing sarcoma [245No clinical trial yet 
HMT inhibitors Tazemetostat Inhibits EZH2 Solid tumors and hematological malignancies [246], refractory follicular lymphoma [247, 248], thyroid cancer [249], colorectal cancer [250], breast cancer [251• Malignant mesothelioma: Phase II (completed- NCT02860286)
• Metastatic prostate cancer: Phase I (recruiting, NCT04179864, NCT04846478)
• Solid tumors harboring an ARID1A mutation: Phase II (not yet recruiting- NCT05023655)
• Refractory INI1-Negative tumors or synovial sarcoma: Phase I (completed- NCT02601937)
• Peripheral nerve sheath tumor: Phase II (recruiting- NCT04917042)
• Metastatic melanoma: Phase II (recruiting- NCT04557956)
• Advanced colorectal carcinoma, advanced soft-tissue sarcoma, advanced pancreatic adenocarcinoma: Phase II (recruiting- NCT04705818) 
 CPI-1205 Inhibits EZH2 Small intestinal neuroendocrine tumors [252], B-cell lymphomas [253• B-cell lymphoma: Phase I (completed- NCT02395601)
• Metastatic castration-resistant prostate cancer: Phase I/II (active, not recruiting- NCT03480646) 
 PF-0682149 Inhibits EZH2 Brain cancers [254• Small-cell lung cancer, follicular lymphoma and castration-resistant prostate cancer: Phase I (active, recruiting- NCT03460977) 
 EPZ-5676 (pinometostat) Inhibits DOT1L Leukemia [255,256• Refractory leukemias bearing a rearrangement of the MLL gene: Phase I (completed- NCT02141828) 
BET inhibitors JQ1 Inhibits BET proteins BRD2, BRD3, BRD4, and BRDT Nuclear protein in testis (NUT)-midline carcinoma (NMC) [257], AML [258], medulloblastoma [259], breast cancer [260], and lung cancer [261• No clinical trial due to low oral bioavailability 
 JQ1 analog RO6870810 (TEN-010/JQ2) Inhibits BET proteins BRD2, BRD3, BRD4, and BRDT Advance multiple myeloma [262], AML and MDS [263• AML and MDS: Phase I (completed- NCT02308761)
• NUT carcinoma, diffuse large B cell lymphoma and other advanced solid tumors: Phase I (completed- NCT01987362) 
 OTX015 Inhibits BET proteins BRD2, BRD3, BRD4 Neuroblastoma [264], mesothelioma [265], multiple myeloma [266], and B-cell lymphoma [267• AML, diffuse large B-cell lymphoma, ALL, multiple myeloma: Phase I (completed- NCT01713582) 
 I-BET62 (Molibresib, GSK525762) Inhibits BET proteins BRD2, BRD3, BRD4 Multiple myeloma[268], pancreatic adenocarcinoma [269], and neuroblastoma [270• Metastatic breast cancer: Phase I (completed- NCT02964507)
• Castrate-resistant prostate cancer: Phase I (completed- NCT03150056)
• Refractory hematologic malignancies: Phase II (completed- NCT01943851) 

Use of DNMT inhibitors in hindering chemoresistance

DNMT inhibitors are extensively used as tools for hypomethylating the genome in preclinical and clinical studies and the treatment of different types of cancers. Currently, two DNMT inhibitors 5-Azacytidine (AZA, vidaza) and its deoxyribose analog, 5-aza-2′-deoxycytidine (decitabine), are approved by Food and Drug Administration (FDA) as anti-cancer agents. AZA and decitabine are the only epigenetic modulators permitted as therapeutic agents for patients with acute myeloid leukemia (AML) with resistance to conventional chemotherapy. AZA has also been approved for the treatment of chronic myelomonocytic leukemia (CMML) by the European Medicines Agency (EMA) and FDA [271].

Several studies have utilized AZA and decitabine to promote cell cycle arrest and cytotoxicity in resistant cancer cells [272,273]. In multiple myeloma, AZA reportedly leads to modulation of p16 expression, cleavage of caspase and PARP, and cell cycle arrest at G0/G1 phase [274]. Similarly, decitabine induces G0/G1 and G2/M arrest by modulating p21 and p38 [275]; the effectivity of decitabine at the clinical level has been well established in AML, chronic myelogenous leukemia (CML), and myelodysplasia [276].

Clinical trial status

A novel compound, termed SGI-110 (combination of 5-aza-2′-deoxycytidine and guanosine), functions as a prodrug for decitabine and is reported to display promising results against drug-resistant AML and myelodysplasia in Phase III clinical trials (NCT02348489) [277] (https://clinicaltrials.gov/). Both AZA and decitabine have undergone many clinical trials before being accepted by FDA for the treatment of several hematological cancers (Table 1). Currently, clinical trials are going on to elucidate their dosage and potency in treating various solid tumors. Subsequent quests for more efficacious DNMT inhibitors have led to the synthesis of MG98, a second-generation phosphorothioate antisense oligodeoxynucleotide that prevents translation of DNMT1 mRNA. MG98 is currently undergoing phase I/II clinical trials in solid tumors’ patients (NCT00003890) [278]. The clinical trial status of various DNMT inhibitors is listed in Table 1.

HDACis as epi-drugs against chemoresistant cancers

The HDAC family constitutes a major target of the epigenome-modifying drugs as HDAC inhibition can trigger a magnitude of cellular responses, promoting intracellular stress and eventually cell death. The role of HDACs in maintaining chemoresistance has been well established by various studies [49,279,280], and the use of HDACis has been in light for several years now. The two HDACis suberoylanilide hydroxamic acid (SAHA, class I and II inhibitor) and depsipeptide (romidepsin, class I inhibitor) have been permitted by FDA for treating T-cell lymphomas [281,282]. SAHA is also reported to induce apoptosis in multiple myelomas through dephosphorylation of Rb, attenuation of Bcl2, and augmentation of p21 and p53 [283]. In multiple myeloma, a combination of Panobinostat, a non-selective pan HDACi, and bortezomib is reported to have synergistic activity against Dexamethasone resistance [284]. The active DNA repair pathway is one of the foremost causes behind chemotherapeutic resistance [285]. Studies on various HDACis in different solid and liquid cancers have shown pronounced effects against this active DNA repair system, ultimately hindering the drug resistance ability of the cancer cells. For instance, in AML, a co-treatment with HDACi belinostat and NEDD8-activating enzyme inhibitor pevonedistat reportedly inhibits expression of the DNA repair proteins 53BP1 and RAD51 [205]. In neuroblastoma, SAHA is reported to repress DNA repair-mediated acquisition of resistance by inhibiting Ku-86, a significant contributor of non-homologous end-joining DNA damage repair [286]. The pan HDACi valproic acid has been used in clinical trials singly or in combination with other drugs as a tool to target cancer cell resistance. For instance, in neuroendocrine carcinoma and central nervous system (CNS) tumors, treatment with valproic acid displayed significant stability in the patients [287,288]. Studies reporting valproic acid treatment in combination with various drugs like S-1 (a novel oral fluoropyrimidine derivative), decitabine, bevacizumab have indicated effective anti-tumor activity [289–291].

Clinical trial status

Compared with DNMTis, HDACis have been studied in several diseases, including hematological malignancies, solid tumors and inflammatory disorders, and have elicited positive outcomes. The three FDA-approved HDACis SAHA, romidepsin, and belinostat, are approved as a therapeutic tool in cutaneous T-cell lymphoma. They are being extensively assessed in the clinical trials for other solid and hematological cancers, either alone or co-treatment with other anti-cancer drugs. For instance, in glioblastoma patients, SAHA is currently under clinical trial combined with radiotherapy and chemotherapeutic drug temozolomide (NCT00731731). A Phase II clinical using romidepsin in peripheral T-cell lymphoma showed an overall 38% response rate [292]. According to https://clinicaltrials.gov/, approximately 30 clinical trials are ongoing with romidepsin, either as a single agent or combined with other drugs. In Phase II clinical trials, belinostat has been utilized to sensitize platinum-resistant epithelial cancer and micropapillary ovarian tumors [293]. Another phase I/II clinical trial of belinostat in thymic epithelial tumors indicated potency of belinostat-PAC (P: cisplatin, A: adriamycin/doxorubicin C: cyclophosphamide) co-treatment in reducing tumor resistance [294].

The preclinical and clinical studies using the different HDACis are enlisted in Table 1.

HMT inhibitors in anti-cancer therapeutics

Reversible histone methylation plays a crucial role in the development and progression of various cancers and histone methyltransferase inhibitors (HMTis) are emerging as potential epigenome-modifying tools with promising outcomes in the arena of clinical oncology. Some of the lysine and arginine methyltransferases widely studied in anti-cancer therapeutics include G9a, EZH2, DOT1L (disruptor of telomeric silencing 1-like protein) and PRMT1 (protein arginine methyltransferases 1), inhibitors of which are being extensively investigated [295].

The expression and activity of EZH2, one of the most well-known HMTs, have been associated with the tumor progression of various cancers [295]. Enhanced EZH2 activity is reported to induce tumor growth and resistance to conventional chemotherapy, leading to poor prognostic outcomes [296]. The EZH2 inhibitor tazemetostat, the first FDA-approved HMT inhibitor for cancer treatment, is currently one of the most investigated compounds in the clinical trial, with its anti-cancer properties established in preclinical studies [297,298]. Other selective inhibitors of EZH2 such as PF-06821497, GSK2816126, are also reported to reduce tumor growth and progression in preclinical models [299,300].

EPZ-5676 or pinometostat, a specific inhibitor against H3K79 methyltransferase DOT1L, has been under preclinical investigation for some years and has shown potency in treating the leukemia model [301].

PRMTs, the HMTs of the arginine methylase family, play a crucial role in cancer signaling pathways. Different classes of PRMTs are reported to be up-regulated and play pro-cancer roles in various solid and hematological cancers [302–304]. Inhibition of PRMTs has displayed anti-cancer potential in preclinical studies. PF-06939999 and EPZ015666 (GSK3235025), potent orally bioavailable PRMT5 inhibitors, have been reported to show significant antitumor activity [305].

Clinical trial status

The EZH2 inhibitors tazemetostat CPI-0209, CPI-1205, GSK2816126, DS-3201, and PF-06821497 have been in clinical trials. Tazemetostat, the most investigated EZH2 inhibitor, has undergone several clinical trials in different types of cancers and has demonstrated potent anti-tumor effects (enlisted in Table 1). In patients with refractory malignant mesothelioma with BAP1 inactivation, a Phase II clinical study using tazemetostat has established its activity as an anti-cancer agent with low toxicity (NCT02860286). However, several other clinical trials of tazemetostat in various cancers are still in progress.

EPZ-5676 has been under clinical trials in patients with refractory leukemia (NCT01684150, NCT02141828) and has shown modest clinical activity and an acceptable safety profile (Table 1) [255].

Three PRMT5 inhibitors, PF-06939999, GSK3326595 and JNJ-6461978, are currently under clinical trials in patients with solid and hematologic cancers. GSK3368715, the only developed inhibitor of PRMT1, has recently entered Phase I clinical trials in patients with various solid and hematologic malignancies (NCT03666988).

Bromodomain inhibitors as anti-cancer agents

Bromodomain (BRD) proteins are ‘readers’ of histone acetylation, which target transcription factors and chromatin-remodeling enzymes to target sites on the chromatin to regulate gene expression. BRD proteins are generally associated with the augmentation of target genes [306]. Bromo- and extra-terminal domain (BET) proteins, a crucial subfamily of BRD proteins, constitute BRDT, BRD2, BRD3 and BRD4. Due to their diverse genome-wide targets, BET proteins play significant roles in cellular survival and homeostasis, dysregulation of which leads to malignancy [307]. Over the recent years, targeting BET proteins through small molecule inhibitors has become a highlight of anti-cancer research to sensitize drug-resistant cancer cells towards therapy. JQ1 and I-BET762 (GSK525762A/molibresib/I-BET) are the two foremost BET inhibitors that have been extensively studied as cancer therapeutics over the past few years. JQ1 has been reported to induce anti-cancer effects in nuclear protein in testis (NUT) midline carcinoma (NMC), AML, medulloblastoma, breast cancer, and lung cancer [257–261]. Due to the low oral bioavailability and short half-life of JQ1, it has not been used in clinical trials [307]. OTX015 (Birabresib), another small molecule similar to JQ1, inhibits BRD2, BRD3 and BRD4 supports oral administration and has been reported to display anti-cancer effects in preclinical studies of drug-resistant neuroblastoma [264], mesothelioma [265], multiple myeloma [266], and B-cell lymphoma [267]. I-BET62 is orally bioavailable and has shown efficacy as an anti-tumor agent in preclinical models, including multiple myeloma [268], pancreatic adenocarcinoma [269], and neuroblastoma [270].

Clinical trial status

Consecutive research has led to the development of JQ1 analog TEN-010 (JQ2), which has completed Phase I clinical trial in various advanced solid tumors (NCT01987362), AML, and MDS patients (NCT02308761) [262,263] (Table 1). However, OTX015 ha snot produced any significant positive outcome in clinical trials as the patients administered with OTX015 displayed symptoms of dose-limited toxicities like fatigue and headache, gastrointestinal disorders, anemia, hyperbilirubinemia etc. [308,309] (Table 1). I-BET62 is currently under clinical trial to specify doses [310].

Usage of epi-drugs in targeting CSCs

Since CSCs are the major contributors behind drug resistance within the tumor mass, several studies have utilized epi-drugs to specifically target CSCs and sensitize them towards conventional therapies to achieve significant tumor regression (Figure 2). For instance, DNMT inhibitors AZA and decitabine have been reported to selectively target CSCs of leukemia and bladder cancer, thereby improving the efficacy of chemotherapy [311,312]. Inhibition of EZH2 and BET proteins by specific inhibitors have also been reported to repress CSC-associated tumor aggressiveness in different solid cancers [313–315]. Similarly, potent HDACis including valproic acid, SAHA and trichostatin A (TSA) have been shown to selectively target CSCs and repress their self-renewal, proliferation, migratory and therapy-resistant properties [316–318]. In a nutshell, all these reports indicate towards the crucial significance of targeting CSCs to alleviate therapy resistance and achieve overall tumor regression.

Immunotherapy or enhancing the patient’s immune response to eradicate cancer cells has become a breakthrough in anti-cancer therapeutics. But, a substantial percentage of nonresponding patients and systemic toxicities pose an obstacle to its therapeutic success [319]. An immunosuppressive tumor milieu can lead to epigenetic modification of tumor-associated immune cells, promoting therapeutic failure [132]. In this context, modulating epigenomic signatures of cancer cells and cancer-associated immune cells show promising prognostic outcomes in patient’s responses to immunotherapy.

Epigenetic reprogramming of tumor cells, TME, and tumor-associated immune cells play crucial roles in developing an antitumor immune response, patient’s response to immunotherapy, and overall prognostic outcome of the patient. For instance, derepression of endogenous retroviruses (ERVs) by LSD1 inhibitors have been reported to lead to accumulation of double-stranded RNAs (dsRNAs) in cancer cells, which simulates viral infection, in turn triggering antitumoral immunity through interferon response within the tumor milieu [320]. Another study has reported that dual inhibition of DNMTs and methyltransferase G9a in ovarian and hematological cancers leads to an augmentation of ERV transcripts and consequent induction of viral defense genes like IRF7 and STAT1 [321]. AZA treatment in colorectal cancer has been reported to promote the expression of interferon-response factors like OASL and IRF7 by inducing up-regulation of dsRNA and stimulation of the MDA5/MAVS/IRF7 pathway [322]. Inhibition of EZH2 in the ovarian cancer model has been reported to enhance the expression of Th1 chemokine genes (Cxcl9 and Cxcl10), promoting CD8+ T-cell infiltration and leading to the efficacy of anti-PD-L1 antibody immunotherapy [132]. Moreover, due to the crucial role of EZH2 in the differentiation, and maintenance of Treg cells and the expansion of effector T cells, makes it a potential target of immune-modulation therapy [323]. Repression of EZH2 in tumor-infiltrating Treg cells thus promote pro-inflammatory signaling and enhance recruitment of CD8+ and CD4+ effector T cells, ultimately leading to tumor regression [324].

All these reports indicate that targeting epigenetic modulators can prompt significant anti-tumor immunity and thus can serve as a complementary therapeutic approach to current conventional therapies.

Cancer is the principal cause of global morbidity and mortality, with billions of people succumbing to it each year. The most routinely used therapeutic regimen for most advanced cancers is chemotherapy, which has led to no significant improvement in the treatment scenario and mortality worldwide. Approximately 70–80% of patients with advanced carcinomas experience recurrence, which leads to an even more aggressive form of the disease and is mostly fatal. The development of the recurrent tumor is mainly attributed to drug resistance as the therapy-resistant cells remaining after first-line therapy, serve as the seed for future relapse. Drug resistance involves a complex architecture of molecular events, including heightened drug efflux, drug inactivation, inhibition of apoptosis, and enhanced DNA repair mechanisms. This resistance towards the conventional anti-cancer stratagems, be inherent or acquired, has hindered favorable prognosis since the emergence of the chemotherapy and seems to be a prime impediment in the effective management of breast cancer. Extensive research for the past decade has established that CSCs are the primary drug-resistant cells within the tumor mass, which remains unaffected even after the chemotherapeutic cycles and finally gives rise to recurrent tumors. This indicates the importance of developing CSC targeting therapies that could sensitize the drug-resistant tumor towards conventional chemotherapy and hence will be able to complement it. Combinatorial therapies involving the usage of various drugs or pathway inhibitors and chemotherapy have been highlighted for the past few years due to their efficacy in reducing tumor burden and attenuation of drug resistance.

Epigenetic modifications play a crucial role in the acquisition of therapy resistance. However, even these combinatorial therapies often fall prey to the ingenious epigenetic reprogramming of the cancer cells, leading to treatment failure after a certain point in time. In this scenario, epigenome modifying drugs seem to be an attractive option to serve as a co-treatment procedure along with mainstream/combinatorial therapies. Hence, it is crucial to continue unravelling the intricate epigenetic regulatory mechanisms underlying the acquisition of drug resistance to keep developing more efficacious treatment possibilities. This review encompasses the currently known epigenetic landscape, which regulates the mechanisms of drug resistance in cancer. Further studies are required to enhance overall understanding and devise novel therapeutic strategies to target drug-resistant cancers.

The authors declare that there are no competing interests associated with the manuscript.

This work was supported by the Swarnajayanti Fellowship by the Department of Science and Technology (DST) [grant number DST/SJF/LSA-02/2017-18]; the S. Ramachandran National Bioscience Award for Career Development 2019 by the Department of Biotechnology (DBT) [grant number BT/HRD-NBA-NWB/38/2019-20]; the Basic and Applied Research in Biophysics and Material Science by the Department of Atomic Energy (DAE), Govt. of India [grant number RSI 4002 (to C.D.)]; the National Postdoctoral Fellowship from DST SERB [grant number PDF/2021/004045 (to A.B.)]; the Council of Scientific and Industrial Research (CSIR), Govt. of India, for fellowship support [grant number 31/002(1101)/2018-EMR-I (to Santanu Adhikary)]; First-time Faculty Recruitment Award from the Cancer Prevention and Research Institute of Texas (CPRIT) [grant number RR170020 (to S.S.G.)]; and the the Lizanell and Colbert Coldwell Foundation [grant number E243285 (to S.S.G.)].

ABC

ATP-binding cassette transporter

ALDH1

aldehyde dehydrogenase 1

AML

acute myeloid leukemia

ATM

ataxia telangiectasia mutation

AZA

5-azacyticine

CSC

cancer stem cell

DME

drug-metabolizing enzyme

DNMT

DNA methyltransferase

DSB

double-strand break

EMT

epithelial-to-mesenchymal transition

FDA

Food and Drug Administration

GST

glutathione S-transferase

HDAC

histone deacetylase

HDACi

HDAC inhibitor

HMT

histone methyltransferase

MDSC

myeloid-derived suppressor cell

NK

natural killer

NSCC

non-stem cancer cell

ROS

reactive oxygen species

SAHA

suberoylanilide hydroxamic acid

TAM

tumor-associated macrophage

TME

tumor microenvironment

1.
Miller
K.
,
Wang
M.
,
Gralow
J.
,
Dickler
M.
,
Cobleigh
M.
,
Perez
E.A.
et al.
(
2007
)
Paclitaxel plus bevacizumab versus paclitaxel alone for metastatic breast cancer
.
N. Engl. J. Med.
357
,
2666
2676
[PubMed]
2.
Robert
N.J.
,
Diéras
V.
,
Glaspy
J.
,
Brufsky
A.M.
,
Bondarenko
I.
,
Lipatov
O.N.
et al.
(
2011
)
RIBBON-1: randomized, double-blind, placebo-controlled, phase III trial of chemotherapy with or without bevacizumab for first-line treatment of human epidermal growth factor receptor 2-negative, locally recurrent or metastatic breast cancer
.
J. Clin. Oncol.
29
,
1252
1260
[PubMed]
3.
D’Alterio
C.
,
Scala
S.
,
Sozzi
G.
,
Roz
L.
and
Bertolini
G.
(
2020
)
Paradoxical effects of chemotherapy on tumor relapse and metastasis promotion
.
Semin Cancer Biol.
60
,
351
361
,
4.
Miller
K.D.
,
Siegel
R.L.
,
Lin
C.C.
,
Mariotto
A.B.
,
Kramer
J.L.
,
Rowland
J.H.
et al.
(
2016
)
Cancer treatment and survivorship statistics
.
CA Cancer J. Clin.
66
,
271
289
[PubMed]
5.
Easwaran
H.
,
Tsai
H.C.
and
Baylin
S.B.
(
2014
)
Cancer epigenetics: tumor heterogeneity, plasticity of stem-like states, and drug resistance
.
Mol. Cell
54
,
716
727
https://pubmed.ncbi.nlm.nih.gov/24905005/
6.
Singh
A.
and
Settleman
J.
(
2010
)
EMT, cancer stem cells and drug resistance: an emerging axis of evil in the war on cancer
.
Oncogene
29
,
4741
4751
7.
Shibue
T.
and
Weinberg
R.A.
(
2017
)
EMT, CSCs, and drug resistance: the mechanistic link and clinical implications
.
Nat. Rev. Clin. Oncol.
14
,
611
629
/pmc/articles/PMC5720366/
8.
Jin
W.
,
Liao
X.
,
Lv
Y.
,
Pang
Z.
,
Wang
Y.
,
Li
Q.
et al.
(
2017
)
MUC1 induces acquired chemoresistance by upregulating ABCB1 in EGFR-dependent manner
.
Cell Death Dis.
8
,
e2980
9.
Kim
H.K.
,
Choi
I.J.
,
Kim
C.G.
,
Kim
H.S.
,
Oshima
A.
,
Michalowski
A.
et al.
(
2011
)
A gene expression signature of acquired chemoresistance to cisplatin and fluorouracil combination chemotherapy in gastric cancer patients
.
PLoS ONE
6
,
e16694
10.
Housman
G.
,
Byler
S.
,
Heerboth
S.
,
Lapinska
K.
,
Longacre
M.
,
Snyder
N.
et al.
(
2014
)
Drug resistance in cancer: an overview
.
Cancers (Basel)
6
,
1769
1792
[PubMed]
11.
Jacobson
L.O.
,
Spurr
C.L.
,
Barron
E.S.G.
,
Smith
T.
,
Lushbaugh
C.
and
Dick
G.F.
(
1946
)
Nitrogen mustard therapy; studies on the effect of methyl-bis (beta-chloroethyl) amine hydrochloride on neoplastic diseases and allied disorders of the hemopoietic system
.
J. Am. Med. Assoc.
132
,
263
271
[PubMed]
12.
Kreso
A.
and
Dick
J.E.
(
2014
)
Evolution of the cancer stem cell model
.
Cell Stem Cell
14
,
275
291
,
https://pubmed.ncbi.nlm.nih.gov/24607403/
[PubMed]
13.
Dallas
N.A.
,
Xia
L.
,
Fan
F.
,
Gray
M.J.
,
Gaur
P.
,
Van Buren
G.
et al.
(
2009
)
Chemoresistant colorectal cancer cells, the cancer stem cell phenotype, and increased sensitivity to insulin-like growth factor-I receptor inhibition
.
Cancer Res.
69
,
1951
1957
,
https://pubmed.ncbi.nlm.nih.gov/19244128/
[PubMed]
14.
Bao
S.
,
Wu
Q.
,
McLendon
R.E.
,
Hao
Y.
,
Shi
Q.
,
Hjelmeland
A.B.
et al.
(
2006
)
Glioma stem cells promote radioresistance by preferential activation of the DNA damage response
.
Nature
444
,
756
760
,
https://pubmed.ncbi.nlm.nih.gov/17051156/
[PubMed]
15.
Li
X.
,
Lewis
M.T.
,
Huang
J.
,
Gutierrez
C.
,
Osborne
C.K.
,
Wu
M.F.
et al.
(
2008
)
Intrinsic resistance of tumorigenic breast cancer cells to chemotherapy
.
J. Natl. Cancer Inst.
100
,
672
679
[PubMed]
16.
Li
Y.
,
Wang
Z.
,
Ajani
J.A.
and
Song
S.
(
2021
)
Drug resistance and cancer stem cells
.
Cell Commun. Signal.
19
,
19
https://pubmed.ncbi.nlm.nih.gov/33588867/
17.
Phi
L.T.H.
,
Sari
I.N.
,
Yang
Y.G.
,
Lee
S.H.
,
Jun
N.
,
Kim
K.S.
et al.
(
2018
)
Cancer stem cells (CSCs) in drug resistance and their therapeutic implications in cancer treatment
.
Stem Cells Int.
https://pubmed.ncbi.nlm.nih.gov/29681949/
[PubMed]
18.
Lin
Y.
,
Dong
C.
and
Zhou
B.
(
2014
)
Epigenetic regulation of EMT: the Snail story
.
Curr. Pharm. Des.
20
,
1698
1705
[PubMed]
19.
Mani
S.A.
,
Guo
W.
,
Liao
M.J.
,
Eaton
E.N.
,
Ayyanan
A.
,
Zhou
A.Y.
et al.
(
2008
)
The epithelial-mesenchymal transition generates cells with properties of stem cells
.
Cell
133
,
704
715
https://pubmed.ncbi.nlm.nih.gov/18485877/
[PubMed]
20.
Du
B.
and
Shim
J.S.
(
2016
)
Targeting epithelial-mesenchymal transition (EMT) to overcome drug resistance in cancer
.
Molecules
21
,
965
21.
Della Corte
C.M.
,
Bellevicine
C.
,
Vicidomini
G.
,
Vitagliano
D.
,
Malapelle
U.
,
Accardo
M.
et al.
(
2015
)
SMO gene amplification and activation of the hedgehog pathway as novel mechanisms of resistance to anti-epidermal growth factor receptor drugs in human lung cancer
.
Clin. Cancer Res.
21
,
4686
4697
22.
Wu
Y.
,
Ginther
C.
,
Kim
J.
,
Mosher
N.
,
Chung
S.
,
Slamon
D.
et al.
(
2012
)
Expression of Wnt3 activates Wnt/β-catenin pathway and promotes EMT-like phenotype in trastuzumab-resistant HER2-overexpressing breast cancer cells
.
Mol. Cancer Res.
10
,
1597
1606
23.
Oshimori
N.
,
Oristian
D.
and
Fuchs
E.
(
2015
)
TGF-β promotes heterogeneity and drug resistance in squamous cell carcinoma
.
Cell
160
,
963
976
24.
Bhola
N.E.
,
Balko
J.M.
,
Dugger
T.C.
,
Kuba
M.G.
,
Sánchez
V.
,
Sanders
M.
et al.
(
2013
)
TGF-β inhibition enhances chemotherapy action against triple-negative breast cancer
.
J. Clin. Invest.
123
,
1348
1358
25.
Li
J.
,
Liu
H.
,
Yu
J.
and
Yu
H.
(
2015
)
Chemoresistance to doxorubicin induces epithelial-mesenchymal transition via upregulation of transforming growth factor β signaling in HCT116 colon cancer cells
.
Mol. Med. Rep.
12
,
192
198
26.
Uramoto
H.
,
Shimokawa
H.
,
Hanagiri
T.
,
Kuwano
M.
and
Ono
M.
(
2011
)
Expression of selected gene for acquired drug resistance to EGFR-TKI in lung adenocarcinoma
.
Lung Cancer
73
,
361
365
[PubMed]
27.
Xie
M.
,
Zhang
L.
,
He
C.S.
,
Xu
F.
,
Liu
J.L.
,
Hu
Z.H.
et al.
(
2012
)
Activation of Notch-1 enhances epithelial-mesenchymal transition in gefitinib-acquired resistant lung cancer cells
.
J. Cell. Biochem.
113
,
1501
1513
https://pubmed.ncbi.nlm.nih.gov/22173954/
[PubMed]
28.
Siebzehnrubl
F.A.
,
Silver
D.J.
,
Tugertimur
B.
,
Deleyrolle
L.P.
,
Siebzehnrubl
D.
,
Sarkisian
M.R.
et al.
(
2013
)
The ZEB1 pathway links glioblastoma initiation, invasion and chemoresistance
.
EMBO Mol. Med.
5
,
1196
1212
https://pubmed.ncbi.nlm.nih.gov/23818228/
[PubMed]
29.
Zheng
X.
,
Carstens
J.L.
,
Kim
J.
,
Scheible
M.
,
Kaye
J.
,
Sugimoto
H.
et al.
(
2015
)
Epithelial-to-mesenchymal transition is dispensable for metastasis but induces chemoresistance in pancreatic cancer
.
Nature
527
,
525
530
30.
Hirschmann-Jax
C.
,
Foster
A.E.
,
Wulf
G.G.
,
Nuchtern
J.G.
,
Jax
T.W.
,
Gobel
U.
et al.
(
2004
)
A distinct “side population” of cells with high drug efflux capacity in human tumor cells
.
Proc. Natl. Acad. Sci. U.S.A.
101
,
14228
14233
https://scholars.houstonmethodist.org/en/publications/a-distinct-side-population-of-cells-with-high-drug-efflux-capacit
31.
Haraguchi
N.
,
Utsunomiya
T.
,
Inoue
H.
,
Tanaka
F.
,
Mimori
K.
,
Barnard
G.F.
et al.
(
2006
)
Characterization of a side population of cancer cells from human gastrointestinal system
.
Stem Cells
24
,
506
513
https://pubmed.ncbi.nlm.nih.gov/16239320/
[PubMed]
32.
Moitra
K.
(
2015
)
Overcoming multidrug resistance in cancer stem cells
.
Biomed Res. Int.
,
2015
33.
Das
S.
,
Mukherjee
P.
,
Chatterjee
R.
,
Jamal
Z.
and
Chatterji
U.
(
2019
)
Enhancing chemosensitivity of breast cancer stem cells by downregulating SOX2 and ABCG2 using Wedelolactone-encapsulated Nanoparticles
.
Mol. Cancer Ther.
18
,
680
692
https://pubmed.ncbi.nlm.nih.gov/30587555/
[PubMed]
34.
Bleau
A.M.
,
Hambardzumyan
D.
,
Ozawa
T.
,
Fomchenko
E.I.
,
Huse
J.T.
,
Brennan
C.W.
et al.
(
2009
)
PTEN/PI3K/Akt pathway regulates the side population phenotype and ABCG2 activity in glioma tumor stem-like cells
.
Cell Stem Cell
4
,
226
235
https://pubmed.ncbi.nlm.nih.gov/19265662/
[PubMed]
35.
Raha
D.
,
Wilson
T.R.
,
Peng
J.
,
Peterson
D.
,
Yue
P.
,
Evangelista
M.
et al.
(
2014
)
The cancer stem cell marker aldehyde dehydrogenase is required to maintain a drug-tolerant tumor cell subpopulation
.
Am. Assoc. Cancer Res.
74
,
3579
3590
https://cancerres.aacrjournals.org/content/74/13/3579
36.
Huang
C.P.
,
Tsai
M.F.
,
Chang
T.H.
,
Tang
W.C.
,
Chen
S.Y.
,
Lai
H.H.
et al.
(
2013
)
ALDH-positive lung cancer stem cells confer resistance to epidermal growth factor receptor tyrosine kinase inhibitors
.
Cancer Lett.
328
,
144
151
https://pubmed.ncbi.nlm.nih.gov/22935675/
[PubMed]
37.
Carnero
A.
,
Garcia-Mayea
Y.
,
Mir
C.
,
Lorente
J.
,
Rubio
I.T.
and
LLeonart
M.E.
(
2016
)
The cancer stem-cell signaling network and resistance to therapy
.
Cancer Treat. Rev.
49
,
25
36
https://pubmed.ncbi.nlm.nih.gov/27434881/
[PubMed]
38.
Kleffel
S.
and
Schatton
T.
(
2013
)
Tumor dormancy and cancer stem cells: two sides of the same coin?
Adv. Exp. Med. Biol.
734
,
145
179
https://pubmed.ncbi.nlm.nih.gov/23143979/
[PubMed]
39.
Kurtova
A.V.
,
Xiao
J.
,
Mo
Q.
,
Pazhanisamy
S.
,
Krasnow
R.
,
Lerner
S.P.
et al.
(
2015
)
Blocking PGE2-induced tumour repopulation abrogates bladder cancer chemoresistance
.
Nature
517
,
209
213
https://pubmed.ncbi.nlm.nih.gov/25470039/
[PubMed]
40.
Domingo-Domenech
J.
,
Vidal
S.J.
,
Rodriguez-Bravo
V.
,
Castillo-Martin
M.
,
Quinn
S.A.
,
Rodriguez-Barrueco
R.
et al.
(
2012
)
Suppression of acquired docetaxel resistance in prostate cancer through depletion of notch- and hedgehog-dependent tumor-initiating cells
.
Cancer Cell
22
,
373
388
https://pubmed.ncbi.nlm.nih.gov/22975379/
[PubMed]
41.
Shetzer
Y.
,
Solomon
H.
,
Koifman
G.
,
Molchadsky
A.
,
Horesh
S.
and
Rotter
V.
(
2014
)
The paradigm of mutant p53-expressing cancer stem cells and drug resistance
.
Carcinogenesis
35
,
1196
1208
https://pubmed.ncbi.nlm.nih.gov/24658181/
[PubMed]
42.
Diehn
M.
,
Cho
R.W.
,
Lobo
N.A.
,
Kalisky
T.
,
Dorie
M.J.
,
Kulp
A.N.
et al.
(
2009
)
Association of reactive oxygen species levels and radioresistance in cancer stem cells
.
Nature
458
,
780
783
https://pubmed.ncbi.nlm.nih.gov/19194462/
[PubMed]
43.
Sharma
S.V.
,
Lee
D.Y.
,
Li
B.
,
Quinlan
M.P.
,
Takahashi
F.
,
Maheswaran
S.
et al.
(
2010
)
A chromatin-mediated reversible drug-tolerant state in cancer cell subpopulations
.
Cell
141
,
69
80
https://pubmed.ncbi.nlm.nih.gov/20371346/
[PubMed]
44.
Liau
B.B.
,
Sievers
C.
,
Donohue
L.K.
,
Gillespie
S.M.
,
Flavahan
W.A.
,
Miller
T.E.
et al.
(
2017
)
Adaptive chromatin remodeling drives glioblastoma stem cell plasticity and drug tolerance
.
NIH Public Access
20
,
233
246
pmc/articles/PMC5291795/
45.
Bell
C.C.
and
Gilan
O.
(
2019
)
Principles and mechanisms of non-genetic resistance in cancer
.
Nature
122
,
465
472
,
https://www.nature.com/articles/s41416-019-0648-6
46.
Toh
T.B.
,
Lim
J.J.
and
Chow
E.K.H.
(
2017
)
Epigenetics in cancer stem cells
.
Mol. Cancer
16
,
29
29
,
https://europepmc.org/articles/PMC5286794
47.
Vasiliou
V.
,
Vasiliou
K.
and
Nebert
D.W.
(
2009
)
Human ATP-binding cassette (ABC) transporter family
.
Hum. Genomics
3
,
281
290
[PubMed]
48.
Cho
Y.
and
Kim
Y.K.
(
2020
)
Cancer stem cells as a potential target to overcome multidrug resistance
.
Front. Oncol.
10
,
764
49.
Bhattacharya
A.
,
Mukherjee
S.
,
Khan
P.
,
Banerjee
S.
,
Dutta
A.
,
Banerjee
N.
et al.
(
2020
)
SMAR1 repression by pluripotency factors and consequent chemoresistance in breast cancer stem-like cells is reversed by aspirin
.
Sci. Signal.
13
,
[PubMed]
50.
Zappe
K.
and
Cichna-Markl
M.
(
2020
)
Aberrant DNA methylation of ABC transporters in cancer
.
Cells
9
,
2281
[PubMed]
51.
Wang
X.K.
,
He
J.H.
,
Xu
J.H.
,
Ye
S.
,
Wang
F.
,
Zhang
H.
et al.
(
2014
)
Afatinib enhances the efficacy of conventional chemotherapeutic agents by eradicating cancer stem-like cells
.
Cancer Res.
74
,
4431
4445
[PubMed]
52.
Martín
V.
,
Sanchez-Sanchez
A.M.
,
Herrera
F.
,
Gomez-Manzano
C.
,
Fueyo
J.
,
Alvarez-Vega
M.A.
et al.
(
2013
)
Melatonin-induced methylation of the ABCG2/BCRP promoter as a novel mechanism to overcome multidrug resistance in brain tumour stem cells
.
Br. J. Cancer
108
,
2005
2012
[PubMed]
53.
Bram
E.E.
,
Stark
M.
,
Raz
S.
and
Assaraf
Y.G.
(
2009
)
Chemotherapeutic drug-induced ABCG2 promoter demethylation as a novel mechanism of acquired multidrug resistance
.
Neoplasia
11
,
1359
1370
https://pubmed.ncbi.nlm.nih.gov/20019844/
[PubMed]
54.
Ferreri
A.J.M.
,
Dell'Oro
S.
,
Capello
D.
,
Ponzoni
M.
,
Iuzzolino
P.
,
Rossi
D.
et al.
(
2004
)
Aberrant methylation in the promoter region of the reduced folate carrier gene is a potential mechanism of resistance to methotrexate in primary central nervous system lymphomas
.
Br. J. Haematol.
126
,
657
664
,
https://pubmed.ncbi.nlm.nih.gov/15327516/
[PubMed]
55.
Peng
L.
and
Zhong
X.
(
2015
)
Epigenetic regulation of drug metabolism and transport
.
Acta Pharm. Sin. B.
5
,
106
112
56.
Mukherjee
S.
,
Adhikary
S.
,
Gadad
S.S.
,
Mondal
P.
,
Sen
S.
,
Choudhari
R.
et al.
(
2020
)
Suppression of poised oncogenes by ZMYND8 promotes chemo-sensitization
.
Nature
11
,
1
18
57.
Wang
J.
,
Yu
L.
,
Jiang
H.
,
Zheng
X.
and
Zeng
S.
(
2020
)
Epigenetic regulation of differentially expressed drug-metabolizing enzymes in cancer
.
Drug Metab. Dispos.
48
,
759
768
[PubMed]
58.
Moreb
J.
(
2008
)
Aldehyde dehydrogenase as a marker for stem cells
.
Curr. Stem Cell Res. Ther.
3
,
237
246
[PubMed]
59.
Honoki
K.
,
Fujii
H.
,
Kubo
A.
,
Kido
A.
,
Mori
T.
,
Tanaka
Y.
et al.
(
2010
)
Possible involvement of stem-like populations with elevated ALDH1 in sarcomas for chemotherapeutic drug resistance
.
Oncol. Rep.
24
,
501
505
[PubMed]
60.
Peitzsch
C.
,
Cojoc
M.
,
Hein
L.
,
Kurth
I.
,
Mäbert
K.
,
Trautmann
F.
et al.
(
2016
)
An epigenetic reprogramming strategy to resensitize radioresistant prostate cancer cells
.
Cancer Res.
76
,
2637
2651
[PubMed]
61.
Habano
W.
,
Kawamura
K.
,
Iizuka
N.
,
Terashima
J.
,
Sugai
T.
and
Ozawa
S.
(
2015
)
Analysis of DNA methylation landscape reveals the roles of DNA methylation in the regulation of drug metabolizing enzymes
.
Verlag
7
,
1
11
https://clinicalepigeneticsjournal.biomedcentral.com/articles/10.1186/s13148-015-0136-7
62.
Luo
W.
,
Karpf
A.R.
,
Deeb
K.K.
,
Muindi
J.R.
,
Morrison
C.D.
,
Johnson
C.S.
et al.
(
2010
)
Epigenetic regulation of vitamin D 24-hydroxylase/CYP24A1 in human prostate cancer
.
Cancer Res.
70
,
5953
5962
,
https://pubmed.ncbi.nlm.nih.gov/20587525/
[PubMed]
63.
Widschwendter
M.
,
Siegmund
K.D.
,
Müller
H.M.
,
Fiegl
H.
,
Marth
C.
,
Müller-Holzner
E.
et al.
(
2004
)
Association of breast cancer DNA methylation profiles with hormone receptor status and response to tamoxifen
.
Cancer Res.
64
,
3807
3813
https://cancerres.aacrjournals.org/content/64/11/3807
64.
Kim
S.J.
,
Kang
H.S.
,
Jung
S.Y.
,
Min
S.Y.
,
Lee
S.
,
Kim
S.W.
et al.
(
2010
)
Methylation patterns of genes coding for drug-metabolizing enzymes in tamoxifen-resistant breast cancer tissues
.
J. Mol. Med.
88
,
1123
1131
https://snucm.elsevierpure.com/en/publications/methylation-patterns-of-genes-coding-for-drug-metabolizing-enzyme
65.
Huang
R.X.
and
Zhou
P.K.
(
2020
)
DNA damage response signaling pathways and targets for radiotherapy sensitization in cancer
.
Signal. Transduct. Target Ther.
5
,
60
https://pubmed.ncbi.nlm.nih.gov/32355263/
66.
Nair
N.
,
Shoaib
M.
and
Sørensen
C.S.
(
2017
)
Chromatin dynamics in genome stability: roles in suppressing endogenous DNA damage and facilitating DNA repair
.
Int. J. Mol. Sci.
18
,
1486
https://pubmed.ncbi.nlm.nih.gov/28698521/
67.
Stadler
J.
and
Richly
H.
(
2017
)
Regulation of DNA repair mechanisms: how the chromatin environment regulates the DNA damage response
.
Int. J. Mol. Sci.
18
,
1715
https://pubmed.ncbi.nlm.nih.gov/28783053/
[PubMed]
68.
Cann
K.L.
and
Dellaire
G.
(
2011
)
Heterochromatin and the DNA damage response: the need to relax
.
Biochem. Cell Biol.
89
,
45
60
[PubMed]
69.
Takata
H.
,
Hanafusa
T.
,
Mori
T.
,
Shimura
M.
,
Iida
Y.
,
Ishikawa
K.
et al.
(
2013
)
Chromatin compaction protects genomic DNA from radiation damage
.
PLoS ONE
8
,
e75622
70.
Schuster-Böckler
B.
and
Lehner
B.
(
2012
)
Chromatin organization is a major influence on regional mutation rates in human cancer cells
.
Nature
488
,
504
507
[PubMed]
71.
Bouwman
P.
and
Jonkers
J.
(
2012
)
The effects of deregulated DNA damage signalling on cancer chemotherapy response and resistance
.
Nat. Rev. Cancer
12
,
587
598
[PubMed]
72.
Kandoth
C.
,
McLellan
M.D.
,
Vandin
F.
,
Ye
K.
,
Niu
B.
,
Lu
C.
et al.
(
2013
)
Mutational landscape and significance across 12 major cancer types
.
Nature
502
,
333
339
[PubMed]
73.
Plass
C.
,
Pfister
S.M.
,
Lindroth
A.M.
,
Bogatyrova
O.
,
Claus
R.
and
Lichter
P.
(
2013
)
Mutations in regulators of the epigenome and their connections to global chromatin patterns in cancer
.
Nat. Rev. Genet.
14
,
765
780
[PubMed]
74.
Goodarzi
A.A.
,
Noon
A.T.
,
Deckbar
D.
,
Ziv
Y.
,
Shiloh
Y.
,
Löbrich
M.
et al.
(
2008
)
ATM signaling facilitates repair of DNA double-strand breaks associated with heterochromatin
.
Mol. Cell
31
,
167
177
[PubMed]
75.
Choi
M.
,
Kipps
T.
and
Kurzrock
R.
(
2016
)
ATM mutations in cancer: therapeutic implications
.
Mol. Cancer Ther.
15
,
1781
1791
[PubMed]
76.
Herman
J.G.
,
Umar
A.
,
Polyak
K.
,
Graff
J.R.
,
Ahuja
N.
,
Issa
J.P.J.
et al.
(
1998
)
Incidence and functional consequences of hMLH1 promoter hypermethylation in colorectal carcinoma
.
Proc. Natl. Acad. Sci. U.S.A.
95
,
6870
6875
[PubMed]
77.
Mossman
D.
,
Kim
K.T.
and
Scott
R.J.
(
2010
)
Demethylation by 5-aza-2’-deoxycytidine in colorectal cancer cells targets genomic DNA whilst promoter CpG island methylation persists
.
BMC Cancer
10
,
366
[PubMed]
78.
Steinbichler
T.B.
,
Dudás
J.
,
Skvortsov
S.
,
Ganswindt
U.
,
Riechelmann
H.
and
Skvortsova
I.I.
(
2018
)
Therapy resistance mediated by cancer stem cells
.
Semin. Cancer Biol.
53
,
156
167
[PubMed]
79.
Das
P.K.
,
Islam
F.
and
Lam
A.K.
(
2020
)
The roles of cancer stem cells and therapy resistance in colorectal carcinoma
.
Cells
9
,
1392
,
80.
Pistritto
G.
,
Trisciuoglio
D.
,
Ceci
C.
,
Alessia
G.
and
D'Orazi
G.
(
2016
)
Apoptosis as anticancer mechanism: function and dysfunction of its modulators and targeted therapeutic strategies
.
Aging (Albany N.Y.)
8
,
603
619
https://pubmed.ncbi.nlm.nih.gov/27019364/
[PubMed]
81.
Hanahan
D.
and
Weinberg
R.A.
(
2011
)
Hallmarks of cancer: the next generation
.
Cell
144
,
646
674
https://pubmed.ncbi.nlm.nih.gov/21376230/
[PubMed]
82.
Pfeffer
C.M.
and
Singh
A.T.K.
(
2018
)
Apoptosis: a target for anticancer therapy
.
Int. J. Mol. Sci.
19
,
448
https://pubmed.ncbi.nlm.nih.gov/29393886/
83.
Ozyerli-Goknar
E.
and
Bagci-Onder
T.
(
2021
)
Epigenetic deregulation of apoptosis in cancers
.
Cancers (Basel)
13
,
3210
[PubMed]
84.
Fernald
K.
and
Kurokawa
M.
(
2013
)
Evading apoptosis in cancer
.
Trends Cell Biol.
23
,
620
633
[PubMed]
85.
Vardabasso
C.
,
Hasson
D.
,
Ratnakumar
K.
,
Chung
C.Y.
,
Duarte
L.F.
and
Bernstein
E.
(
2014
)
Histone variants: emerging players in cancer biology
.
Cell. Mol. Life Sci.
71
,
379
404
[PubMed]
86.
Yang
B.
,
Tong
R.
,
Liu
H.
,
Wu
J.
,
Chen
D.
,
Xue
Z.
et al.
(
2018
)
H2A.Z regulates tumorigenesis, metastasis and sensitivity to cisplatin in intrahepatic cholangiocarcinoma
.
Int. J. Oncol.
52
,
1235
1245
[PubMed]
87.
Th'ng
J.P.H.
(
2001
)
Histone modifications and apoptosis: cause or consequence?
Biochem. Cell Biol.
79
,
305
311
[PubMed]
88.
Lu
J.
,
Feng
Y.
,
Wang
X.
,
Xu
L.
,
Pan
H.
,
Zhu
S.
et al.
(
2009
)
The transcription factor ZBP-89 suppresses p16 expression through a histone modification mechanism to affect cell senescence
.
FEBS J.
276
,
4197
4206
[PubMed]
89.
Yang
X.
,
Karuturi
R.K.M.
,
Sun
F.
,
Aau
M.
,
Yu
K.
,
Shao
R.
et al.
(
2009
)
CDKN1C (p57) is a direct target of EZH2 and suppressed by multiple epigenetic mechanisms in breast cancer cells
.
PLoS ONE
4
,
e5011
90.
Richon
V.M.
,
Sandhoff
T.W.
,
Rifkind
R.A.
and
Marks
P.A.
(
2000
)
Histone deacetylase inhibitor selectively induces p21WAF1 expression and gene-associated histone acetylation
.
Proc. Natl. Acad. Sci. U.S.A.
97
,
10014
10019
91.
Stucki
M.
(
2009
)
Histone H2A.X Tyr142 phosphorylation: a novel sWItCH for apoptosis?
DNA Rep. (Amst.)
8
,
873
876
[PubMed]
92.
Stucki
M.
,
Clapperton
J.A.
,
Mohammad
D.
,
Yaffe
M.B.
,
Smerdon
S.J.
and
Jackson
S.P.
(
2005
)
MDC1 directly binds phosphorylated histone H2AX to regulate cellular responses to DNA double-strand breaks
.
Cell
123
,
1213
1226
[PubMed]
93.
Ehrlich
M.
(
2009
)
DNA hypomethylation in cancer cells
.
Epigenomics
1
,
239
94.
Wang
Y.
,
Yu
Q.
,
Cho
A.H.
,
Rondeau
G.
,
Welsh
J.
,
Adamson
E.
et al.
(
2005
)
Survey of differentially methylated promoters in prostate cancer cell lines
.
Neoplasia
7
,
748
760
[PubMed]
95.
Murphy
T.M.
,
Sullivan
L.
,
Lane
C.
,
O'Connor
L.
,
Barrett
C.
,
Hollywood
D.
et al.
(
2011
)
In silico analysis and DHPLC screening strategy identifies novel apoptotic gene targets of aberrant promoter hypermethylation in prostate cancer
.
Prostate
71
,
1
17
[PubMed]
96.
Pompeia
C.
,
Hodge
D.R.
,
Plass
C.
,
Wu
Y.Z.
,
Marquez
V.E.
,
Kelley
J.A.
et al.
(
2004
)
Microarray analysis of epigenetic silencing of gene expression in the KAS-6/1 multiple myeloma cell line
.
Cancer Res.
64
,
3465
3473
[PubMed]
97.
San José-Eneriz
E.
,
Agirre
X.
,
Jiménez-Velasco
A.
,
Cordeu
L.
,
Martín
V.
,
Arqueros
V.
et al.
(
2009
)
Epigenetic down-regulation of BIM expression is associated with reduced optimal responses to imatinib treatment in chronic myeloid leukaemia
.
Eur. J. Cancer
45
,
1877
1889
[PubMed]
98.
Malekzadeh
K.
,
Sobti
R.C.
,
Nikbakht
M.
,
Shekari
M.
,
Hosseini
S.A.
,
Tamandani
D.K.
et al.
(
2009
)
Methylation patterns of Rb1 and Casp-8 promoters and their impact on their expression in bladder cancer
.
Cancer Invest.
27
,
70
80
[PubMed]
99.
Cho
S.
,
Lee
J.H.
,
Cho
S.B.
,
Yoon
K.W.
,
Park
S.Y.
,
Lee
W.S.
et al.
(
2010
)
Epigenetic methylation and expression of caspase 8 and survivin in hepatocellular carcinoma
.
Pathol. Int.
60
,
203
211
[PubMed]
100.
Hervouet
E.
,
Vallette
F.M.
and
Cartron
P.F.
(
2010
)
Impact of the DNA methyltransferases expression on the methylation status of apoptosis-associated genes in glioblastoma multiforme
.
Cell Death Dis.
1
,
e8
[PubMed]
101.
Shivapurkar
N.
,
Toyooka
S.
,
Eby
M.T.
,
Huang
C.X.
,
Sathyanarayana
U.G.
,
Cunningham
H.T.
et al.
(
2002
)
Differential inactivation of caspase-8 in lung cancers
.
Cancer Biol. Ther.
1
,
65
69
[PubMed]
102.
Harada
K.
,
Toyooka
S.
,
Shivapurkar
N.
,
Maitra
A.
,
Reddy
J.L.
,
Matta
H.
et al.
(
2002
)
Deregulation of caspase 8 and 10 expression in pediatric tumors and cell lines
.
Cancer Res.
62
,
5897
5901
[PubMed]
103.
Xu
J.D.
,
Cao
X.X.
,
Long
Z.W.
,
Liu
X.P.
,
Furuya
T.
,
Xu
J.W.
et al.
(
2011
)
BCL2L10 protein regulates apoptosis/proliferation through differential pathways in gastric cancer cells
.
J Pathol.
223
,
400
409
104.
Sugita
H.
,
Iida
S.
,
Inokuchi
M.
,
Kato
K.
,
Ishiguro
M.
,
Ishikawa
T.
et al.
(
2011
)
Methylation of BNIP3 and DAPK indicates lower response to chemotherapy and poor prognosis in gastric cancer
.
Oncol. Rep.
25
,
513
518
[PubMed]
105.
Obata
T.
,
Toyota
M.
,
Satoh
A.
,
Sasaki
Y.
,
Ogi
K.
,
Akino
K.
et al.
(
2003
)
Identification of HRK as a target of epigenetic inactivation in colorectal and gastric cancer
.
Clin. Cancer Res.
9
,
6410
6418
[PubMed]
106.
Agirre
X.
,
Novo
F.J.
,
Calasanz
M.J.
,
Larráyoz
M.J.
,
Lahortiga
I.
,
Valgañón
M.
et al.
(
2003
)
TP53 is frequently altered by methylation, mutation, and/or deletion in acute lymphoblastic leukaemia
.
Mol. Carcinog.
38
,
201
208
[PubMed]
107.
Wen
Y.C.
,
Wang
D.H.
,
RayWhay
C.Y.
,
Luo
J.
,
Gu
W.
and
Baylin
S.B.
(
2005
)
Tumor suppressor HIC1 directly regulates SIRT1 to modulate p53-dependent DNA-damage responses
.
Cell
123
,
437
448
[PubMed]
108.
Esteller
M.
(
2007
)
Cancer epigenomics: DNA methylomes and histone-modification maps
.
Nat. Rev. Genet.
8
,
286
298
[PubMed]
109.
Furukawa
Y.
,
Sutheesophon
K.
,
Wada
T.
,
Nishimura
M.
,
Saifo
Y.
,
Ishii
H.
et al.
(
2005
)
Methylation silencing of the Apaf-1 gene in acute leukemia
.
Mol. Cancer Res.
3
,
325
334
[PubMed]
110.
Wu
J.
and
Wood
G.S.
(
2011
)
Reduction of Fas/CD95 promoter methylation, upregulation of Fas protein, and enhancement of sensitivity to apoptosis in cutaneous T-cell lymphoma
.
Arch. Dermatol.
147
,
443
449
[PubMed]
111.
Navarro
F.
and
Lieberman
J.
(
2015
)
miR-34 and p53: new insights into a complex functional relationship
.
PLoS ONE
10
,
e0132767
https://pubmed.ncbi.nlm.nih.gov/26177460/
112.
Ji
Q.
,
Hao
X.
,
Meng
Y.
,
Zhang
M.
,
DeSano
J.
,
Fan
D.
et al.
(
2008
)
Restoration of tumor suppressor miR-34 inhibits human p53-mutant gastric cancer tumorspheres
.
BMC Cancer
8
,
266
[PubMed]
113.
Lodygin
D.
,
Tarasov
V.
,
Epanchintsev
A.
,
Berking
C.
,
Knyazeva
T.
,
Körner
H.
et al.
(
2008
)
Inactivation of miR-34a by aberrant CpG methylation in multiple types of cancer
.
Cell Cycle
7
,
2591
2600
[PubMed]
114.
Sampath
D.
,
Liu
C.
,
Vasan
K.
,
Sulda
M.
,
Puduvalli
V.K.
,
Wierda
W.G.
et al.
(
2012
)
Histone deacetylases mediate the silencing of miR-15a, miR-16, and miR-29b in chronic lymphocytic leukemia
.
Blood
119
,
1162
1172
[PubMed]
115.
Ratert
N.
,
Meyer
H.A.
,
Jung
M.
,
Mollenkopf
H.J.
,
Wagner
I.
,
Miller
K.
et al.
(
2012
)
Reference miRNAs for miRNAome analysis of urothelial carcinomas
.
PLoS ONE
7
,
e39309
[PubMed]
116.
Li
Y.
,
Gao
L.
,
Luo
X.
,
Wang
L.
,
Gao
X.
,
Wang
W.
et al.
(
2013
)
Epigenetic silencing of microRNA-193a contributes to leukemogenesis in t(8; 21) acute myeloid leukemia by activating the PTEN/PI3K signal pathway
.
Blood
121
,
499
509
[PubMed]
117.
Saito
Y.
,
Suzuki
H.
,
Tsugawa
H.
,
Nakagawa
I.
,
Matsuzaki
J.
,
Kanai
Y.
et al.
(
2009
)
Chromatin remodeling at Alu repeats by epigenetic treatment activates silenced microRNA-512-5p with downregulation of Mcl-1 in human gastric cancer cells
.
Oncogene
28
,
2738
2744
[PubMed]
118.
Crawford
M.
,
Batte
K.
,
Yu
L.
,
Wu
X.
,
Nuovo
G.J.
,
Marsh
C.B.
et al.
(
2009
)
MicroRNA 133B targets pro-survival molecules MCL-1 and BCL2L2 in lung cancer
.
Biochem. Biophys. Res. Commun.
388
,
483
489
[PubMed]
119.
Ghasemi
A.
,
Fallah
S.
and
Ansari
M.
(
2016
)
MiR-153 as a tumor suppressor in glioblastoma multiforme is downregulated by DNA methylation
.
Clin. Lab.
62
,
573
580
[PubMed]
120.
Garofalo
M.
,
Di Leva
G.
,
Romano
G.
,
Nuovo
G.
,
Suh
S.S.
,
Ngankeu
A.
et al.
(
2009
)
miR-221&222 regulate TRAIL resistance and enhance tumorigenicity through PTEN and TIMP3 downregulation
.
Cancer Cell
16
,
498
509
[PubMed]
121.
Zhang
C.Z.
,
Zhang
J.X.
,
Zhang
A.L.
,
Shi
Z.D.
,
Han
L.
,
Jia
Z.F.
et al.
(
2010
)
MiR-221 and miR-222 target PUMA to induce cell survival in glioblastoma
.
Mol. Cancer
9
,
229
122.
Jin
X.
,
Cai
L.
,
Wang
C.
,
Deng
X.
,
Yi
S.
,
Lei
Z.
et al.
(
2018
)
CASC2/miR-24/miR-221 modulates the TRAIL resistance of hepatocellular carcinoma cell through caspase-8/caspase-3
.
Cell Death Dis.
9
,
318
123.
Lu
Q.
,
Lu
C.
,
Zhou
G.P.
,
Zhang
W.
,
Xiao
H.
and
Wang
X.R.
(
2010
)
MicroRNA-221 silencing predisposed human bladder cancer cells to undergo apoptosis induced by TRAIL
.
Urol. Oncol.
28
,
635
641
[PubMed]
124.
Choy
E.Y.W.
,
Siu
K.L.
,
Kok
K.H.
,
Lung
R.W.M.
,
Tsang
C.M.
,
To
K.F.
et al.
(
2008
)
An Epstein-Barr virus-encoded microRNA targets PUMA to promote host cell survival
.
J. Exp. Med.
205
,
2551
2560
[PubMed]
125.
Duan
S.
,
Dong
X.
,
Hai
J.
,
Jiang
J.
,
Wang
W.
,
Yang
J.
et al.
(
2018
)
MicroRNA-135a-3p is downregulated and serves as a tumour suppressor in ovarian cancer by targeting CCR2
.
Biomed. Pharmacother.
107
,
712
720
[PubMed]
126.
Xu
H.
and
Wen
Q.
(
2018
)
Downregulation of miR-135a predicts poor prognosis in acute myeloid leukemia and regulates leukemia progression via modulating HOXA10 expression
.
Mol. Med. Rep.
18
,
1134
1140
[PubMed]
127.
Hou
P.
,
Kapoor
A.
,
Zhang
Q.
,
Li
J.
,
Wu
C.J.
,
Li
J.
et al.
(
2020
)
Tumor microenvironment remodeling enables bypass of oncogenic KRAS dependency in pancreatic cancer
.
Cancer Discov.
10
,
1058
1077
[PubMed]
128.
Chang
Y.C.
,
Chen
T.C.
,
Lee
C.T.
,
Yang
C.Y.
,
Wang
H.W.
,
Wang
C.C.
et al.
(
2008
)
Epigenetic control of MHC class II expression in tumor-associated macrophages by decoy receptor 3
.
Blood
111
,
5054
5063
[PubMed]
129.
Gabrilovich
D.I.
and
Nagaraj
S.
(
2009
)
Myeloid-derived suppressor cells as regulators of the immune system [Internet]
.
Nat. Rev. Immunol.
9
,
162
174
https://pubmed.ncbi.nlm.nih.gov/19197294/
[PubMed]
130.
Zhou
J.
,
Liu
M.
,
Sun
H.
,
Feng
Y.
,
Xu
L.
,
Chan
A.W.H.
et al.
(
2018
)
Hepatoma-intrinsic CCRK inhibition diminishes myeloid-derived suppressor cell immunosuppression and enhances immune-checkpoint blockade efficacy
.
Gut
67
,
931
944
[PubMed]
131.
Burr
M.L.
,
Sparbier
C.E.
,
Chan
K.L.
,
Chan
Y.C.
,
Kersbergen
A.
,
Lam
E.Y.N.
et al.
(
2019
)
An evolutionarily conserved function of polycomb silences the MHC class I antigen presentation pathway and enables immune evasion in cancer
.
Cancer Cell
36
,
385.e8
401.e8
[PubMed]
132.
Peng
D.
,
Kryczek
I.
,
Nagarsheth
N.
,
Zhao
L.
,
Wei
S.
,
Wang
W.
et al.
(
2015
)
Epigenetic silencing of TH1-type chemokines shapes tumour immunity and immunotherapy
.
Nature
527
,
249
253
[PubMed]
133.
Xu
H.
,
Cheung
I.Y.
,
Guo
H.F.
and
Cheung
N.K.V.
(
2009
)
MicroRNA miR-29 modulates expression of immunoinhibitory molecule B7-H3: potential implications for immune based therapy of human solid tumors
.
Cancer Res.
69
,
6275
6281
[PubMed]
134.
Kim
T.D.
,
Lee
S.U.
,
Yun
S.
,
Sun
H.N.
,
Lee
S.H.
,
Kim
J.W.
et al.
(
2011
)
Human microRNA-27a* targets Prf1 and GzmB expression to regulate NK-cell cytotoxicity
.
Blood
118
,
5476
5486
[PubMed]
135.
Hanson
J.A.
,
Gillespie
J.W.
,
Grover
A.
,
Tangrea
M.A.
,
Chuaqui
R.F.
,
Emmert-Buck
M.R.
et al.
(
2006
)
Gene promoter methylation in prostate tumor-associated stromal cells
.
J. Natl. Cancer Inst.
98
,
255
261
[PubMed]
136.
Justus
C.R.
,
Sanderlin
E.J.
and
Yang
L.V.
(
2015
)
Molecular connections between cancer cell metabolism and the tumor microenvironment
.
Int. J. Mol. Sci.
16
,
11055
11086
[PubMed]
137.
Najafi
M.
,
Farhood
B.
and
Mortezaee
K.
(
2019
)
Cancer stem cells (CSCs) in cancer progression and therapy
.
J. Cell. Physiol.
234
,
8381
8395
[PubMed]
138.
Prieto-Vila
M.
,
Takahashi
R.U.
,
Usuba
W.
,
Kohama
I.
and
Ochiya
T.
(
2017
)
Drug resistance driven by cancer stem cells and their niche
.
Int. J. Mol. Sci.
18
,
2674
[PubMed]
139.
Banerjee
S.
,
Halder
K.
,
Bose
A.
,
Bhattacharya
P.
,
Gupta
G.
,
Karmahapatra
S.
et al.
(
2011
)
TLR signaling-mediated differential histone modification at IL-10 and IL-12 promoter region leads to functional impairments in tumor-associated macrophages
.
Carcinogenesis
32
,
1789
1797
[PubMed]
140.
Pan
W.
,
Zhu
S.
,
Qu
K.
,
Meeth
K.
,
Cheng
J.
,
He
K.
et al.
(
2017
)
The DNA methylcytosine dioxygenase Tet2 sustains immunosuppressive function of tumor-infiltrating myeloid cells to promote melanoma progression
.
Immunity
47
,
284.e5
297.e5
141.
Boulding
T.
,
McCuaig
R.D.
,
Tan
A.
,
Hardy
K.
,
Wu
F.
,
Dunn
J.
et al.
(
2018
)
LSD1 activation promotes inducible EMT programs and modulates the tumour microenvironment in breast cancer
.
Sci. Rep.
8
,
73
[PubMed]
142.
Cooks
T.
,
Pateras
I.S.
,
Jenkins
L.M.
,
Patel
K.M.
,
Robles
A.I.
,
Morris
J.
et al.
(
2018
)
Mutant p53 cancers reprogram macrophages to tumor supporting macrophages via exosomal miR-1246
.
Nat. Commun.
9
,
771
[PubMed]
143.
Larionova
I.
,
Cherdyntseva
N.
,
Liu
T.
,
Patysheva
M.
,
Rakina
M.
and
Kzhyshkowska
J.
(
2019
)
Interaction of tumor-associated macrophages and cancer chemotherapy
.
Oncoimmunology
8
,
[PubMed]
144.
Quagliano
A.
,
Gopalakrishnapillai
A.
and
Barwe
S.P.
(
2017
)
Epigenetic drug combination overcomes osteoblast-induced chemoprotection in pediatric acute lymphoid leukemia
.
Leuk. Res.
56
,
36
43
[PubMed]
145.
Cheng
J.
,
Kang
X.
,
Zhang
S.
and
Yeh
E.T.H.
(
2007
)
SUMO-specific protease 1 is essential for stabilization of HIF1alpha during hypoxia
.
Cell
131
,
584
595
[PubMed]
146.
Kietzmann
T.
,
Mennerich
D.
and
Dimova
E.Y.
(
2016
)
Hypoxia-inducible factors (HIFs) and phosphorylation: impact on stability, localization, and transactivity
.
Front. Cell Dev. Biol.
4
,
11
[PubMed]
147.
Pedrosa
L.
,
Esposito
F.
,
Thomson
T.M.
and
Maurel
J.
(
2019
)
The tumor microenvironment in colorectal cancer therapy
.
Cancers (Basel)
11
,
1172
[PubMed]
148.
Roma-Rodrigues
C.
,
Mendes
R.
,
Baptista
P.V.
and
Fernandes
A.R.
(
2019
)
Targeting tumor microenvironment for cancer therapy
.
Int. J. Mol. Sci.
20
,
840
[PubMed]
149.
Chakraborty
C.
,
Sharma
A.R.
,
Sharma
G.
and
Lee
S.S.
(
2020
)
The interplay among miRNAs, major cytokines, and cancer-related inflammation
.
Mol. Ther. Nucleic Acids
20
,
606
620
[PubMed]
150.
Xia
L.
,
Zhang
D.
,
Du
R.
,
Pan
Y.
,
Zhao
L.
,
Sun
S.
et al.
(
2008
)
miR-15b and miR-16 modulate multidrug resistance by targeting BCL2 in human gastric cancer cells
.
Int. J. Cancer
123
,
372
379
[PubMed]
151.
Dejean
E.
,
Renalier
M.H.
,
Foisseau
M.
,
Agirre
X.
,
Joseph
N.
,
De Paiva
G.R.
et al.
(
2011
)
Hypoxia-microRNA-16 downregulation induces VEGF expression in anaplastic lymphoma kinase (ALK)-positive anaplastic large-cell lymphomas
.
Leukemia
25
,
1882
1890
[PubMed]
152.
Lin
S.C.
,
Wang
C.C.
,
Wu
M.H.
,
Yang
S.H.
,
Li
Y.H.
and
Tsai
S.J.
(
2012
)
Hypoxia-induced microRNA-20a expression increases ERK phosphorylation and angiogenic gene expression in endometriotic stromal cells
.
J. Clin. Endocrinol. Metab.
97
,
E1515
E1523
153.
Chai
H.
,
Liu
M.
,
Tian
R.
,
Li
X.
and
Tang
H.
(
2011
)
miR-20a targets BNIP2 and contributes chemotherapeutic resistance in colorectal adenocarcinoma SW480 and SW620 cell lines
.
Acta Biochim. Biophys. Sin. (Shanghai)
43
,
217
225
[PubMed]
154.
Meng
F.
,
Henson
R.
,
Lang
M.
,
Wehbe
H.
,
Maheshwari
S.
,
Mendell
J.T.
et al.
(
2006
)
Involvement of human micro-RNA in growth and response to chemotherapy in human cholangiocarcinoma cell lines
.
Gastroenterology
130
,
2113
2129
[PubMed]
155.
Carrick
S.
,
Parker
S.
,
Thornton
C.E.
,
Ghersi
D.
,
Simes
J.
and
Wilcken
N.
(
2009
)
Single agent versus combination chemotherapy for metastatic breast cancer
.
Cochrane Database Syst. Rev.
2009
,
CD003372
156.
Albain
K.S.
,
Nag
S.M.
,
Calderillo-Ruiz
G.
,
Jordaan
J.P.
,
Llombart
A.C.
,
Pluzanska
A.
et al.
(
2008
)
Gemcitabine plus Paclitaxel versus Paclitaxel monotherapy in patients with metastatic breast cancer and prior anthracycline treatment
.
J. Clin. Oncol.
26
,
3950
3957
[PubMed]
157.
Mounier
N.
,
Briere
J.
,
Gisselbrecht
C.
,
Emile
J.F.
,
Lederlin
P.
,
Sebban
C.
et al.
(
2003
)
Rituximab plus CHOP (R-CHOP) overcomes bcl-2–associated resistance to chemotherapy in elderly patients with diffuse large B-cell lymphoma (DLBCL)
.
Blood
101
,
4279
4284
[PubMed]
158.
Kim
A.
,
Ueda
Y.
,
Naka
T.
and
Enomoto
T.
(
2012
)
Therapeutic strategies in epithelial ovarian cancer
.
J. Exp. Clin. Cancer Res.
31
,
14
[PubMed]
159.
Blanquicett
C.
,
Roman
J.
and
Hart
C.M.
(
2008
)
Thiazolidinediones as anti-cancer agents
.
Cancer Ther.
6
,
25
34
[PubMed]
160.
Samuel
S.M.
,
Varghese
E.
,
Koklesová
L.
,
Líšková
A.
,
Kubatka
P.
and
Büsselberg
D.
(
2020
)
Counteracting chemoresistance with metformin in breast cancers: targeting cancer stem cells
.
Cancers (Basel)
12
,
1
52
161.
Zhou
Q.M.
,
Sun
Y.
,
Lu
Y.Y.
,
Zhang
H.
,
Chen
Q.L.
and
Su
S.B.
(
2017
)
Curcumin reduces mitomycin C resistance in breast cancer stem cells by regulating Bcl-2 family-mediated apoptosis
.
Cancer Cell Int.
17
,
84
162.
Ramasamy
T.S.
,
Ayob
A.Z.
,
Myint
H.H.L.
,
Thiagarajah
S.
and
Amini
F.
(
2015
)
Targeting colorectal cancer stem cells using curcumin and curcumin analogues: insights into the mechanism of the therapeutic efficacy
.
Cancer Cell Int.
15
,
96
[PubMed]
163.
Wang
J.
,
Wang
C.
and
Bu
G.
(
2018
)
Curcumin inhibits the growth of liver cancer stem cells through the phosphatidylinositol 3-kinase/protein kinase B/mammalian target of rapamycin signaling pathway
.
Exp. Ther. Med.
15
,
3650
3658
[PubMed]
164.
Bentires-Alj
M.
,
Barbu
V.
,
Fillet
M.
,
Chariot
A.
,
Relic
B.
,
Jacobs
N.
et al.
(
2003
)
NF-kappaB transcription factor induces drug resistance through MDR1 expression in cancer cells
.
Oncogene
22
,
90
97
[PubMed]
165.
Lagunas
V.M.
and
Meléndez-Zajgla
J.
(
2008
)
Nuclear factor-kappa B as a resistance factor to platinum-based antineoplasic drugs
.
Met Based Drugs
2008
,
576104
[PubMed]
166.
Li
L.
,
Hu
M.
,
Wang
T.
,
Chen
H.
and
Xu
L.
(
2020
)
Repositioning aspirin to treat lung and breast cancers and overcome acquired resistance to targeted therapy
.
Front. Oncol.
9
,
1503
167.
Fu
J.
,
Xu
Y.
,
Yang
Y.
,
Liu
Y.
,
Ma
L.
and
Zhang
Y.
(
2019
)
Aspirin suppresses chemoresistance and enhances antitumor activity of 5-Fu in 5-Fu-resistant colorectal cancer by abolishing 5-Fu-induced NF-κB activation
.
Sci. Rep.
9
,
16937
168.
Saha
S.
,
Mukherjee
S.
,
Khan
P.
,
Kajal
K.
,
Mazumdar
M.
,
Manna
A.
et al.
(
2016
)
Aspirin suppresses the acquisition of chemoresistance in breast cancer by disrupting an NFκB-IL6 signaling axis responsible for the generation of cancer stem cells
.
Cancer Res.
76
,
2000
2012
[PubMed]
169.
Xue
J.
,
Chi
Y.
,
Chen
Y.
,
Huang
S.
,
Ye
X.
,
Niu
J.
et al.
(
2016
)
MiRNA-621 sensitizes breast cancer to chemotherapy by suppressing FBXO11 and enhancing p53 activity
.
Oncogene
35
,
448
458
[PubMed]
170.
Huang
J.W.
,
Wang
Y.
,
Dhillon
K.K.
,
Calses
P.
,
Villegas
E.
,
Mitchell
P.S.
et al.
(
2013
)
Systematic screen identifies miRNAs that target RAD51 and RAD51D to enhance chemosensitivity
.
Mol. Cancer Res.
11
,
1564
1573
[PubMed]
171.
Zhu
W.
,
Shan
X.
,
Wang
T.
,
Shu
Y.
and
Liu
P.
(
2010
)
miR-181b modulates multidrug resistance by targeting BCL2 in human cancer cell lines
.
Int. J. Cancer
127
,
2520
2529
[PubMed]
172.
Pouliot
L.M.
,
Chen
Y.C.
,
Bai
J.
,
Guha
R.
,
Martin
S.E.
,
Gottesman
M.M.
et al.
(
2012
)
Cisplatin sensitivity mediated by WEE1 and CHK1 is mediated by miR-155 and the miR-15 family
.
Cancer Res.
72
,
5945
5955
[PubMed]
173.
Keyvani-Ghamsari
S.
,
Khorsandi
K.
,
Rasul
A.
and
Zaman
M.K.
(
2021
)
Current understanding of epigenetics mechanism as a novel target in reducing cancer stem cells resistance
.
Clin. Epigenetics
13
,
120
pmc/articles/PMC8164819/
174.
Ahuja
N.
,
Sharma
A.R.
and
Baylin
S.B.
(
2016
)
Epigenetic therapeutics: a new weapon in the war against cancer
.
Annu. Rev. Med.
67
,
73
89
[PubMed]
175.
Zhang
T.
,
Pilko
A.
and
Wollman
R.
(
2020
)
Loci specific epigenetic drug sensitivity
.
Nucleic Acids Res.
48
,
4797
4810
[PubMed]
176.
Amorim
S.
,
Stathis
A.
,
Gleeson
M.
,
Iyengar
S.
,
Magarotto
V.
,
Leleu
X.
et al.
(
2016
)
Bromodomain inhibitor OTX015 in patients with lymphoma or multiple myeloma: a dose-escalation, open-label, pharmacokinetic, phase 1 study
.
Lancet Haematol.
3
,
e196
e204
[PubMed]
177.
Pastore
F.
and
Levine
R.L.
(
2016
)
Epigenetic regulators and their impact on therapy in acute myeloid leukemia
.
Haematologica
101
,
269
278
[PubMed]
178.
Gelato
K.A.
,
Shaikhibrahim
Z.
,
Ocker
M.
and
Haendler
B.
(
2016
)
Targeting epigenetic regulators for cancer therapy: modulation of bromodomain proteins, methyltransferases, demethylases, and microRNAs
.
Expert Opin. Ther. Targets
20
,
783
799
[PubMed]
179.
Al-Jamal
H.A.N.
,
Mat Jusoh
S.A.
,
Hassan
R.
and
Johan
M.F.
(
2015
)
Enhancing SHP-1 expression with 5-azacytidine may inhibit STAT3 activation and confer sensitivity in lestaurtinib (CEP-701)-resistant FLT3-ITD positive acute myeloid leukemia
.
BMC Cancer
15
,
869
https://pubmed.ncbi.nlm.nih.gov/26547689/
[PubMed]
180.
Sabatino
M.A.
,
Geroni
C.
,
Ganzinelli
M.
,
Ceruti
R.
and
Broggini
M.
(
2013
)
Zebularine partially reverses GST methylation in prostate cancer cells and restores sensitivity to the DNA minor groove binder brostallicin
.
Epigenetics
8
,
656
665
https://pubmed.ncbi.nlm.nih.gov/23771052/
[PubMed]
181.
Dimopoulos
K.
,
Søgaard Helbo
A.
,
Fibiger Munch-Petersen
H.
,
Sjö
L.
,
Christensen
J.
,
Sommer Kristensen
L.
et al.
(
2018
)
Dual inhibition of DNMTs and EZH2 can overcome both intrinsic and acquired resistance of myeloma cells to IMiDs in a cereblon-independent manner
.
Mol. Oncol.
12
,
180
195
https://pubmed.ncbi.nlm.nih.gov/29130642/
[PubMed]
182.
Füller
M.
,
Klein
M.
,
Schmidt
E.
,
Rohde
C.
,
Göllner
S.
,
Schulze
I.
et al.
(
2015
)
5-azacytidine enhances efficacy of multiple chemotherapy drugs in AML and lung cancer with modulation of CpG methylation
.
Int. J. Oncol.
46
,
1192
1204
https://pubmed.ncbi.nlm.nih.gov/25501798/
[PubMed]
183.
Xylinas
E.
,
Hassler
M.R.
,
Zhuang
D.
,
Krzywinski
M.
,
Erdem
Z.
,
Robinson
B.D.
et al.
(
2016
)
An epigenomic approach to improving response to neoadjuvant cisplatin chemotherapy in bladder cancer
.
Biomolecules
6
,
37
https://pubmed.ni.nlm.nih.gov/27598218/
[PubMed]
184.
Cui
Y.
,
Naz
A.
,
Thompson
D.H.
and
Irudayaraj
J.
(
2015
)
Decitabine nanoconjugate sensitizes human glioblastoma cells to temozolomide
.
Mol. Pharm.
12
,
1279
1288
https://pubmed.ncbi.nlm.nih.gov/25751281/
[PubMed]
185.
Eramo
A.
,
Pallini
R.
,
Lotti
F.
,
Sette
G.
,
Patti
M.
,
Bartucci
M.
et al.
(
2005
)
Inhibition of DNA methylation sensitizes glioblastoma for tumor necrosis factor-related apoptosis-inducing ligand-mediated destruction
.
Cancer Res.
65
,
11469
11477
https://pubmed.ncbi.nlm.nih.gov/16357155/
[PubMed]
186.
Shire
A.
,
Lomberk
G.
,
Lai
J.-P.
,
Zou
H.
,
Tsuchiya
N.
,
Aderca
I.
et al.
(
2015
)
Restoration of epigenetically silenced SULF1 expression by 5-aza-2-deoxycytidine sensitizes hepatocellular carcinoma cells to chemotherapy-induced apoptosis
.
Med. Epigenet.
3
,
1
18
https://pubmed.ncbi.nlm.nih.gov/26236329/
[PubMed]
187.
Reu
F.J.
,
Leaman
D.W.
,
Maitra
R.R.
,
Bae
S.I.
,
Cherkassky
L.
,
Fox
M.W.
et al.
(
2006
)
Expression of RASSF1A, an epigenetically silenced tumor suppressor, overcomes resistance to apoptosis induction by interferons
.
Cancer Res.
66
,
2785
2793
https://pubmed.ncbi.nlm.nih.gov/16510600/
[PubMed]
188.
Niitsu
N.
,
Hayashi
Y.
,
Sugita
K.
and
Honma
Y.
(
2001
)
Sensitization by 5-aza-2’-deoxycytidine of leukaemia cells with MLL abnormalities to induction of differentiation by all-trans retinoic acid and 1alpha,25-dihydroxyvitamin D3
.
Br. J. Haematol.
112
,
315
326
https://pubmed.ncbi.nlm.nih.gov/11167824/
[PubMed]
189.
Li
Y.C.
,
Wang
Y.
,
Li
D.D.
,
Zhang
Y.
,
Zhao
T.C.
and
Li
C.F.
(
2018
)
Procaine is a specific DNA methylation inhibitor with anti-tumor effect for human gastric cancer
.
J. Cell. Biochem.
119
,
2440
2449
https://pubmed.ncbi.nlm.nih.gov/28926119/
[PubMed]
190.
Zhou
H.
,
Xu
M.
,
Luo
G.
and
Zhang
Y.
(
2007
)
Effects of procaine on human nasopharyngeal carcinoma cell strain CNE-2Z
.
Lin Chung Er Bi Yan Hou Tou Jing Wai Ke Za Zhi
21
,
1118
1121
https://pubmed.ncbi.nlm.nih.gov/18330258/
[PubMed]
191.
Balch
C.
,
Yan
P.
,
Craft
T.
,
Young
S.
,
Skalnik
D.G.
,
Huang
T.H.M.
et al.
(
2005
)
Antimitogenic and chemosensitizing effects of the methylation inhibitor zebularine in ovarian cancer
.
Mol. Cancer Ther.
4
,
1505
1514
https://pubmed.ncbi.nlm.nih.gov/16227399/
[PubMed]
192.
Okabe
S.
,
Tauchi
T.
,
Tanaka
Y.
,
Kimura
S.
,
Maekawa
T.
and
Ohyashiki
K.
(
2013
)
Activity of histone deacetylase inhibitors and an Aurora kinase inhibitor in BCR-ABL-expressing leukemia cells: combination of HDAC and Aurora inhibitors in BCR-ABL-expressing cells
.
Cancer Cell Int.
13
,
32
https://pubmed.ncbi.nlm.nih.gov/23556431/
[PubMed]
193.
Galanis
E.
,
Keith Anderson
S.
,
Miller
C.R.
,
Sarkaria
J.N.
,
Jaeckle
K.
,
Buckner
J.C.
et al.
(
2018
)
Phase I/II trial of vorinostat combined with temozolomide and radiation therapy for newly diagnosed glioblastoma: results of Alliance N0874/ABTC 02
.
Neuro Oncol.
20
,
546
556
,
https://pubmed.ncbi.nlm.nih.gov/29016887/
[PubMed]
194.
Butler
L.M.
,
Liapis
V.
,
Bouralexis
S.
,
Welldon
K.
,
Hay
S.
,
Thai
L.M.
et al.
(
2006
)
The histone deacetylase inhibitor, suberoylanilide hydroxamic acid, overcomes resistance of human breast cancer cells to Apo2L/TRAIL
.
Int. J. Cancer
119
,
944
954
,
https://pubmed.ncbi.nlm.nih.gov/16550602/
[PubMed]
195.
Frew
A.J.
,
Lindemann
R.K.
,
Martin
B.P.
,
Clarke
C.J.P.
,
Sharkey
J.
,
Anthony
D.A.
et al.
(
2008
)
Combination therapy of established cancer using a histone deacetylase inhibitor and a TRAIL receptor agonist
.
Proc. Natl. Acad. Sci. U.S.A.
105
,
11317
11322
,
https://pubmed.ncbi.nlm.nih.gov/18685088/
[PubMed]
196.
Jazirehi
A.R.
,
Kurdistani
S.K.
and
Economou
J.S.
(
2014
)
Histone deacetylase inhibitor sensitizes apoptosis-resistant melanomas to cytotoxic human T lymphocytes through regulation of TRAIL/DR5 pathway
.
J. Immunol.
192
,
3981
3989
,
https://pubmed.ncbi.nlm.nih.gov/24639349/
[PubMed]
197.
Susanto
J.M.
,
Colvin
E.K.
,
Pinese
M.
,
Chang
D.K.
,
Pajic
M.
,
Mawson
A.
et al.
(
2015
)
The epigenetic agents suberoylanilide hydroxamic acid and 5-AZA-2’ deoxycytidine decrease cell proliferation, induce cell death and delay the growth of MiaPaCa2 pancreatic cancer cells in vivo
.
Int. J. Oncol.
46
,
2223
2230
,
https://pubmed.ncbi.nlm.nih.gov/25695794/
[PubMed]
198.
Mueller
S.
,
Yang
X.
,
Sottero
T.L.
,
Gragg
A.
,
Prasad
G.
,
Polley
M.Y.
et al.
(
2011
)
Cooperation of the HDAC inhibitor vorinostat and radiation in metastatic neuroblastoma: efficacy and underlying mechanisms
.
Cancer Lett.
306
,
223
229
,
https://pubmed.ncbi.nlm.nih.gov/21497989/
[PubMed]
199.
Poulaki
V.
,
Mitsiades
C.S.
,
Kotoula
V.
,
Negri
J.
,
McMullan
C.
,
Miller
J.W.
et al.
(
2009
)
Molecular sequelae of histone deacetylase inhibition in human retinoblastoma cell lines: clinical implications
.
Invest. Ophthalmol. Vis. Sci.
50
,
4072
4079
,
https://pubmed.ncbi.nlm.nih.gov/19387079/
[PubMed]
200.
Li
C.T.
,
Hsiao
Y.M.
,
Wu
T.C.
,
Lin
Y.W.
,
Yeh
K.T.
and
Ko
J.L.
(
2011
)
Vorinostat, SAHA, represses telomerase activity via epigenetic regulation of telomerase reverse transcriptase in non-small cell lung cancer cells
.
J. Cell. Biochem.
112
,
3044
3053
,
https://pubmed.ncbi.nlm.nih.gov/21678477/
[PubMed]
201.
Gressette
M.
,
Vérillaud
B.
,
Jimenez-Pailhès
A.S.
,
Lelièvre
H.
,
Lo
K.W.
,
Ferrand
F.R.
et al.
(
2014
)
Treatment of nasopharyngeal carcinoma cells with the histone-deacetylase inhibitor abexinostat: cooperative effects with cis-platin and radiotherapy on patient-derived xenografts
.
PLoS ONE
9
,
e91325
https://pubmed.ncbi.nlm.nih.gov/24618637/
[PubMed]
202.
Salvador
M.A.
,
Wicinski
J.
,
Cabaud
O.
,
Toiron
Y.
,
Finetti
P.
,
Josselin
E.
et al.
(
2013
)
The histone deacetylase inhibitor abexinostat induces cancer stem cells differentiation in breast cancer with low Xist expression
.
Clin. Cancer Res.
19
,
6520
6531
,
https://pubmed.ncbi.nlm.nih.gov/24141629/
[PubMed]
203.
To
K.K.W.
,
Tong
W.S.
and
Fu
L.W.
(
2017
)
Reversal of platinum drug resistance by the histone deacetylase inhibitor belinostat
.
Lung Cancer
103
,
58
65
,
https://pubmed.ncbi.nlm.nih.gov/28024697/
[PubMed]
204.
Nguyen
T.
,
Parker
R.
,
Hawkins
E.
,
Holkova
B.
,
Yazbeck
V.
,
Kolluri
A.
et al.
(
2019
)
Correction: Synergistic interactions between PLK1 and HDAC inhibitors in non-Hodgkin’s lymphoma cells occur in vitro and in vivo and proceed through multiple mechanisms
.
Oncotarget
10
,
5383
5384
,
https://pubmed.ncbi.nlm.nih.gov/31523396/
[PubMed]
205.
Zhou
L.
,
Chen
S.
,
Zhang
Y.
,
Kmieciak
M.
,
Leng
Y.
,
Li
L.
et al.
(
2016
)
The NAE inhibitor pevonedistat interacts with the HDAC inhibitor belinostat to target AML cells by disrupting the DDR
.
Blood
127
,
2219
2230
[PubMed]
206.
Greve
G.
,
Schiffmann
I.
,
Pfeifer
D.
,
Pantic
M.
,
Schüler
J.
and
Lübbert
M.
(
2015
)
The pan-HDAC inhibitor panobinostat acts as a sensitizer for erlotinib activity in EGFR-mutated and -wildtype non-small cell lung cancer cells
.
BMC Cancer
15
,
947
,
https://pubmed.ncbi.nlm.nih.gov/26675484/
[PubMed]
207.
Wang
L.
,
Syn
N.L.X.
,
Subhash
V.V.
,
Any
Y.
,
Thuya
W.L.
,
Cheow
E.S.H.
et al.
(
2018
)
Pan-HDAC inhibition by panobinostat mediates chemosensitization to carboplatin in non-small cell lung cancer via attenuation of EGFR signaling
.
Cancer Lett.
417
,
152
160
,
https://pubmed.ncbi.nlm.nih.gov/29306016/
[PubMed]
208.
Liu
Y.
,
Mondello
P.
,
Erazo
T.
,
Tannan
N.B.
,
Asgari
Z.
,
de Stanchina
E.
et al.
(
2018
)
NOXA genetic amplification or pharmacologic induction primes lymphoma cells to BCL2 inhibitor-induced cell death
.
Proc. Natl. Acad. Sci. U.S.A.
115
,
12034
12039
,
https://pubmed.ncbi.nlm.nih.gov/30404918/
[PubMed]
209.
Qi
W.
,
Zhang
W.
,
Edwards
H.
,
Chu
R.
,
Madlambayan
G.J.
,
Taub
J.W.
et al.
(
2015
)
Synergistic anti-leukemic interactions between panobinostat and MK-1775 in acute myeloid leukemia ex vivo
.
Cancer Biol. Ther.
16
,
1784
1793
,
https://pubmed.ncbi.nlm.nih.gov/26529495/
[PubMed]
210.
Uy
G.L.
,
Duncavage
E.J.
,
Chang
G.S.
,
Jacoby
M.A.
,
Miller
C.A.
,
Shao
J.
et al.
(
2017
)
Dynamic changes in the clonal structure of MDS and AML in response to epigenetic therapy
.
Leukemia
31
,
872
881
,
https://pubmed.ncbi.nlm.nih.gov/27740633/
[PubMed]
211.
Tan
P.
,
Wei
A.
,
Mithraprabhu
S.
,
Cummings
N.
,
Liu
H.B.
,
Perugini
M.
et al.
(
2014
)
Dual epigenetic targeting with panobinostat and azacitidine in acute myeloid leukemia and high-risk myelodysplastic syndrome
.
Blood Cancer J.
4
,
e170
https://pubmed.ncbi.nlm.nih.gov/24413064/
[PubMed]
212.
Wilson
A.J.
,
Sarfo-Kantanka
K.
,
Barrack
T.
,
Steck
A.
,
Saskowski
J.
,
Crispens
M.A.
et al.
(
2016
)
Panobinostat sensitizes cyclin E high, homologous recombination-proficient ovarian cancer to olaparib
.
Gynecol. Oncol.
143
,
143
151
,
https://pubmed.ncbi.nlm.nih.gov/27444036/
[PubMed]
213.
Kalal
B.S.
,
Modi
P.K.
,
Najar
M.A.
,
Behera
S.K.
,
Upadhya
D.
,
Prasad
T.S.K.
et al.
(
2021
)
Hyperphosphorylation of HDAC2 promotes drug resistance in a novel dual drug resistant mouse melanoma cell line model: an in vitro study
.
Am. J. Cancer Res.
11
,
5881
5901
,
https://pubmed.ncbi.nlm.nih.gov/35018231/
[PubMed]
214.
Roos
W.P.
,
Jöst
E.
,
Belohlavek
C.
,
Nagel
G.
,
Fritz
G.
and
Kaina
B.
(
2011
)
Intrinsic anticancer drug resistance of malignant melanoma cells is abrogated by IFN-β and valproic acid
.
Cancer Res.
71
,
4150
4160
,
https://pubmed.ncbi.nlm.nih.gov/21493591/
[PubMed]
215.
Jo
Y.K.
,
Park
N.Y.
,
Shin
J.H.
,
Jo
D.S.
,
Bae
J.E.
,
Choi
E.S.
et al.
(
2018
)
Up-regulation of UVRAG by HDAC1 inhibition attenuates 5FU-induced cell death in HCT116 colorectal cancer cells
.
Anticancer Res.
38
,
271
277
,
https://pubmed.ncbi.nlm.nih.gov/29277783/
[PubMed]
216.
Iannelli
F.
,
Roca
M.S.
,
Lombardi
R.
,
Ciardiello
C.
,
Grumetti
L.
,
De Rienzo
S.
et al.
(
2020
)
Synergistic antitumor interaction of valproic acid and simvastatin sensitizes prostate cancer to docetaxel by targeting CSCs compartment via YAP inhibition
.
J. Exp. Clin. Cancer Res.
39
,
213
,
https://pubmed.ncbi.nlm.nih.gov/33032653/
[PubMed]
217.
Sharda
A.
,
Rashid
M.
,
Shah
S.G.
,
Sharma
A.K.
,
Singh
S.R.
,
Gera
P.
et al.
(
2020
)
Elevated HDAC activity and altered histone phospho-acetylation confer acquired radio-resistant phenotype to breast cancer cells
.
Clin. Epigenetics
12
,
4
,
https://pubmed.ncbi.nlm.nih.gov/31900196/
[PubMed]
218.
Heers
H.
,
Stanislaw
J.
,
Harrelson
J.
and
Lee
M.W.
(
2018
)
Valproic acid as an adjunctive therapeutic agent for the treatment of breast cancer
.
Eur. J. Pharmacol.
835
,
61
74
,
https://pubmed.ncbi.nlm.nih.gov/30075223/
[PubMed]
219.
Guo
W.
,
Chen
Z.
,
Tan
L.
,
Wu
Q.
,
Ren
X.
,
Fu
C.
et al.
(
2020
)
l-Cysteine decorated nanoscale metal-organic frameworks delivering valproic acid/cisplatin for drug-resistant lung cancer therapy
.
Chem. Commun. (Camb.)
56
,
3919
3922
,
https://pubmed.ncbi.nlm.nih.gov/32149283/
[PubMed]
220.
Iannelli
F.
,
Zotti
A.I.
,
Roca
M.S.
,
Grumetti
L.
,
Lombardi
R.
,
Moccia
T.
et al.
(
2020
)
Valproic acid synergizes with cisplatin and cetuximab in vitro and in vivo in head and neck cancer by targeting the mechanisms of resistance
.
Front. Cell. Dev. Biol.
8
,
732
https://pubmed.ncbi.nlm.nih.gov/33015030/
[PubMed]
221.
Xu
Y.
,
Xu
D.
,
Zhu
S.J.
,
Ye
B.
,
Dong
J. Da
,
Zhang
Y.L.
et al.
(
2015
)
Induction of apoptosis and autophagy in metastatic thyroid cancer cells by valproic acid (VPA)
.
Int. J. Clin. Exp. Pathol.
8
,
8291
8297
,
https://pubmed.ncbi.nlm.nih.gov/26339399/
[PubMed]
222.
Fu
Y.-T.
,
Zheng
H.-B.
,
Zhou
L.
,
Zhang
D.-Q.
,
Liu
X.-L.
and
Sun
H.
(
2017
)
Valproic acid, targets papillary thyroid cancer through inhibition of c-Met signalling pathway
.
Am. J. Transl. Res.
9
,
3138
3147
,
https://pubmed.ncbi.nlm.nih.gov/28670399/
[PubMed]
223.
Wu
A.
,
Wu
K.
,
Li
M.
,
Bao
L.
,
Shen
X.
,
Li
S.
et al.
(
2015
)
Upregulation of microRNA-492 induced by epigenetic drug treatment inhibits the malignant phenotype of clear cell renal cell carcinoma in vitro
.
Mol. Med. Rep.
12
,
1413
1420
,
https://pubmed.ncbi.nlm.nih.gov/25815441/
[PubMed]
224.
Denlinger
C.E.
,
Rundall
B.K.
and
Jones
D.R.
(
2005
)
Inhibition of phosphatidylinositol 3-kinase/Akt and histone deacetylase activity induces apoptosis in non-small cell lung cancer in vitro and in vivo
.
J. Thorac. Cardiovasc. Surg.
130
,
1422
1429
,
https://pubmed.ncbi.nlm.nih.gov/16256798/
[PubMed]
225.
Lee
J.
,
Bartholomeusz
C.
,
Mansour
O.
,
Humphries
J.
,
Hortobagyi
G.N.
,
Ordentlich
P.
et al.
(
2014
)
A class I histone deacetylase inhibitor, entinostat, enhances lapatinib efficacy in HER2-overexpressing breast cancer cells through FOXO3-mediated Bim1 expression
.
Breast Cancer Res. Treat.
146
,
259
272
,
https://pubmed.ncbi.nlm.nih.gov/24916181/
[PubMed]
226.
Sabnis
G.J.
,
Goloubeva
O.
,
Chumsri
S.
,
Nguyen
N.
,
Sukumar
S.
and
Brodie
A.M.H.
(
2011
)
Functional activation of the estrogen receptor-α and aromatase by the HDAC inhibitor entinostat sensitizes ER-negative tumors to letrozole
.
Cancer Res.
71
,
1893
1903
,
https://pubmed.ncbi.nlm.nih.gov/21245100/
[PubMed]
227.
Shen
L.
,
Ciesielski
M.
,
Ramakrishnan
S.
,
Miles
K.M.
,
Ellis
L.
,
Sotomayor
P.
et al.
(
2012
)
Class I histone deacetylase inhibitor entinostat suppresses regulatory T cells and enhances immunotherapies in renal and prostate cancer models
.
PLoS ONE
7
,
e30815
.
https://pubmed.ncbi.nlm.nih.gov/22303460/
228.
Topper
M.J.
,
Vaz
M.
,
Chiappinelli
K.B.
,
DeStefano Shields
C.E.
,
Niknafs
N.
,
Yen
R.W.C.
et al.
(
2017
)
Epigenetic therapy ties MYC depletion to reversing immune evasion and treating lung cancer
.
Cell
171
,
1284.e21
1300.e21
,
https://pubmed.ncbi.nlm.nih.gov/29195073/
[PubMed]
229.
Cooper
S.J.
,
Von Roemeling
C.A.
,
Kang
K.H.
,
Marlow
L.A.
,
Grebe
S.K.
,
Menefee
M.E.
et al.
(
2012
)
Reexpression of tumor suppressor, sFRP1, leads to antitumor synergy of combined HDAC and methyltransferase inhibitors in chemoresistant cancers
.
Mol. Cancer Ther.
11
,
2105
2115
,
https://pubmed.ncbi.nlm.nih.gov/22826467/
[PubMed]
230.
Li
H.
,
Ma
L.
,
Bian
X.
,
Lv
Y.
and
Lin
W.
(
2021
)
FK228 sensitizes radioresistant small cell lung cancer cells to radiation
.
Clin Epigenetics
13
,
41
,
https://pubmed.ncbi.nlm.nih.gov/33632300/
231.
Hirokawa
Y.
,
Arnold
M.
,
Nakajima
H.
,
Zalcberg
J.
and
Maruta
H.
(
2005
)
Signal therapy of breast cancers by the HDAC inhibitor FK228 that blocks the activation of PAK1 and abrogates the tamoxifen-resistance
.
Cancer Biol. Ther.
4
,
956
960
,
https://pubmed.ncbi.nlm.nih.gov/16082189/
[PubMed]
232.
Choudhary
S.
,
Sood
S.
and
Wang
H.C.R.
(
2013
)
Synergistic induction of cancer cell death and reduction of clonogenic resistance by cisplatin and FK228
.
Biochem. Biophys. Res. Commun.
436
,
325
330
,
https://pubmed.ncbi.nlm.nih.gov/23743194/
[PubMed]
233.
McGraw
A.L.
(
2013
)
Romidepsin for the treatment of T-cell lymphomas
.
Am. J. Health Syst. Pharm.
70
,
1115
1122
,
https://pubmed.ncbi.nlm.nih.gov/23784158/
[PubMed]
234.
Mukhopadhyay
N.K.
,
Weisberg
E.
,
Gilchrist
D.
,
Bueno
R.
,
Sugarbaker
D.J.
and
Jaklitsch
M.T.
(
2006
)
Effectiveness of trichostatin A as a potential candidate for anticancer therapy in non-small-cell lung cancer
.
Ann. Thorac. Surg.
81
,
1034
1042
,
https://pubmed.ncbi.nlm.nih.gov/16488717/
[PubMed]
235.
Xu
Q.
,
Liu
X.
,
Zhu
S.
,
Hu
X.
,
Niu
H.
,
Zhang
X.
et al.
(
2018
)
Hyper-acetylation contributes to the sensitivity of chemo-resistant prostate cancer cells to histone deacetylase inhibitor Trichostatin A
.
J. Cell. Mol. Med.
22
,
1909
1922
,
https://pubmed.ncbi.nlm.nih.gov/29327812/
[PubMed]
236.
El-Zawahry
A.
,
Lu
P.
,
White
S.J.
and
Voelkel-Johnson
C.
(
2006
)
In vitro efficacy of AdTRAIL gene therapy of bladder cancer is enhanced by trichostatin A-mediated restoration of CAR expression and downregulation of cFLIP and Bcl-XL
.
Cancer Gene Ther.
13
,
281
289
,
https://pubmed.ncbi.nlm.nih.gov/16167063/
[PubMed]
237.
Karagiannis
T.C.
,
Harikrishnan
K.N.
and
El-Osta
A.
(
2005
)
The histone deacetylase inhibitor, Trichostatin A, enhances radiation sensitivity and accumulation of gammaH2A.X
.
Cancer Biol. Ther.
4
,
787
793
,
https://pubmed.ncbi.nlm.nih.gov/16082178/
[PubMed]
238.
Höring
E.
,
Podlech
O.
,
Silkenstedt
B.
,
Rota
I.A.
,
Adamopoulou
E.
and
Naumann
U.
(
2013
)
The histone deacetylase inhibitor trichostatin a promotes apoptosis and antitumor immunity in glioblastoma cells
.
Anticancer Res.
33
,
1351
1360
,
https://pubmed.ncbi.nlm.nih.gov/23564772/
[PubMed]
239.
Boulding
T.
,
McCuaig
R.D.
,
Tan
A.
,
Hardy
K.
,
Wu
F.
,
Dunn
J.
et al.
(
2018
)
LSD1 activation promotes inducible EMT programs and modulates the tumour microenvironment in breast cancer
.
Sci. Rep.
8
,
73
https://pubmed.ncbi.nlm.nih.gov/29311580/
[PubMed]
240.
Lee
H.T.
,
Choi
M.R.
,
Doh
M.S.
,
Jung
K.H.
and
Chai
Y.G.
(
2013
)
Effects of the monoamine oxidase inhibitors pargyline and tranylcypromine on cellular proliferation in human prostate cancer cells
.
Oncol. Rep.
30
,
1587
1592
,
https://pubmed.ncbi.nlm.nih.gov/23900512/
[PubMed]
241.
Dalvi
P.S.
,
MacHeleidt
I.F.
,
Lim
S.Y.
,
Meemboor
S.
,
Müller
M.
,
Eischeid-Scholz
H.
et al.
(
2019
)
LSD1 inhibition attenuates tumor growth by disrupting PLK1 mitotic pathway
.
Mol. Cancer Res.
17
,
1326
1337
,
https://pubmed.ncbi.nlm.nih.gov/30760542/
[PubMed]
242.
Feng
S.
,
Jin
Y.
,
Cui
M.
and
Zheng
J.
(
2016
)
Lysine-specific demethylase 1 (LSD1) inhibitor S2101 induces autophagy via the AKT/mTOR pathway in SKOV3 ovarian cancer cells
.
Med. Sci. Monit.
22
,
4742
4748
,
https://pubmed.ncbi.nlm.nih.gov/27914215/
[PubMed]
243.
Benedetti
R.
,
Dell’aversana
C.
,
De Marchi
T.
,
Rotili
D.
,
Liu
N.Q.
,
Novakovic
B.
et al.
(
2019
)
Inhibition of histone demethylases LSD1 and UTX regulates ERα signaling in breast cancer
.
Cancers (Basel)
11
,
2027
https://pubmed.ncbi.nlm.nih.gov/31888209/
[PubMed]
244.
Wang
W.
,
Oguz
G.
,
Lee
P.L.
,
Bao
Y.
,
Wang
P.
,
Terp
M.G.
et al.
(
2018
)
KDM4B-regulated unfolded protein response as a therapeutic vulnerability in PTEN-deficient breast cancer
.
J. Exp. Med.
215
,
2833
2849
,
https://pubmed.ncbi.nlm.nih.gov/30266800/
[PubMed]
245.
Parrish
J.K.
,
McCann
T.S.
,
Sechler
M.
,
Sobral
L.M.
,
Ren
W.
,
Jones
K.L.
et al.
(
2018
)
The Jumonji-domain histone demethylase inhibitor JIB-04 deregulates oncogenic programs and increases DNA damage in Ewing Sarcoma, resulting in impaired cell proliferation and survival, and reduced tumor growth
.
Oncotarget
9
,
33110
33123
,
https://pubmed.ncbi.nlm.nih.gov/30237855/
[PubMed]
246.
Marsh
S.
and
Jimeno
A.
(
2020
)
Tazemetostat for the treatment of multiple types of hematological malignancies and solid tumors
.
Drugs Today (Barc.)
56
,
377
387
,
https://pubmed.ncbi.nlm.nih.gov/32525136/
[PubMed]
247.
von Keudell
G.
and
Salles
G.
(
2021
)
The role of tazemetostat in relapsed/refractory follicular lymphoma
.
Ther. Adv. Hematol.
12
,
https://pubmed.ncbi.nlm.nih.gov/34104370/
[PubMed]
248.
Mondello
P.
and
Ansell
S.M.
(
2022
)
Tazemetostat: a treatment option for relapsed/refractory follicular lymphoma
.
Expert Opin. Pharmacother.
23
,
295
301
,
https://pubmed.ncbi.nlm.nih.gov/34904909/
[PubMed]
249.
Fu
H.
,
Cheng
L.
,
Sa
R.
,
Jin
Y.
and
Chen
L.
(
2020
)
Combined tazemetostat and MAPKi enhances differentiation of papillary thyroid cancer cells harbouring BRAF V600E by synergistically decreasing global trimethylation of H3K27
.
J. Cell. Mol. Med.
24
,
3336
3345
,
https://pubmed.ncbi.nlm.nih.gov/31970877/
[PubMed]
250.
Tan
X.
,
Zhang
Z.
,
Liu
P.
,
Yao
H.
,
Shen
L.
and
Tong
J.S.
(
2020
)
Inhibition of EZH2 enhances the therapeutic effect of 5-FU via PUMA upregulation in colorectal cancer
.
Cell Death Dis.
11
,
1061
,
https://pubmed.ncbi.nlm.nih.gov/33311453/
251.
Yang
Q.
,
Zhao
S.
,
Shi
Z.
,
Cao
L.
,
Liu
J.
,
Pan
T.
et al.
(
2021
)
Chemotherapy-elicited exosomal miR-378a-3p and miR-378d promote breast cancer stemness and chemoresistance via the activation of EZH2/STAT3 signaling
.
J. Exp. Clin. Cancer Res.
40
,
120
,
https://pubmed.ncbi.nlm.nih.gov/33823894/
252.
Barazeghi
E.
,
Hellman
P.
,
Norlén
O.
,
Westin
G.
and
Stålberg
P.
(
2021
)
EZH2 presents a therapeutic target for neuroendocrine tumors of the small intestine
.
Sci. Rep.
11
,
22733
https://pubmed.ncbi.nlm.nih.gov/34815475/
[PubMed]
253.
Vaswani
R.G.
,
Gehling
V.S.
,
Dakin
L.A.
,
Cook
A.S.
,
Nasveschuk
C.G.
,
Duplessis
M.
et al.
(
2016
)
Identification of (R)-N-((4-Methoxy-6-methyl-2-oxo-1,2-dihydropyridin-3-yl)methyl)-2-methyl-1-(1-(1-(2,2,2-trifluoroethyl)piperidin-4-yl)ethyl)-1H-indole-3-carboxamide (CPI-1205), a potent and selective inhibitor of histone methyltransferase EZH2, suitable for phase I clinical trials for B-cell lymphomas
.
J. Med. Chem.
59
,
9928
9941
,
https://pubmed.ncbi.nlm.nih.gov/27739677/
[PubMed]
254.
Zhang
Y.
,
Zhou
L.
,
Safran
H.
,
Borsuk
R.
,
Lulla
R.
,
Tapinos
N.
et al.
(
2021
)
EZH2i EPZ-6438 and HDACi vorinostat synergize with ONC201/TIC10 to activate integrated stress response, DR5, reduce H3K27 methylation, ClpX and promote apoptosis of multiple tumor types including DIPG
.
Neoplasia
23
,
792
810
,
https://pubmed.ncbi.nlm.nih.gov/34246076/
[PubMed]
255.
Stein
E.M.
,
Garcia-Manero
G.
,
Rizzieri
D.A.
,
Tibes
R.
,
Berdeja
J.G.
,
Savona
M.R.
et al.
(
2018
)
The DOT1L inhibitor pinometostat reduces H3K79 methylation and has modest clinical activity in adult acute leukemia
.
Blood
131
,
2662
2669
,
https://pubmed.ncbi.nlm.nih.gov/29724899/
256.
Richter
W.F.
,
Shah
R.N.
and
Ruthenburg
A.J.
(
2021
)
Non-canonical H3K79me2-dependent pathways promote the survival of MLL-rearranged leukemia
.
eLife
10
,
e64960
https://pubmed.ncbi.nlm.nih.gov/34263728/
257.
Filippakopoulos
P.
,
Qi
J.
,
Picaud
S.
,
Shen
Y.
,
Smith
W.B.
,
Fedorov
O.
et al.
(
2010
)
Selective inhibition of BET bromodomains
.
Nature
468
,
1067
1073
,
https://pubmed.ncbi.nlm.nih.gov/20871596/
[PubMed]
258.
Zuber
J.
,
Shi
J.
,
Wang
E.
,
Rappaport
A.R.
,
Herrmann
H.
,
Sison
E.A.
et al.
(
2011
)
RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia
.
Nature
478
,
524
528
,
https://pubmed.ncbi.nlm.nih.gov/21814200/
[PubMed]
259.
Bandopadhayay
P.
,
Bergthold
G.
,
Nguyen
B.
,
Schubert
S.
,
Gholamin
S.
,
Tang
Y.
et al.
(
2014
)
BET bromodomain inhibition of MYC-amplified medulloblastoma
.
Clin. Cancer Res.
20
,
912
925
,
https://pubmed.ncbi.nlm.nih.gov/24297863/
[PubMed]
260.
Shu
S.
,
Lin
C.Y.
,
He
H.H.
,
Witwicki
R.M.
,
Tabassum
D.P.
,
Roberts
J.M.
et al.
(
2016
)
Response and resistance to BET bromodomain inhibitors in triple-negative breast cancer
.
Nature
529
,
413
417
,
https://pubmed.ncbi.nlm.nih.gov/26735014/
[PubMed]
261.
Klingbeil
O.
,
Lesche
R.
,
Gelato
K.A.
,
Haendler
B.
and
Lejeune
P.
(
2016
)
Inhibition of BET bromodomain-dependent XIAP and FLIP expression sensitizes KRAS-mutated NSCLC to pro-apoptotic agents
.
Cell Death Dis.
7
,
e2365
,
https://pubmed.ncbi.nlm.nih.gov/27607580/
[PubMed]
262.
Ramasamy
K.
,
Nooka
A.
,
Quach
H.
,
Htut
M.
,
Popat
R.
,
Liedtke
M.
et al.
(
2021
)
A phase 1b dose-escalation/expansion study of BET inhibitor RO6870810 in patients with advanced multiple myeloma
.
Blood Cancer J.
11
,
149
,
https://pubmed.ncbi.nlm.nih.gov/34480019/
263.
Roboz
G.J.
,
Desai
P.
,
Lee
S.
,
Ritchie
E.K.
,
Winer
E.S.
,
DeMario
M.
et al.
(
2021
)
A dose escalation study of RO6870810/TEN-10 in patients with acute myeloid leukemia and myelodysplastic syndrome
.
Leuk. Lymphoma
62
,
1740
1748
,
https://pubmed.ncbi.nlm.nih.gov/33586590/
[PubMed]
264.
Henssen
A.
,
Althoff
K.
,
Odersky
A.
,
Beckers
A.
,
Koche
R.
,
Speleman
F.
et al.
(
2016
)
Targeting MYCN-driven transcription By BET-bromodomain inhibition
.
Clin. Cancer Res.
22
,
2470
2781
,
https://pubmed.ncbi.nlm.nih.gov/26631615/
[PubMed]
265.
Vázquez
R.
,
Licandro
S.A.
,
Astorgues-Xerri
L.
,
Lettera
E.
,
Panini
N.
,
Romano
M.
et al.
(
2017
)
Promising in vivo efficacy of the BET bromodomain inhibitor OTX015/MK-8628 in malignant pleural mesothelioma xenografts
.
Int. J. Cancer
140
,
197
207
,
https://pubmed.ncbi.nlm.nih.gov/27594045/
[PubMed]
266.
Shi
J.
,
Song
S.
,
Han
H.
,
Xu
H.
,
Huang
M.
,
Qian
C.
et al.
(
2018
)
Potent activity of the bromodomain inhibitor OTX015 in multiple myeloma
.
Mol. Pharm.
15
,
4139
4147
,
https://pubmed.ncbi.nlm.nih.gov/30048594/
[PubMed]
267.
Boi
M.
,
Gaudio
E.
,
Bonetti
P.
,
Kwee
I.
,
Bernasconi
E.
,
Tarantelli
C.
et al.
(
2015
)
The BET bromodomain inhibitor OTX015 affects pathogenetic pathways in preclinical B-cell tumor models and synergizes with targeted drugs
.
Clin. Cancer Res.
21
,
1628
1638
,
https://pubmed.ncbi.nlm.nih.gov/25623213/
[PubMed]
268.
Chaidos
A.
,
Caputo
V.
,
Gouvedenou
K.
,
Liu
B.
,
Marigo
I.
,
Chaudhry
M.S.
et al.
(
2014
)
Potent antimyeloma activity of the novel bromodomain inhibitors I-BET151 and I-BET762
.
Blood
123
,
697
705
,
https://pubmed.ncbi.nlm.nih.gov/24335499/
[PubMed]
269.
Xie
F.
,
Huang
M.
,
Lin
X.
,
Liu
C.
,
Liu
Z.
,
Meng
F.
et al.
(
2018
)
The BET inhibitor I-BET762 inhibits pancreatic ductal adenocarcinoma cell proliferation and enhances the therapeutic effect of gemcitabine
.
Sci. Rep.
8
,
8102
https://pubmed.ncbi.nlm.nih.gov/29802402/
270.
Wyce
A.
,
Ganji
G.
,
Smitheman
K.N.
,
Chung
C.W.
,
Korenchuk
S.
,
Bai
Y.
et al.
(
2013
)
BET inhibition silences expression of MYCN and BCL2 and induces cytotoxicity in neuroblastoma tumor models
.
PLoS ONE
8
,
e72967
https://pubmed.ncbi.nlm.nih.gov/24009722/
[PubMed]
271.
Gros
C.
,
Fahy
J.
,
Halby
L.
,
Dufau
I.
,
Erdmann
A.
,
Gregoire
J.M.
et al.
(
2012
)
DNA methylation inhibitors in cancer: recent and future approaches
.
Biochimie
94
,
2280
2296
,
https://pubmed.ncbi.nlm.nih.gov/22967704/
[PubMed]
272.
Moreaux
J.
,
Rème
T.
,
Leonard
W.
,
Veyrune
J.L.
,
Requirand
G.
,
Goldschmidt
H.
et al.
(
2012
)
Development of gene expression-based score to predict sensitivity of multiple myeloma cells to DNA methylation inhibitors
.
Mol. Cancer Ther.
11
,
2685
2692
[PubMed]
273.
Maes
K.
,
De Smedt
E.
,
Lemaire
M.
,
De Raeve
H.
,
Menu
E.
,
Van Valckenborgh
E.
et al.
(
2014
)
The role of DNA damage and repair in decitabine-mediated apoptosis in multiple myeloma
.
Oncotarget
5
,
3115
3129
[PubMed]
274.
Maes
K.
,
Menu
E.
,
Van Valckenborgh
E.
,
Van Riet
I.
,
Vanderkerken
K.
and
de Bruyne
E.
(
2013
)
Epigenetic modulating agents as a new therapeutic approach in multiple myeloma
.
Cancers (Basel)
5
,
430
461
[PubMed]
275.
Lavelle
D.
,
DeSimone
J.
,
Hankewych
M.
,
Kousnetzova
T.
and
Chen
Y.H.
(
2003
)
Decitabine induces cell cycle arrest at the G1 phase via p21(WAF1) and the G2/M phase via the p38 MAP kinase pathway
.
Leuk. Res.
27
,
999
1007
[PubMed]
276.
Issa
J.P.J.
,
Garcia-Manero
G.
,
Giles
F.J.
,
Mannari
R.
,
Thomas
D.
,
Faderl
S.
et al.
(
2004
)
Phase 1 study of low-dose prolonged exposure schedules of the hypomethylating agent 5-aza-2’-deoxycytidine (decitabine) in hematopoietic malignancies
.
Blood
103
,
1635
1640
[PubMed]
277.
Chuang
J.C.
,
Warner
S.L.
,
Vollmer
D.
,
Vankayalapati
H.
,
Redkar
S.
,
Bearss
D.J.
et al.
(
2010
)
S110, a 5-Aza-2’-deoxycytidine-containing dinucleotide, is an effective DNA methylation inhibitor in vivo and can reduce tumor growth
.
Mol. Cancer Ther.
9
,
1443
1450
[PubMed]
278.
Plummer
R.
,
Vidal
L.
,
Griffin
M.
,
Lesley
M.
,
De Bono
J.
,
Coulthard
S.
et al.
(
2009
)
Phase I study of MG98, an oligonucleotide antisense inhibitor of human DNA methyltransferase 1, given as a 7-day infusion in patients with advanced solid tumors
.
Clin. Cancer Res.
15
,
3177
3183
,
https://pubmed.ncbi.nlm.nih.gov/19383817/
[PubMed]
279.
Chen
D.Q.
,
Pan
B.Z.
,
Huang
J.Y.
,
Zhang
K.
,
Cui
S.Y.
,
De
W.
et al.
(
2014
)
HDAC 1/4-mediated silencing of microRNA-200b promotes chemoresistance in human lung adenocarcinoma cells
.
Oncotarget
5
,
3333
3349
[PubMed]
280.
Ridinger
J.
,
Koeneke
E.
,
Kolbinger
F.R.
,
Koerholz
K.
,
Mahboobi
S.
,
Hellweg
L.
et al.
(
2018
)
Dual role of HDAC10 in lysosomal exocytosis and DNA repair promotes neuroblastoma chemoresistance
.
Sci. Rep.
8
,
10039
[PubMed]
281.
West
A.C.
and
Johnstone
R.W.
(
2014
)
New and emerging HDAC inhibitors for cancer treatment
.
J. Clin. Invest.
124
,
30
39
[PubMed]
282.
Bose
P.
,
Dai
Y.
and
Grant
S.
(
2014
)
Histone deacetylase inhibitor (HDACI) mechanisms of action: emerging insights
.
Pharmacol. Ther.
143
,
323
336
[PubMed]
283.
Mitsiades
N.
,
Mitsiades
C.S.
,
Richardson
P.G.
,
McMullan
C.
,
Poulaki
V.
,
Fanourakis
G.
et al.
(
2003
)
Molecular sequelae of histone deacetylase inhibition in human malignant B cells
.
Blood
101
,
4055
4062
[PubMed]
284.
Rajkumar
S.V.
(
2011
)
Treatment of multiple myeloma
.
Nat. Rev. Clin. Oncol.
8
,
479
491
[PubMed]
285.
Turgeon
M.O.
,
Perry
N.J.S.
and
Poulogiannis
G.
(
2018
)
DNA damage, repair, and cancer metabolism
.
Front. Oncol.
8
,
15
[PubMed]
286.
Mueller
S.
,
Yang
X.
,
Sottero
T.L.
,
Gragg
A.
,
Prasad
G.
,
Polley
M.Y.
et al.
(
2011
)
Cooperation of the HDAC inhibitor vorinostat and radiation in metastatic neuroblastoma: efficacy and underlying mechanisms
.
Cancer Lett.
306
,
223
229
[PubMed]
287.
Su
J.M.
,
Li
X.N.
,
Thompson
P.
,
Ou
C.N.
,
Ingle
A.M.
,
Russell
H.
et al.
(
2011
)
Phase 1 study of valproic acid in pediatric patients with refractory solid or CNS tumors: a children's oncology group report
.
Clin. Cancer Res.
17
,
589
597
[PubMed]
288.
Mohammed
T.A.
,
Holen
K.D.
,
Jaskula‐Sztul
R.
,
Mulkerin
D.
,
Lubner
S.J.
,
Schelman
W.R.
et al.
(
2011
)
A pilot phase II study of valproic acid for treatment of low-grade neuroendocrine carcinoma
.
Oncologist
16
,
835
843
[PubMed]
289.
Wheler
J.J.
,
Janku
F.
,
Falchook
G.S.
,
Jackson
T.L.
,
Fu
S.
,
Naing
A.
et al.
(
2014
)
Phase I study of anti-VEGF monoclonal antibody bevacizumab and histone deacetylase inhibitor valproic acid in patients with advanced cancers
.
Cancer Chemother. Pharmacol.
73
,
495
501
[PubMed]
290.
Chu
B.F.
,
Karpenko
M.J.
,
Liu
Z.
,
Aimiuwu
J.
,
Villalona-Calero
M.A.
,
Chan
K.K.
et al.
(
2013
)
Phase I study of 5-aza-2’-deoxycytidine in combination with valproic acid in non-small-cell lung cancer
.
Cancer Chemother. Pharmacol.
71
,
115
121
[PubMed]
291.
Iwahashi
S.
,
Utsunomiya
T.
,
Imura
S.
,
Morine
Y.
,
Ikemoto
T.
,
Arakawa
Y.
et al.
(
2014
)
Effects of valproic acid in combination with S-1 on advanced pancreatobiliary tract cancers: clinical study phases I/II
.
Anticancer Res.
34
,
5187
5191
[PubMed]
292.
Piekarz
R.L.
,
Frye
R.
,
Prince
H.M.
,
Kirschbaum
M.H.
,
Zain
J.
,
Allen
S.L.
et al.
(
2011
)
Phase 2 trial of romidepsin in patients with peripheral T-cell lymphoma
.
Blood
117
,
5827
5834
,
https://pubmed.ncbi.nlm.nih.gov/21355097/
[PubMed]
293.
Mackay
H.J.
,
Hirte
H.
,
Colgan
T.
,
Covens
A.
,
MacAlpine
K.
,
Grenci
P.
et al.
(
2010
)
Phase II trial of the histone deacetylase inhibitor belinostat in women with platinum resistant epithelial ovarian cancer and micropapillary (LMP) ovarian tumours
.
Eur. J. Cancer
46
,
1573
1579
,
https://pubmed.ncbi.nlm.nih.gov/20304628/
[PubMed]
294.
Thomas
A.
,
Rajan
A.
,
Szabo
E.
,
Tomita
Y.
,
Carter
C.A.
,
Scepura
B.
et al.
(
2014
)
A phase I/II trial of belinostat in combination with cisplatin, doxorubicin, and cyclophosphamide in thymic epithelial tumors: a clinical and translational study
.
Clin. Cancer Res.
20
,
5392
5402
,
https://pubmed.ncbi.nlm.nih.gov/25189481/
[PubMed]
295.
Rugo
H.S.
,
Jacobs
I.
,
Sharma
S.
,
Scappaticci
F.
,
Paul
T.A.
,
Jensen-Pergakes
K.
et al.
(
2020
)
The promise for histone methyltransferase inhibitors for epigenetic therapy in clinical oncology: a narrative review
.
Adv. Ther.
37
,
3059
3082
,
https://pubmed.ncbi.nlm.nih.gov/32445185/
[PubMed]
296.
Gardner
E.E.
,
Lok
B.H.
,
Schneeberger
V.E.
,
Desmeules
P.
,
Miles
L.A.
,
Arnold
P.K.
et al.
(
2017
)
Chemosensitive relapse in small cell lung cancer proceeds through an EZH2-SLFN11 axis
.
Cancer Cell
31
,
286
299
,
https://pubmed.ncbi.nlm.nih.gov/28196596/
[PubMed]
297.
Kawano
S.
,
Grassian
A.R.
,
Tsuda
M.
,
Knutson
S.K.
,
Warholic
N.M.
,
Kuznetsov
G.
et al.
(
2016
)
Preclinical evidence of anti-tumor activity induced by EZH2 inhibition in human models of synovial sarcoma
.
PLoS ONE
11
,
e0158888
https://pubmed.ncbi.nlm.nih.gov/27391784/
298.
Knutson
S.K.
,
Kawano
S.
,
Minoshima
Y.
,
Warholic
N.M.
,
Huang
K.C.
,
Xiao
Y.
et al.
(
2014
)
Selective inhibition of EZH2 by EPZ-6438 leads to potent antitumor activity in EZH2-mutant non-Hodgkin lymphoma
.
Mol. Cancer Ther.
13
,
842
854
,
https://pubmed.ncbi.nlm.nih.gov/24563539/
[PubMed]
299.
Kung
P.P.
,
Bingham
P.
,
Brooun
A.
,
Collins
M.
,
Deng
Y.L.
,
Dinh
D.
et al.
(
2018
)
Optimization of orally bioavailable enhancer of zeste homolog 2 (EZH2) inhibitors using ligand and property-based design strategies: identification of development candidate (R)-5,8-Dichloro-7-(methoxy(oxetan-3-yl)methyl)-2-((4-methoxy-6-methyl-2-oxo-1,2-dihydropyridin-3-yl)methyl)-3,4-dihydroisoquinolin-1(2H)-one (PF-06821497)
.
J. Med. Chem.
61
,
650
665
,
https://pubmed.ncbi.nlm.nih.gov/29211475/
[PubMed]
300.
McCabe
M.T.
,
Ott
H.M.
,
Ganji
G.
,
Korenchuk
S.
,
Thompson
C.
,
Van Aller
G.S.
et al.
(
2012
)
EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations
.
Nature
492
,
108
112
,
https://pubmed.ncbi.nlm.nih.gov/23051747/
[PubMed]
301.
Daigle
S.R.
,
Olhava
E.J.
,
Therkelsen
C.A.
,
Basavapathruni
A.
,
Jin
L.
,
Boriack-Sjodin
P.A.
et al.
(
2013
)
Potent inhibition of DOT1L as treatment of MLL-fusion leukemia
.
Blood
122
,
1017
1025
,
https://pubmed.ncbi.nlm.nih.gov/23801631/
[PubMed]
302.
Poulard
C.
,
Corbo
L.
and
Le Romancer
M.
(
2016
)
Protein arginine methylation/demethylation and cancer
.
Oncotarget
7
,
67532
67550
,
https://pubmed.ncbi.nlm.nih.gov/27556302/
[PubMed]
303.
Cheng
H.
,
Qin
Y.
,
Fan
H.
,
Su
P.
,
Zhang
X.
,
Zhang
H.
et al.
(
2013
)
Overexpression of CARM1 in breast cancer is correlated with poorly characterized clinicopathologic parameters and molecular subtypes
.
Diagn. Pathol.
8
,
129
,
https://pubmed.ncbi.nlm.nih.gov/23915145/
304.
Hong
H.
,
Kao
C.
,
Jeng
M.H.
,
Eble
J.N.
,
Koch
M.O.
,
Gardner
T.A.
et al.
(
2004
)
Aberrant expression of CARM1, a transcriptional coactivator of androgen receptor, in the development of prostate carcinoma and androgen-independent status
.
Cancer
101
,
83
89
,
https://pubmed.ncbi.nlm.nih.gov/15221992/
[PubMed]
305.
Chan-Penebre
E.
,
Kuplast
K.G.
,
Majer
C.R.
,
Boriack-Sjodin
P.A.
,
Wigle
T.J.
,
Johnston
L.D.
et al.
(
2015
)
A selective inhibitor of PRMT5 with in vivo and in vitro potency in MCL models
.
Nat. Chem. Biol.
11
,
432
437
,
https://pubmed.ncbi.nlm.nih.gov/25915199/
[PubMed]
306.
Pérez-Salvia
M.
and
Esteller
M.
(
2017
)
Bromodomain inhibitors and cancer therapy: from structures to applications
.
Epigenetics
12
,
323
339
,
https://pubmed.ncbi.nlm.nih.gov/27911230/
[PubMed]
307.
Shorstova
T.
,
Foulkes
W.D.
and
Witcher
M.
(
2021
)
Achieving clinical success with BET inhibitors as anti-cancer agents
.
Br. J. Cancer
124
,
1478
1490
,
https://pubmed.ncbi.nlm.nih.gov/33723398/
308.
Berthon
C.
,
Raffoux
E.
,
Thomas
X.
,
Vey
N.
,
Gomez-Roca
C.
,
Yee
K.
et al.
(
2016
)
Bromodomain inhibitor OTX015 in patients with acute leukaemia: a dose-escalation, phase 1 study
.
Lancet Haematol.
3
,
e186
e195
,
https://pubmed.ncbi.nlm.nih.gov/27063977/
[PubMed]
309.
Amorim
S.
,
Stathis
A.
,
Gleeson
M.
,
Iyengar
S.
,
Magarotto
V.
,
Leleu
X.
et al.
(
2016
)
Bromodomain inhibitor OTX015 in patients with lymphoma or multiple myeloma: a dose-escalation, open-label, pharmacokinetic, phase 1 study
.
Lancet Haematol.
3
,
e196
e204
,
https://pubmed.ncbi.nlm.nih.gov/27063978/
[PubMed]
310.
Piha-Paul
S.A.
,
Hann
C.L.
,
French
C.A.
,
Cousin
S.
,
Braña
I.
,
Cassier
P.A.
et al.
(
2019
)
Phase 1 Study of Molibresib (GSK525762), a bromodomain and extra-terminal domain protein inhibitor, in NUT carcinoma and other solid tumors
.
JNCI Cancer Spectr.
4
,
pkz093
,
https://pubmed.ncbi.nlm.nih.gov/32328561/
[PubMed]
311.
Pollyea
D.A.
,
Stevens
B.M.
,
Jones
C.L.
,
Winters
A.
,
Pei
S.
,
Minhajuddin
M.
et al.
(
2018
)
Venetoclax with azacitidine disrupts energy metabolism and targets leukemia stem cells in patients with acute myeloid leukemia
.
Nat. Med.
24
,
1859
1866
,
https://pubmed.ncbi.nlm.nih.gov/30420752/
[PubMed]
312.
Wu
M.
,
Sheng
L.
,
Cheng
M.
,
Zhang
H.
,
Jiang
Y.
,
Lin
S.
et al.
(
2019
)
Low doses of decitabine improve the chemotherapy efficacy against basal-like bladder cancer by targeting cancer stem cells
.
Oncogene
38
,
5425
5439
,
https://pubmed.ncbi.nlm.nih.gov/30918330/
[PubMed]
313.
Adhikary
G.
,
Grun
D.
,
Balasubramanian
S.
,
Kerr
C.
,
Huang
J.M.
and
Eckert
R.L.
(
2015
)
Survival of skin cancer stem cells requires the Ezh2 polycomb group protein
.
Carcinogenesis
36
,
800
810
,
https://pubmed.ncbi.nlm.nih.gov/25969142/
[PubMed]
314.
Yu
T.
,
Wang
Y.
,
Hu
Q.
,
Wu
W.N.
,
Wu
Y.
,
Wei
W.
et al.
(
2017
)
The EZH2 inhibitor GSK343 suppresses cancer stem-like phenotypes and reverses mesenchymal transition in glioma cells
.
Oncotarget
8
,
98348
98359
,
https://pubmed.ncbi.nlm.nih.gov/29228694/
[PubMed]
315.
Tian
T.
,
Guo
T.
,
Zhen
W.
,
Zou
J.
and
Li
F.
(
2020
)
BET degrader inhibits tumor progression and stem-like cell growth via Wnt/β-catenin signaling repression in glioma cells
.
Cell Death Dis.
11
,
900
,
https://pubmed.ncbi.nlm.nih.gov/33093476/
316.
Lee
S.H.
,
Nam
H.J.
,
Kang
H.J.
,
Samuels
T.L.
,
Johnston
N.
and
Lim
Y.C.
(
2015
)
Valproic acid suppresses the self-renewal and proliferation of head and neck cancer stem cells
.
Oncol. Rep.
34
,
2065
2071
,
https://pubmed.ncbi.nlm.nih.gov/26239260/
[PubMed]
317.
Nalls
D.
,
Tang
S.N.
,
Rodova
M.
,
Srivastava
R.K.
and
Shankar
S.
(
2011
)
Targeting epigenetic regulation of miR-34a for treatment of pancreatic cancer by inhibition of pancreatic cancer stem cells
.
PLoS ONE
6
,
e24099
https://pubmed.ncbi.nlm.nih.gov/21909380/
318.
Cai
M.H.
,
Xu
X.G.
,
Yan
S.L.
,
Sun
Z.
,
Ying
Y.
,
Wang
B.K.
et al.
(
2018
)
Depletion of HDAC1, 7 and 8 by histone deacetylase inhibition confers elimination of pancreatic cancer stem cells in combination with gemcitabine
.
Sci. Rep.
8
,
162
,
https://pubmed.ncbi.nlm.nih.gov/29374219/
319.
Villanueva
L.
,
Álvarez-Errico
D.
and
Esteller
M.
(
2020
)
The contribution of epigenetics to cancer immunotherapy
.
Trends Immunol.
41
,
676
691
https://pubmed.ncbi.nlm.nih.gov/32622854/
[PubMed]
320.
Liu
M.
,
Thomas
S.L.
,
DeWitt
A.K.
,
Zhou
W.
,
Madaj
Z.B.
,
Ohtani
H.
et al.
(
2018
)
Dual inhibition of DNA and histone methyltransferases increases viral mimicry in ovarian cancer cells
.
Cancer Res.
78
,
5754
5760
https://pubmed.ncbi.nlm.nih.gov/30185548/
[PubMed]
321.
Daskalakis
M.
,
Brocks
D.
,
Sheng
Y.H.
,
Islam
M.S.
,
Ressnerova
A.
,
Assenov
Y.
et al.
(
2018
)
Reactivation of endogenous retroviral elements via treatment with DNMT- and HDAC-inhibitors
.
Cell Cycle
17
,
811
822
https://pubmed.ncbi.nlm.nih.gov/29633898/
[PubMed]
322.
Roulois
D.
,
Loo Yau
H.
,
Singhania
R.
,
Wang
Y.
,
Danesh
A.
,
Shen
S.Y.
et al.
(
2015
)
DNA-demethylating agents target colorectal cancer cells by inducing viral mimicry by endogenous transcripts
.
Cell
162
,
961
973
https://pubmed.ncbi.nlm.nih.gov/26317465/
[PubMed]
323.
Yang
X.P.
,
Jiang
K.
,
Hirahara
K.
,
Vahedi
G.
,
Afzali
B.
,
Sciume
G.
et al.
(
2015
)
EZH2 is crucial for both differentiation of regulatory T cells and T effector cell expansion
.
Sci. Rep.
5
,
10643
https://pubmed.ncbi.nlm.nih.gov/26090605/
324.
Wang
D.
,
Quiros
J.
,
Mahuron
K.
,
Pai
C.C.
,
Ranzani
V.
,
Young
A.
et al.
(
2018
)
Targeting EZH2 reprograms intratumoral regulatory T cells to enhance cancer immunity
.
Cell Rep
23
,
3262
3274
,
[PubMed]

Author notes

*

These authors contributed equally to this work.

This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).