Next Article in Journal
Phosphorous Magnetic Resonance Spectroscopy and Molecular Markers in IDH1 Wild Type Glioblastoma
Next Article in Special Issue
Interactions Networks for Primary Heart Sarcomas
Previous Article in Journal
Consistent Major Differences in Sex- and Age-Specific Diagnostic Performance among Nine Faecal Immunochemical Tests Used for Colorectal Cancer Screening
Previous Article in Special Issue
Evaluation of Collagen Alterations in Early Precursor Lesions of High Grade Serous Ovarian Cancer by Second Harmonic Generation Microscopy and Mass Spectrometry
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

HDAC Inhibitors: Dissecting Mechanisms of Action to Counter Tumor Heterogeneity

by
Dimitris Karagiannis
1,* and
Theodoros Rampias
2,*
1
Department of Genetics and Development, Columbia University Medical Center, New York, NY 10032, USA
2
Biomedical Research Foundation of the Academy of Athens, 11527 Athens, Greece
*
Authors to whom correspondence should be addressed.
Cancers 2021, 13(14), 3575; https://doi.org/10.3390/cancers13143575
Submission received: 16 June 2021 / Revised: 9 July 2021 / Accepted: 13 July 2021 / Published: 16 July 2021
(This article belongs to the Special Issue Tumor Evolution: Progression, Metastasis and Therapeutic Response)

Abstract

:

Simple Summary

Tumor heterogeneity promotes the development of drug resistance in cancer. HDAC inhibitors modulate several processes that contribute to intra-tumoral heterogeneity. With careful consideration of the underlying biology, HDAC inhibitors can be utilized to improve therapeutic efficacy.

Abstract

Intra-tumoral heterogeneity presents a major obstacle to cancer therapeutics, including conventional chemotherapy, immunotherapy, and targeted therapies. Stochastic events such as mutations, chromosomal aberrations, and epigenetic dysregulation, as well as micro-environmental selection pressures related to nutrient and oxygen availability, immune infiltration, and immunoediting processes can drive immense phenotypic variability in tumor cells. Here, we discuss how histone deacetylase inhibitors, a prominent class of epigenetic drugs, can be leveraged to counter tumor heterogeneity. We examine their effects on cellular processes that contribute to heterogeneity and provide insights on their mechanisms of action that could assist in the development of future therapeutic approaches.

1. Introduction

1.1. HDAC Inhibitors in Cancer Therapy

Histone deacetylase inhibitors (HDACi) were the first class of epigenetic drugs to be approved for cancer therapy. Currently, four HDACi, vorinostat, belinostat, Panobinostat, and romidepsin, are approved by the FDA for cancer treatment, and several others are under investigation in the clinic (Table 1). Vorinostat is approved for use in refractory cutaneous T-cell lymphoma (CTCL) [1], belinostat for refractory peripheral T-cell lymphoma (PTCL) [2], and romidepsin for CTCL and PTCL patients that have received one prior therapy [3,4]. Panobinostat is approved as a third-line treatment in multiple myeloma (MM) in conjunction with bortezomib and dexamethasone. Since their discovery, a huge effort has been underway to expand their use in cancer therapy. Despite their initial success as monotherapy in these hematological cancers, the current clinical investigations using HDACi have been reshaped to drug combination strategies (reviewed here [5,6,7]). So far, ongoing and completed studies have examined combinations of HDACi with chemotherapy, DNA methylation inhibitors, tyrosine kinase inhibitors, estrogen inhibitors, immunotherapy, etc. Prominent phase III clinical trials include studies on the treatment of hormone receptor-positive and HER2-negative breast carcinoma with Class I-selective HDACi entinostat and tucidinostat in combination with aromatase inhibitor exemestane (NCT02482753 [8], NCT02115282 [9]), and studies examining the activity of vorinostat and VPA in pediatric high-grade glioma in combination with temozolomide (NCT03243461, NCT01236560). In this review we give an overview of the current knowledge on the mechanisms of action of HDAC inhibitors and propose ways they could be leveraged in future clinical applications, in the context of tumor heterogeneity and therapy resistance.
Histone acetylation is an epigenetic modification important for gene expression, genome maintenance, and DNA replication. Several lysine residues in histones 3 (e.g., H3K27 and H3K9) and 4 (e.g., H4K16, H4K12) are acetylated, and are generally associated with a permissive chromatin state. These modifications are deposited by histone acetyltransferase complexes, such as CBP/p300 and GCN5, and removed by histone deacetylases, which include HDAC and Sirtuin (SIRT) enzymes (Figure 1) (reviewed in [19]). Class I HDACs include HDAC 1, 2, 3, and 8, which are ubiquitously expressed, localize primarily in the nucleus, and are components of multiple repressor complexes [20]. Class II HDACs display a tissue-specific pattern of expression, localize in both the nucleus and the cytoplasm, and thus have non-histone deacetylation activity [20]. They are divided into two subclasses: IIa (HDAC4, 5, 7, 9) and IIb (HDAC6, 10) [20]. Class III HDACs include SIRT deacetylases 1–7 and belong to a distinct family of enzymes that are not targeted by the HDACi discussed in this review. Class IV HDACs are comprised solely of HDAC11. Class I, II, and IV histone deacetylases, henceforth referred to as HDACs, rely on a zinc ion to bind their substrate and catalyze deacetylation (Figure 2). Importantly, HDACs deacetylate several other non-histone proteins important in cancer, including p53 and c-Myc [21,22] (Figure 1).
HDAC inhibitors, such as the hydroxamic acids vorinostat (a.k.a SAHA) and trichostatin A (TSA), act by binding to the catalytic site metal Zn2+ ion and thus blocking substrate accessibility [23] (Figure 2). Structurally, most of the HDAC inhibitors follow a general pharmacophore which mimics the natural peptide substrates and consists of: (i) a protein-surface-interacting moiety that occupies the entrance area of the active site (cap group), (ii) a hydrophobic linker that mimics the N-alkyl side chain of lysine, and (iii) a metal binding moiety that interacts with the catalytic site Zn2+ ion [24,25,26,27]. Hydroxamic acids, as metal binding groups, generally lead to strong inhibition. Early structural studies in HDAC-8 revealed the mode of binding of the HDAC inhibitors [23]. As shown in Figure 2, the hydroxylamates are complexed by the Zn2+ ion, and the aliphatic chains of the linker fit into the long hydrophobic tunnel of the active site. The terminal capping group interacts with the external portion of the enzyme and the solvent.
First generation HDAC inhibitors such as vorinostat and trichostatin are generally nonselective, targeting most of the metal-dependent isoforms [12,28]. The high degree of sequence similarity among the channels, active sites, and internal cavities of all Class I HDACs makes the design of isoform-specific inhibitors a very challenging task. However, information obtained by X-ray crystallographic data and enzymatic activity assays led to remarkable progress in the field of isoform-selective compounds for both Class I/II HDACs over recent years [25,29]. As a result, there are significant variations among HDACs’ selectivity (Table 1), which seems to stem from their chemical nature. For instance, in HDAC6, the entrance area on the enzyme’s surface is larger compared to other isoforms, and therefore selectivity for this enzyme has been achieved by incorporating large cap groups and benzyl linkers instead of aliphatic chains [30]. Selectivity of compounds across HDACs has been determined primarily through cell-free enzymatic activity assays of HDAC1, HDAC2, HDAC3, and HDAC6; and only few studies have tested all enzymes [10,11,12,13,15,16,17,18]. Interestingly, HDAC11 was reported to be a target of romidepsin, largazole [11], and entinostat [16]; however, this should be confirmed by additional studies. Overall, the complexity of histone acetylation regulation, HDAC activity, and HDACi selectivity renders this field of study particularly challenging.

1.2. Heterogeneity and Therapy Resistance

High grade tumors are comprised of distinct subclones of cancer cells that arise due to factors such as the tumor micro-environment, stromal interactions, and genomic and epigenomic events. During cancer progression and evolution, spontaneous genetic alterations such as mutations and copy number changes that confer fitness advantages propagate and collectively generate a heterogeneous tumor [31,32]. In addition, a wide range of epigenetic mechanisms, including DNA methylation, chromatin remodeling, and histone modifications, contribute to gene expression diversity and transcriptomic heterogeneity within tumors [33,34].
Tumor heterogeneity is recognized as a major contributor to therapy resistance [35]. Survival and adaptation of cancer cells to therapeutics relies on rewiring of biological processes. Therefore, in heterogeneous tumors, the emergence of resistant subclones is more likely than in homogeneous ones. This is supported by the fact that intra-tumoral heterogeneity stemming from mutations and copy number alterations is positively correlated with chemotherapy resistance [35,36].
Chemotherapy has remained a primary treatment for a wide variety of cancers for many decades; however, intrinsic (present at baseline) or acquired (developed after initial treatment) resistance during treatment cycles and high toxicity are major obstacles that severely limit patients’ clinical benefits [37]. Chemotherapy drugs act mainly by inhibiting processes required for cell proliferation, such as DNA replication, genome integrity, and mitosis [38]. Examples of chemotherapy include anti-folates, such as methotrexate, which causes thymidine nucleotide depletion and DNA synthesis inhibition; nucleoside analogues such as gemcitabine, which inhibits DNA synthesis and repair machinery; vinca alkaloids, which inhibit microtubule polymerization and cell division; taxanes such as paclitaxel, an antimitotic compound that promotes microtubule assembly; camptothecins and anthracyclines, which inhibit topoisomerases I and II, respectively; and platinum compounds such as cisplatin and carboplatin, which are DNA-crosslinking agents that form inter-strand crosslinks (ICLs) which induce DNA damage and inhibit DNA replication [38,39]. Radiotherapy in genotoxic cancer therapy causes a broad spectrum of DNA damage types, including double-strand breaks (DSBs), single-strand breaks (SSBs), and oxidized nucleotide adducts such as 8-oxoguanine via the formation of reactive oxygen species (ROS) [40].
Generally, cancer cells become resistant to chemotherapy through alterations in drug transport and metabolism, modulation or mutation of drug targets, and genetic rewiring to bypass or compensate for the targeted pathways [41,42,43]. For example, a well characterized mechanism of methotrexate resistance is caused by a defect in the reduced folate carrier (RFC/SLC19A1) which leads to reduced methotrexate uptake [44,45]. In the case of 5-fluorouracil, resistance has been shown to occur by overexpression of its target, thymidylate synthase [46,47]. For DNA-damaging agents such as cisplatin, increases in the cells’ capacity to repair DNA lesions, such as overexpression of genes involved in the nucleotide excision repair (NER) pathway, have been strongly linked with chemoresistance [48,49].
Analogous resistance mechanisms have been described for targeted therapeutic approaches as well. In endocrine therapy, a widely used treatment for hormone receptor-positive (HR+) breast cancer [50], several mechanisms of resistance have been identified, including loss of estrogen receptor α (ERα) expression, mutations and post-translational modifications of ERα, deregulation of ERα co-activators, and mutations in cancer-associated pathways, such as the cell cycle, tyrosine kinase signaling, and apoptosis [51]. As a result, subclones with advantageous alterations, such as defects in cell cycle checkpoints [52], can cause recurrence. Analogous resistance mechanisms have been described in response to hormone therapy of prostate cancer with anti-androgen inhibitors [53,54]. Patients carrying epidermal growth factor receptor (EGFR) mutations often develop resistance to EGFR tyrosine kinase inhibitors, through selection of subclones with mutations that activate signaling pathways that are parallel to or downstream from EGFR, such as secondary mutations in EGFR or amplification of the MET receptor tyrosine kinase locus [55,56]. In BRCA-mutant breast and ovarian cancer patients, PARP inhibitors are utilized to exploit the inherent deficiency of these tumors in homologous recombination repair [57]. In this targeted therapy, resistance or recurrence can result from subsets of cells with restored BRCA functionality, or with rewiring of DNA repair machinery to restore homologous recombination repair (HRR) function or replication fork protection [58].

1.3. Epigenetic Drugs to Counter Tumor Heterogeneity and Overcome Resistance

Epigenetic deregulation is a hallmark of cancer and has a major contribution to disease development and progression [33]. Transcriptional heterogeneity within tumors has often been correlated with the identification of gene expression profiles and epigenomic signatures that are linked to de-differentiation processes or to poor differentiation states that are typical among embryonic and cancer stem cell populations [59,60,61,62]. These findings suggest that the epigenetic landscape in cancer cells may drive reprogramming and compromise cellular differentiation processes. Therefore, loss of cellular identity by epigenetic deregulation in tumor subclones and cancer stem cell populations (CSCs) within tumors contributes to transcriptional heterogeneity. For instance, single-cell RNA-seq studies in glioblastoma and oligodendroglioma have demonstrated that brain tumors contain subpopulations of undifferentiated cells with stem cell signatures and subsets of cells that have undergone neural differentiation [63,64]. Since the epigenome is important for stem cell status, epigenetic dysregulation during the initial steps of carcinogenesis may also be responsible for the emergence of CSCs from tissue stem cells.
Recent studies provide a direct link between altered epigenome and drug resistance. For instance, KDM5A/JARID1A epigenetically-driven resistance to EGFR tyrosine kinase inhibitors has been reported for non-small cell lung cancer (NSCLC) [65]. Similarly, an adaptive chromatin remodeling mechanism based on the activity of KDM5A/JARID1A has been described to drive resistance to kinase inhibitors in glioblastoma CSCs [66].
Reversibility is a key feature of epigenetic modifications, and therefore targeting the dependence of CSCs and other tumor cells on specific epigenetic factors offers a therapeutic opportunity. In this direction, numerous studies have shown that pharmacological inhibition of epigenetic factors and other chromatin remodeler proteins can inhibit the expression of oncoproteins that maintain the CSC identities or the altered transcriptomes of specific tumor subclones and promote their effective elimination [67,68,69]. In addition, single-cell transcriptomic and epigenomic analyses have shown that tumor cells undergo heterogeneous adaptive responses to chemotherapy, giving rise to independent mechanisms of resistance [70]. Taken together, these findings indicate that epigenetic regulation is a core aspect of resistance emergence. As a result, targeting the epigenome could be a therapeutic strategy to mitigate mechanisms of resistance. Along these lines, Hinohara and colleagues showed that inhibition of KDM5 could mitigate anti-estrogen resistance by inhibiting transcriptional heterogeneity [71]. In addition, several lines of research suggest that epigenetic drugs can re-sensitize tumors to chemotherapy and other types of treatment (reviewed here [72,73]).

2. Effects of HDAC Inhibitors and Therapeutic Implications for Tumor Heterogeneity

Extensive research on HDACi has found that they impact most cancer-related pathways. To date, HDACi have been implicated in the regulation of numerous cellular processes, including chromatin regulation, gene expression, apoptosis, cell cycle progression, genome maintenance, DNA repair, metabolism, phenotypic plasticity, and aspects of the tumor micro-environment. This is due to the broad and complex functions of the HDAC enzymes. Through deacetylation of histones and proteins, these enzymes regulate gene expression, chromatin structure, genome replication and maintenance, and several other cellular pathways. Understanding how HDACi mediate each of their effects will be important for clinical application of these inhibitors. Rational use of HDAC inhibitors, such as in conjunction with other treatments or in selected patients, could be leveraged to reduce tumor heterogeneity and thus mitigate tumor resistance and recurrence mechanisms (Figure 3, Table 2).
Chromatin–Gene Expression. Histone acetylation is an integral part of chromatin regulation and is generally associated with accessible and transcriptionally active chromatin (Figure 1). Although inhibition of HDAC activity leads to increases in the global levels of histone acetylation, it does not lead to corresponding increases in gene transcription or chromatin accessibility. This is likely because the genomic distribution and recruitment of transcription factors and chromatin remodelers do not follow the same pattern.
Upon HDAC inhibition, histone acetylation does not increase uniformly. For example, in colon cancer cells, TSA and sodium butyrate cause a global increase in histone acetylation but loss of H3ac and H4ac at transcription start sites [76]. In a study where the authors evaluated the genomic distribution of H4 acetylation after vorinostat treatment, they observed a preferential increase in gene bodies [77]. Notably, this could be due to the preferential localization of the chemical probe, which has been shown to happen in the case of vorinostat [78]. In addition, several studies have noted that the pre-existing state of chromatin influences local response of chromatin to HDACi [77,79].
In the context of cancer therapy, several studies have investigated whether disrupting regulation of chromatin by HDAC inhibition has tumor-killing effects. In a model of melanoma progression, advanced melanoma displayed loss of histone acetylation in specific loci, such as cancer-associated gene promoters, and vorinostat and entinostat were found to reverse deacetylated sites [80]. HDAC inhibition was more cytotoxic in advanced melanoma, and in melanoma cell lines with reduced H3K27ac at these loci. Notably, it was not determined whether this was due to restoration of gene expression. Early studies on gene expression identified induction of apoptosis-related genes. In a CTCL clinical trial, panobinostat induced histone acetylation and the expression of p21 and other apoptosis-related genes, and downregulated proliferation-associated genes as early as 4 h after treatment [81]. Vorinostat and romidepsin were found to induce pro-apoptotic gene expression within 10 hours in a similar manner [82]. These transcriptional changes could potentially induce cell death, but it is still undetermined whether they are sufficient, and whether they are primary effects of HDAC inhibition.
An emerging concept of chromatin dysregulation in cancer is the sustenance of oncogene transcription by super-enhancers [83]. Super-enhancers, which are large clusters of transcriptional enhancers, are exploited by cancer cells to sustain oncogenic pathway activation, such as MYC. As a result, cancer cells are sensitive to disruption of super-enhancers by epigenetic drugs such as bromodomain (BET) inhibitors. Recently, several studies have reported super-enhancer disruption by HDAC inhibition. Using a systematic chemical screen, Gryder et al. identified HDAC1/2/3 as essential for core regulatory transcription, which is governed by super-enhancer activity [84]. In glioma, several HDAC inhibitors were found to reduce H3K27ac in super-enhancers of oncogenes such as MYC and PIK3C2B [85,86]. One study suggests that this is due to the redistribution of BRD4, an acetylation-binding protein that mediates super-enhancer assembly [87], to gene bodies [77]. In line with this, Kim and colleagues used global run-on sequencing (GRO-seq), a method of high-throughput analysis of nascent transcription, and showed that HDACi represses transcription elongation, thereby selectively inhibiting transcription of highly-expressed genes [88]. In contrast, another study on immediate transcriptional effects found positive effect of TSA on elongation after 10 minutes of treatment [89]. In pancreatic adenocarcinoma, HDACi inhibited TGFβ-regulated gene expression through deactivation of enhancers, and induced expression of MYC and BRD4-bound genes [90]. Overall, although there is conflicting evidence regarding how HDACi affect chromatin and gene expression, it is evident that this is caused at least partly through deregulation of enhancers.
Therapeutic Implications: Transcriptional dysregulation is often a driver of tumor progression, and its disruption can have therapeutic benefit. Thus, HDAC inhibitors could potentially be used to target tumors that rely on the expression of oncogenes such as c-Myc and EGFR. Additionally, transcriptional dysregulation stemming from sub-clonal events, such as genomic rearrangements (e.g., amplifications) or stochastic epigenetic events (e.g., loss of repressive state), is a significant source of intra-tumoral heterogeneity. Markedly, transcriptional heterogeneity has been associated with therapy resistance in breast cancer [71]; and in EGFR mutant NSCLC, MET overexpression is associated with resistance to tyrosine kinase inhibitors(TKI) [91,92]. Thus, HDAC inhibitors have the potential to counter heterogeneity in transcription and resistance by disrupting clonal gene overexpression—such as that of the c-Myc oncogene [86]—and the transcription of genes induced by regional signals in the tumor microenvironment, such as TGFβ [90,93]. Notably, a clinical trial in NSCLC patients examining the combination of entinostat with the EGFR inhibitor erlotinib in comparison to erlotinib alone found a small increase in progression-free survival (PFS) [94]. Currently, another trial is exploring the efficacy of the HDAC and EGFR inhibitor combination in treating TKI-resistant lung cancer (NCT02151721).
Genome maintenance—DNA repair. Genome stability and repair rely on strict regulation of chromatin remodeling, accessibility, and histone and DNA modifications, including histone acetylation [95]. HDACs have been shown to be indispensable for genome maintenance and DNA repair. Bhaskara et al. showed that HDAC3 is important for DNA replication, maintenance of chromatin structure, genome stability, and DNA repair by non-homologous end joining (NHEJ) and homologous recombination repair (HRR) [96,97]. These functions were attributed to modulations of global levels of histone modifications, such as H3 and H4 acetylation and H3K9me3, which are important for DNA repair, DNA replication, and heterochromatin maintenance, respectively. In addition, HDAC3 was found to be important for sister chromatid cohesion through regulation of H3K4ac [98]. HDAC1 and 2 have also been shown to be necessary for DNA replication and NHEJ [99,100]. Class II HDACs such as HDAC4 have also been implicated in DNA repair [101].
As expected, HDAC inhibitors have been found to disrupt both genome stability and DNA repair, and several underlying mechanisms have been proposed. One study reported that vorinostat induced DNA damage through stalling of replication forks, and HDAC3 knock-down had similar effects [102]. This was attributed to aberrant replication origin firing through opening of chromatin by HDACi. Moreover, HDACi were shown to radiosensitize cells by downregulating DNA repair genes [103]. Several studies have reported that inhibition of HDACs leads to downregulation HRR components, potentially through decreased E2F1 recruitment [104,105]. In addition, HDACi were shown to modulate DNA repair by direct deacetylation of the DNA repair proteins Ku70 and PARP1 [106] (Figure 4). In leukemic cells, HDACi induced H2A.X phosphorylation 3 min after treatment, indicating the presence of DNA double-strand breaks [107]. A few studies have attributed this to the generation of ROS [108,109]. Entinostat induced ROS through loss of mitochondrial membrane potential at high concentrations as early as 2 hours after treatment [110]. In summary, it is evident that HDAC inhibitors disrupt genome stability, but the contributions of the various implicated mechanisms are unclear.
Therapeutic Implications. Modulation of DNA repair by HDAC inhibitors has generated great interest in cancer treatment. Several studies have shown that HDACi sensitize cells to irradiation and DNA-damaging agents [103,104]. It has been suggested that this is due to the modulation of DNA repair proteins involved in ICL repair, such as downregulation of BRCA1, Rad51, and FANCD2 [111]. Therefore, HDACi could suppress the emergence of subclones proficient at DNA repair during chemotherapy and mitigate resistance. For instance, in a heterogeneous BRCA-mutant tumor, resistance can emerge via reversion to DNA-repair defects [112]. Several clinical trials are currently evaluating this combinatorial approach, for example, in patients with advanced lymphoma (NCT01796002), glioma (NCT00268385), and solid tumors (NCT00246103).
In addition, impairment of HRR by HDACi has been found to confer sensitivity to PARP inhibition, a known characteristic of HRR-deficient cells [113]. This approach also counters the inherent heterogeneity of tumor cells in their ability to employ HRR and should therefore lead to better and sustained responses to PARP inhibitors. This approach is currently under evaluation in the clinic for ovarian, primary peritoneal, fallopian tube, and breast cancers (NCT03924245, NCT03742245). Notably, suppression of DNA repair can be detrimental for genomically unstable cells, which rely on it for repair of endogenous DNA damage. Therefore, the contributions of subclones with genomic instability to tumor heterogeneity could be eradicated by utilizing HDAC inhibitors.
Cell cycle. A major effect of HDAC inhibitors in cells is activation of tumor suppressor pathways and blocking of cell cycle progression. Cell cycle progression is tightly regulated to ensure homeostasis, and its obstruction can lead to senescence, apoptosis, and other forms of cell death. The activities of major tumor suppressors such as p53 and p21 are tightly linked to cell cycle regulation [114].
A number of studies have shown that HDAC inhibition induces p21 expression and cell cycle arrest [115,116]. This could be due to HDAC1 and HDAC2 inhibition, both of which have been shown to bind and repress p21 transcription [117,118], although it is not clear whether this regulation is direct. Inhibition of p53 deacetylation by HDACi has also been implicated as a potential mechanism [119]. Therefore, it has been suggested that the cell-killing property of HDAC inhibitors is mediated by transcriptional de-repression of the p21 locus. However, this hypothesis is in stark contrast to two other studies where p21 was found to confer a cytoprotective rather than cytotoxic effect in cells treated with HDAC inhibitors [110,120].
Further studies on the effects of HDAC inhibitors on the cell cycle indicated that they induce activation of the G2/M checkpoint, which has been suggested to explain why—similarly to chemotherapy—tumor cells are more sensitive to these agents in comparison to normal tissues [121,122]. HDAC3 directly regulates cell cycle progression and cyclin A degradation by modulating cyclin A acetylation [123], which points to a direct mechanism behind this effect. Overall, it is evident that HDAC inhibitors impede cell cycle progression through direct and/or indirect means, but a clear mechanistic link showing which aspects of HDAC inhibition are necessary and sufficient to block the cell cycle and induce cell death has yet to be uncovered.
Therapeutic Implications: An important aspect of tumor chemoresistance through heterogeneity is cell division. For example, it has been well characterized that more slowly proliferating tumor subpopulations are more resistant to chemotherapy [124]. Several studies and clinical trials have assessed combinations of HDACi with chemotherapies, such as topoisomerase inhibitors [125,126,127,128] and platinum-based compounds [108,129,130], and reported synergistic effects. This suggests that either the effect of HDACi on the cell cycle is not sufficient to promote chemoresistance or that the beneficial effects of HDACi, such as suppression of DNA repair, outweigh this drawback. In addition, HDACi could be a way to reduce cell cycle-heterogeneity in tumors and achieve robust responses to treatment. Computational models suggest that when cell cycle-heterogeneous tumors are subjected to high levels of cell death during mitosis, such as during chemotherapy treatment, slow-cycling resistant subpopulations become more prevalent [131,132]. Therefore, HDACi could suppress this selection process and block the emergence of resistant subpopulations. Further research is needed to evaluate these strategies.
Cell cycle deregulation has been identified as a resistance mechanism in endocrine therapy of ER+ breast cancer [51]. Interestingly, preclinical studies in ER+ breast cancer have associated HDAC inhibitors with suppression of ER signaling, including cyclin D signaling [133], and re-sensitization of endocrine therapy-resistant cells [134]. The combination of HDACi and the aromatase inhibitor exemestane in ER+ breast cancer patients with resistance to prior endocrine therapy provided promising results in phase II clinical trials and is currently being evaluated in two phase III trials [8,9].
Cell Death. Studies on HDAC inhibitor-mediated cell death showed consistent induction of intrinsic and/or extrinsic pathways of apoptosis. This type of cell death is induced by extracellular signals such as tumor necrosis factor α (TNFα) or endogenous signals such as DNA damage. Signaling cascades lead to the activation of members of the caspase family, which then target numerous proteins for proteolysis [135]. The characteristics and determinants of HDACi-mediated apoptosis that were studied varied between reports. Specifically, HDAC inhibitors have been shown to induce caspase-mediated apoptosis [136]. Meanwhile, several other studies have provided a mechanistic link between HDACi driven apoptosis and the generation of reactive oxygen species (ROS) through mitochondrial membrane depolarization in a caspase-independent manner [108,110,137]. Importantly, the latter studies showed that ROS scavengers reduced HDACi-mediated cell death. In contrast, lower concentrations of the same HDACi induced moderate cell death, but did not lead to an increase in ROS [138]. Therefore, ROS generation contributes to but is not the sole cause of cell-death by HDACi. Both vorinostat and romidepsin were found to induce expression of pro-apoptotic genes [139], and trichostatin A (TSA) treatment was shown to downregulate anti-apoptotic proteins [140]. However, it is unclear whether this effect on gene expression is direct.
Several studies have implicated p53 as a mediator of HDACi-induced cell death [21,141]. p53 transcriptional activity was shown to be negatively regulated by acetylation, and HDAC1 has been reported to deacetylate p53. Additionally, HDAC6 was shown to deacetylate p53 in hepatocellular carcinoma [142]. Moreover, downregulation of p53 has been reported by several studies [143,144]. As a result, HDAC inhibition is suggested to mediate cell death by induction of pro-apoptotic genes through p53 activation. Later studies, however, revealed that HDAC inhibitors have p53-independent cytotoxic activity. Specifically, the HDAC inhibitors vorinostat and entinostat displayed equal and partial cytotoxic activity, respectively, in p53-deficient colon cancer cells [145]. This suggests that inhibition of p53 deacetylation is not required for induction of HDACi-mediated apoptosis.
Studies have reported induction of additional types of cell death by HDACi. Vorinostat and butyrate induce apoptotic and autophagic cell death [146]. In contrast, another report noted a reduction in autophagy upon HDAC inhibition [147]. Notably, vorinostat has also been shown to induce necroptosis, a caspase-independent form of cell death [148]. Additionally, it was also suggested that HDACi cell death might be mediated by hyperacetylation of the NHEJ repair component Ku70 [149,150,151]. Ku70 acetylation and its interactions with Bax have been shown to be part of the apoptotic pathway, in a DNA repair-independent manner [152]. Although this hypothesis is appealing, it is not clear whether Ku70 acetylation is a direct effect of HDACi, nor whether it is necessary and sufficient to induce cell death. In conclusion, HDAC inhibitors have been found to induce cell death through several pathways, which could be due to direct roles of HDACs in cell death pathways or results of other effects of HDACi, such as DNA damage or gene expression dysregulation.
Therapeutic Implications: HDAC inhibitors consistently induce cancer cell death in vitro, which prompted their evaluation as single-agent therapies in pre-clinical and clinical settings. Despite initial success in T-cell lymphoma and multiple myeloma, most clinical trials for several solid cancers, including ovarian, NSCLC, colorectal and prostate, indicated that HDAC inhibitor monotherapy is not an effective treatment [153,154,155,156]. This could be due to poor pharmacokinetics and high toxicity profiles [157], or due to distinct genetic or epigenetic alterations present in hematological cancers [79,158]. Thus, future studies in solid tumors should focus on using HDACi in stratified settings, combinatorial approaches, or enhanced drug delivery methods.
In line with this, several groups have investigated how the modulation of p53 can enhance chemotherapy. Vorinostat has been reported to downregulate thymidilate synthase and p53 expression, thereby enhancing 5-fluorouracil chemotherapy [159]. VPA was shown to synergize with fluoropyrimidine-based chemo-radiotherapy in p53 wild-type and mutant colorectal cancer (CRC) cells [160]. These and several other preliminary studies have led to assessments of HDACi-chemotherapy combinations in the clinic (Table 3).
Resistance to therapy can emerge due to heterogeneity in the cell death threshold and kinetics [161,162]. For example, an in-silico study suggested that cell death induced by the tumor necrosis factor-related apoptosis inducing ligand (TRAIL) directly correlates with a threshold of caspase-8 activity, and that expression of anti-apoptotic proteins modulates this threshold [163]. Therefore, HDACi can be employed to lower the apoptotic threshold of subpopulations with high expression levels of anti-apoptotic genes.
Metabolism. HDAC inhibitors have been shown to have strong effects on cell metabolism. Several studies have reported reductions in glycolysis upon HDAC inhibition, and in one study the pentose phosphate pathway was shown to be reduced as well [86,164,165,166]. Importantly, this effect was observed by several HDAC inhibitors in several types of cancer. It is possible that it is driven by inhibition of HDAC3, which represents a major regulator of glucose metabolism and fatty acid oxidation in muscle and adipose tissue [167,168]. In another study, the authors identified a synergistic interaction between HDAC inhibition and inhibition of glycolysis [169,170], although the reason for this is not known. The link between HDAC inhibition and perturbation of metabolism could be related to the regulation of enhancers. In glioblastoma, inhibition of glycolysis by HDACi was attributed to MYC super-enhancer disruption and glycolysis gene downregulation [86]. Notably, in a proteomics approach using a vorinostat affinity probe, vorinostat was found to directly interact with enolase 1 (ENO1) and could therefore potentially inhibit its enzymatic role in glycolysis [171].
Therapeutic Implications: Metabolic reprogramming is known to introduce metabolic liabilities in cancer cells [172]. Therefore, dysregulation of metabolism by HDAC inhibition presents an opportunity for design of combinatorial or targeted therapeutic strategies to achieve enhanced responses. Notably, several studies have identified synergistic interactions between HDACi and inhibitors of metabolic pathways such as glycolysis, fatty acid β-oxidation, and oxidative phosphorylation in glioma [86,169]. It would be interesting to see whether tumors with inherent defects in these pathways due to specific mutations or limited nutrient availability would be more sensitive to HDAC inhibition.
Cell metabolism responds dynamically to intracellular and extracellular cues present in tumor cells, such as nutrient availability, signaling pathway activation, and gene expression. These stimuli constitute a considerable source of tumor heterogeneity that affects responses [173]. Therefore, HDACi could be applied to counter this heterogeneity, by forcing cells to conform to a certain metabolic phenotype. For example, glucose levels generate heterogeneity in cell glycolysis [174], which could be countered by using HDACi to inhibit glycolysis in all cells. This could be a way to introduce a bottleneck that limits cancer cell adaptability and resistance. In addition, induction of glycolysis after chemotherapy has been found to support cell survival and resistance in ovarian cancer [74]. As a result, HDACi-mediated inhibition of glycolysis could be leveraged to inhibit this resistance mechanism.
Tumor Microenvironment. The tumor microenvironment (TME) is central to cancer development, progression, and treatment [175]. HDAC inhibitors have been found to regulate several aspects of the TME that are important for clinical applications, including the stroma, innate immunity, and angiogenesis.
The advancements in immunotherapy in recent years have highlighted the importance of the tumor’s immune system components in cancer therapy [176]. HDAC inhibitors have been found to induce pleiotropic effects in the immune response, but in general they seem to promote its activation in cancer settings [177,178]. Specifically, HDACi have been shown to promote antigen presentation through induction of major histocompatibility complex (MHC) gene expression [179,180]. In addition, they have been found to modulate the expression levels of immune response-related genes in cancer cells and macrophages [181]. HDACi have been shown to promote NFκB signaling by inhibiting the ability of HDAC1 and HDAC2 to repress p65 target gene expression, and by inhibiting p65 deacetylation by HDAC3 [182,183]. Moreover, HDACs and HDACi have been shown to modulate immune cell populations and functions, such as in innate immune control of T helper 1 cells and CD4+ T cell activation [184,185]. Interestingly, a recent study reported that low-dose TSA potentiates the anti-tumor activities of tumor-associated macrophages and immune cell infiltration [186]. Overall, these effects are suggested to stem from the inhibition of HDAC activities, such as modulating gene transcription and transcription factor deacetylation.
Cancer-associated fibroblasts (CAFs), constitute another component of the tumoral stroma that have important roles in cancer treatment [187]. They induce several tumor-promoting effects, including metastasis, immune suppression, and chemoresistance [188]. HDACi have been reported to both suppress [189] and induce [190] CAF activity, which suggests context-dependent function. Further research is needed to draw conclusions.
Angiogenesis is a major aspect of tumor progression and treatment. Generation of blood vessels supports tumor growth by delivering nutrients and oxygen, and regulates other aspects of progression, such as metabolism and metastasis [191]. Several anti-angiogenesis agents are currently used for cancer treatment, and act by targeting vascular endothelial growth factor A (VEGFA) and its receptors [192,193]. HDACi have displayed anti-angiogenic effects in vitro and in vivo [194,195,196,197]. This function is attributed to several mechanisms, including transcriptional regulation of angiogenesis effectors such as VHL, HIF1α, and VEGFA; and direct interactions of HDAC1, HDAC3, HDAC4, and HDAC6 with HIF1α [198,199].
Therapeutic Implications: Tumor and host immune responses are central to cancer therapy and often dictate treatment responses. In the context of immunotherapy, the degree of tumor infiltration by immune cells is recognized as an important predictive factor and can be crucial for its success [200]. In addition, the composition of the immune microenvironment within the tumor can also affect the efficiency of immunotherapy [75]. Therefore, modulation of innate and adaptive immunity by HDACi presents a therapeutic opportunity. Increased tumor cell immunogenicity can boost the effectiveness of immune cells and reduce microenvironment heterogeneity by increasing immune cell recruitment in less infiltrated parts of the tumor. Interestingly, increased immune cell infiltration has been shown to be induced by HDACi [186,201], and several studies have shown that HDACi being combined with immunotherapy improves response in some solid tumors [181,202]. Clinical trials are currently investigating this approach in non-small cell lung cancer and lymphoma patients (NCT01928576; NCT03161223).
Regarding angiogenesis, the suppressive activity of HDACi is a desirable effect in the treatment of solid tumors and has been suggested as a way to suppress tumor progression and enhance therapy [197,198]. Interestingly HDACi have been shown to reduce vascularization in CTCL [203]. In addition, HDACi are under clinical investigation in combination with anti-angiogenesis inhibitors in clear-cell renal cell carcinoma (ccRCC) [204]. ccRCC is driven by VEGF signaling activation, and anti-angiogenesis inhibitors are currently used for its treatment [205,206]. Frequently, tumors develop resistance through activation of angiogenic signaling through other means, such as HIFα activation [207,208]. Therefore, the pleiotropic inhibitory effect of HDACi on angiogenesis pathways is employed to counteract these resistance mechanisms.
Phenotypic Plasticity. The plasticity of tumor cells in phenotypic states, such as epithelial–mesenchymal transition (EMT) and cancer stemness, contributes significantly to tumor heterogeneity [209]. EMT refers to the reversible shift of cells from an epithelial state, characterized by strong cell-to-cell adhesion, to a mesenchymal state, where cells become more migratory and invasive [210,211]. This transition affects several cellular processes and components, including the cytoskeleton, metabolism, innate immunity, proliferation, and apoptosis. In cancer this process is usually partial and is associated with metastatic disease and chemoresistance [212]. Several groups have demonstrated that HDAC inhibitors suppress the EMT transcriptional program in several cancer types, including breast, biliary tract, bladder, and others [213,214,215,216,217]. Importantly, this suppression was evident in cell lines that were predominantly mesenchymal-like, either intrinsically or due to exogenous signals such as TGFβ. This may explain why the opposite effect has also been reported in epithelial-like cancer cell lines [218,219]. Therefore, it is likely that the influence of HDAC inhibitors in EMT is context dependent, specifically on the initial phenotypic state of the cancer cells.
It is well accepted that epigenetic heterogeneity leads to transcriptional plasticity and adaptive responses to chemoresistance in cancer. Cancer stem cells (CSCs) and poorly differentiated cancer cells represent sources of cellular heterogeneity within tumors, and there is strong clinical evidence that these subpopulations are critical to conferring drug resistance [220,221,222,223]. CSCs have the potential to self-renew with symmetric or asymmetric division and are characterized by high tumor-initiating capacity [224]. Moreover, their divisions can generate differentiated progeny and transient amplifying cells, increasing the tumor’s heterogeneity. In addition, CSCs can enter a quiescent state that protects them upon treatment, since chemotherapy is more effective against proliferating cells [225]. Targeted pharmaceutical inhibition of stem cell-related signaling pathways, such as Wnt, Notch, and Hedgehog, causes high levels of toxicity, as normal tissue homeostasis also relies on these pathways [226]. Moreover, compensatory activation of other signaling pathways often confers resistance [227].
Just as in normal cells, self-renewal of cancer stem cells or proliferation of cells with undifferentiated phenotypes is highly dependent on key transcriptional programs that are regulated by specific epigenetic patterns in their chromatin [228,229]. For instance, differential DNA methylation is associated with the expression of stem cell marker genes such as CD44, CD133, and Musashi-1 (MSI1). More specifically, hypomethylation can activate these CSC genes in aggressive tumors [230,231]. Other studies in glioblastomas (GBM) have demonstrated that chromatin in CSCs is characterized by reduced levels of the silencing histone mark, H3K27me3, and possesses a more open conformation compared to non-CSCs, which together allow genes that maintain the stem cell phenotype to be expressed [232]. The overexpression of several HDACs has been associated with cancer stem cell identity, regulation of the Sonic-Hedgehog pathway, and poor survival in GBM, NSCLC, and breast and ovarian cancers [233,234,235,236]. In addition, in acute myeloid leukemia(AML), CSCs were characterized by higher H3K4me3 levels on genes involved in stem cell identity, proliferation, and metabolic reprogramming compared to non-CSCs, indicating that differentiation processes were associated with epigenetic silencing of stem cell identity genes [237].
Therapeutic Implications. Phenotypic plasticity such as that in the form of EMT state contributes significantly to tumor heterogeneity [209], and is considered an important mechanism of therapy resistance because it is accompanied by anti-apoptotic signaling and drug efflux [212,238]. The EMT-suppressive effect of HDACi in mesenchymal-like cells can be employed in tumors where EMT occurs and mediates resistance, such as patient subsets in breast and pancreatic cancer [239,240]. In this setting, EMT inhibition could confer several beneficial effects, including enhancements of the effects of other therapeutics, and suppression of mesenchymal subclone emergence and metastasis.
The importance of epigenetic regulation in CSCs suggests a dependence that could be exploited for cancer treatment. Specifically, disruption of chromatin states and the expression of genes required to maintain cancer stemness could be a way to target CSC populations and reduce heterogeneity [228]. Interestingly, recent studies have demonstrated that targeting the epigenetic state of the CSC pool in tumors via HDAC inhibitors can suppress the growth of cancer stem cells without impairing the functions of normal stem cells [241]. For instance, in triple-negative breast cancer, the Class I HDAC inhibitor entinostat was reported to decrease the CSC population [242]. Similarly, HDAC inhibition has been shown to reduce the cancer stem cell burden in GBM tumors [234,243,244] and NSCLC [236].

3. Dissecting the Variables of HDAC Inhibition

Thus far we have explored the application of HDACi in cancer treatment to counter tumor heterogeneity and achieve greater responses by targeting specific cellular processes (Table 2). However, targeted therapeutic approaches require deep understanding of the effects and mechanisms of action of the candidate drugs. At first glance, the widespread effects of HDAC inhibitors in cell biology do not suggest a unifying mechanism of action. Several questions arise when considering how HDAC inhibitors act: Which HDACs are relevant for the effects of HDACi? Why are HDACi only effective in hematological malignancies? Which effects of HDACi are primary and which are secondary? How does HDACi dosage affect the phenotypes observed? Answering these questions is vital for comprehending why HDAC inhibition is successful or is not in cancer treatment, and for the development of effective therapeutic strategies, such as countering tumor heterogeneity.
Target selectivity. Firstly, we must consider whether the selectivity of HDAC inhibitors influences their biological activity. Despite their name, histone deacetylases have diverse targets, which are often not histones. Moreover, HDAC inhibitors often target more than one HDAC protein, and the target specificity varies between compounds. The drugs used in clinical and pre-clinical settings of cancer treatment vary in their target specificity (Table 1). Some are selective for Class I HDACs, such as romidepsin and entinostat, and others inhibit Classes I, II, and IV to various degree, such as vorinostat and its derivatives [157]. Despite their differences in targeting selectivity, HDAC inhibitors are used in similar clinical settings and seem to have similar phenotypic effects. In addition, they invariably inhibit HDAC family members 1, 2, and 3, which suggests that inhibition of these proteins is what mediates their anti-cancer activity.
This notion is challenged by the fact that several Class II and HDAC6-specific inhibitors have shown promise as anticancer agents as well. HDAC6 deacetylates cytoplasmic proteins such as tubulin and HSP90, and has important functions in tumorigenesis, such as modulation of protein homeostasis through regulation of HSP90 and proteasomal degradation [245,246,247], p53 apoptotic activity [248], and tyrosine kinase signaling [249]. As a result, several groups have contributed to the development of HDAC6-selective inhibitors [250,251,252]. Ricolinostat, an HDAC6-selective inhibitor, was effective in models of multiple myeloma and lymphoma in vitro and in vivo [253,254]. Several clinical trials are currently assessing its efficacy in multiple myeloma and lymphoma [255] (NCT01997840; NCT02091063). Notably, ricolinostat was recently reported to mediate cell death through off-target toxicity [256], which could be attributed to its low selectivity for HDAC6 compared to Class I HDACs [253]. Another HDAC6-specific inhibitor, citarinostat, has also displayed anti-cancer activity and is being investigated in the clinic [257] (NCT02886065).
To our knowledge, few studies have directly compared how HDACi with different selectivities affect various cellular processes. In a study of p53 and NFκB signaling, the Class I/IIa selective inhibitor VPA had different effects from the HDAC6-selective inhibitor marbostat [258]. Vorinostat and romidepsin similarly induced the expression of pro-apoptotic genes [139], and cytokine expression in CTCL [259]. Sonnemann et al. reported a partial difference between entinostat and vorinostat in p53 dependency [145]. Overall, Class I and Class I/II/IV HDACi have mostly similar effects in cells, but it is probably best that HDAC6-specific HDACi are considered distinct agents. Finally, it should be noted that several effects of HDACi have been reproduced by genetic knock-down or knock-out of Class I HDACs [84,86,104,143]. Therefore, it is likely that many or most effects of HDAC inhibitors stem from inhibition of Class I HDACs.
Tissue specificity. Secondly, we must consider whether the activity of HDAC inhibitors is tissue-specific. So far, HDACi are being employed in the clinic only for treatment of a few hematological cancers, which could point to a tissue-specific effect. Class I and Class II HDACs have important roles in T-cell development and differentiation, which could underlie the effectiveness of HDAC inhibitors in T-cell lymphoma [158]. In addition, HDACi have been shown to target blood cancer-specific pathways, such as BCL6 overexpression in B-cell lymphoma [260], aggresome dependency in multiple myeloma [261,262], and HDAC6 overexpression in lymphoma [250].
Therefore, it is possible that HDACi have not achieved the desired efficacy as monotherapies in solid tumors because of tissue specificity, and due to inter- and intra-tumoral heterogeneity. Nevertheless, HDAC inhibitors exhibit useful biological effects in both solid and hematological cancers, that can be employed to target specific cancer alterations and improve the efficacy of other therapies. Additionally, HDACs are found overexpressed in solid cancers as well [263], which also supports the use of HDACi in solid tumor treatment.
Dose-Dependent Effects. Thirdly, we need to consider how HDACi dosage influences the manifestation of their phenotypic effects. Several studies have indicated differences in concentration thresholds among effects such as DNA damage and histone hyperacetylation. ROS generation and apoptosis are generally observed at high-concentrations of HDACi [110,138]. In a study on the dose-dependent effects of the Class I HDACi largazole, cell cycle blocking was observed only with higher concentrations of the inhibitor [264]. In contrast, changes in H3K9ac and H3K27ac abundance and genomic distribution were observed even at the lowest concentrations. Moreover, there was marked difference between high and low-dose HDACi in the subsets of enhancers and transcripts that were affected. In line with this, in a study testing vorinostat dosage,, histone acetylation induction was observed at lower doses than DNA damage, assessed by γH2A.X [102].
In summary, the current evidence suggests that lower levels of HDACi are sufficient to disrupt epigenetic regulation, for example, through enhancer acetylation, but more severe phenotypic effects such as genomic instability, cell cycle blocking, and apoptosis occur only by extensive HDAC inhibition. In the clinic, HDACi are administered at sufficiently high concentrations that tumor cells should display severe phenotypic effects [157]. However, it is likely that poorly vascularized regions of solid tumors are exposed to lower concentrations of HDACi, which likely impairs their cell killing effects. This might explain their success in the treatment of hematological cancers, where drug diffusion is unobstructed. Moreover, this is an additional incentive to look for combinatorial or targeted therapeutic approaches where a low dose of HDACi is sufficient. The disruption of epigenomic and transcriptional regulation observed even at low concentrations of HDACi most likely confers vulnerabilities that could be exploited for therapeutic benefit.
Primary and Secondary effects. Finally, we need to identify which of the effects of HDAC inhibition are primary and which are secondary. Ideally, this could be achieved by construction of a time course for all effects observed, but only a few studies have assessed phenotypic effects earlier than 16 h after HDACi treatment. The earliest events are most likely histone acetylation and DNA damage, which were observed to increase in just 3 minutes of exposure to TSA [107]. The kinetics of DNA damage generation suggest that it is likely caused by genome instability and not by indirect means, such as downregulation of DNA repair components. Generation of ROS was shown to happen as early as 2 hours post-treatment [110], but earlier timepoints have not been assessed, and it is thus not clear whether this is the underlying cause of DNA damage. TSA primed cells for enhanced induction of NFκB signaling after 1 h of treatment [183]. Effects on transcription were found as early as 1 hour post-treatment, when vorinostat was shown to significantly reduce MYC mRNA levels and cause widespread gene expression changes [77]. Two studies reported effects of HDACi on gene expression as early as 4 h and 10 min after treatment respectively, which stemmed from modulation of transcriptional elongation [88,89]. Apoptosis is generally observed after 17 h of treatment, so it is likely a secondary event.
Based on these findings and our knowledge so far, we can speculate on the order of events after HDAC inhibition (Figure 5). With a low dose of a HDACi, the histone acetylation increase modulates the enhancer and promoter’s chromatin structure and activity, which leads to transcriptional changes. This could explain the transcription priming observed in immune response genes. Subsequently, gene downregulation impairs metabolic and DNA repair pathways. With high doses of HDACi, additional effects are observed. Initially, strong HDAC inhibition disrupts DNA replication and/or DNA repair by histone and/or protein hyperacetylation. This leads to DNA damage, H2A.X phosphorylation, and activation of the DNA-damage response and cell cycle checkpoints. Subsequently, excessive DNA damage and an inability to resolve it, due to functional and transcriptional suppression of DNA repair, leads to apoptosis and possibly to activation of innate immunity pathways, such as antigen presentation.

4. Conclusions

Heterogeneity presents a major obstacle to cancer therapy. Epigenetic drugs such as HDAC inhibitors could present a way to reduce tumor heterogeneity and suppress inherent and acquired resistance mechanisms to therapeutic approaches. In this review, we examined how HDAC inhibitors affect cell processes such as chromatin regulation, gene expression, and genome stability. Significant advances have been achieved in our understanding of the underlying mechanisms, but several questions remain unanswered. In addition, we highlighted several aspects of HDACi that could be utilized for the development of new therapeutic approaches. Since their chemotherapeutic efficacy is restricted to a handful of hematopoietic cancers, we suggest that other HDACi effects, such as modulation of gene expression, DNA repair, and innate immunity, could be utilized instead.
Importantly, predictive biomarkers of gene expression, DNA repair capabilities, and the immune system’s interactions with the tumor are subject to assessment prior to treatment selection. In our perspective, significant levels of tumor heterogeneity in such biomarkers could be neutralized through precise use of HDAC inhibitors. Finally, we further considered the kinetics of HDAC inhibition, in order to deconvolute their modes of action and inform treatment regimen design. Overall, deep mechanistic understanding, a precise approach, and regimen optimization will be the key to making the most out of HDAC inhibitors as a tool for cancer treatment.
Table 3. Clinical trials evaluating HDAC inhibitors in cancer treatment. More clinical trials exploring drug combinations of HDACi can be found here [5,6,7]. Abbreviations: CHOP: cyclophosphamide, doxorubicin, vincristine, and prednisone; HR+: hormone receptor positive; HER2: human epidermal growth factor 2 negative; CRPC: castration resistant prostate cancer; SCLC: small cell lung cancer; AML: acute myelodysplastic leukemia; MDS: myelodysplastic syndrome; DLBCL: diffuse large B-cell lymphoma; HGG: high grade glioma.
Table 3. Clinical trials evaluating HDAC inhibitors in cancer treatment. More clinical trials exploring drug combinations of HDACi can be found here [5,6,7]. Abbreviations: CHOP: cyclophosphamide, doxorubicin, vincristine, and prednisone; HR+: hormone receptor positive; HER2: human epidermal growth factor 2 negative; CRPC: castration resistant prostate cancer; SCLC: small cell lung cancer; AML: acute myelodysplastic leukemia; MDS: myelodysplastic syndrome; DLBCL: diffuse large B-cell lymphoma; HGG: high grade glioma.
TreatmentHDACiDrug CombinationCancer TypePhaseID/Ref.
MonotherapyRomidepsin CRPCII[153]
SCLCII[265]
Belinostat Ovarian cancerII[154]
Vorinostat Solid tumorsII[155]
Ovarian cancerII[156]
Panobinostat AMLIINCT00880269
Ricolinostat Lymphoma, Lymphoid malignanciesIb/IINCT02091063
MaintenancePanobinostat AML, MDSIIINCT04326764
ChemotherapyRomidepsinCHOPPTCLIIINCT01796002
VorinostatTemozolomideGliomaINCT00268385
Valproic acidEpirubicin, 5-Fluorouracil, CyclophosphamideSolid tumorsINCT00246103
PARPiEntinostatOlaparibOvarian, Peritoneal, Fallopian tube cancerIINCT03924245
VorinostatOlaparibBreast CancerINCT03742245
Endocrine therapyEntinostatExemestaneHR+ & HER2- Breast cancerIIINCT02115282 [9,266]
TucidinostatHR+ & HER2- Breast cancerIIINCT02482753 [8]
EGFR TKIEntinostatErlotinibNSCLCIINCT00602030 [94]
VorinostatGivinostatNSCLCINCT02151721
Anti-angiogenicVorinostatBevacizumabccRCCI/II[204]
MultiplePanobinostatBortezomid & DexamethasoneMultiple MyelomaIIINCT01023308 [267]
EntinostatNivolumab, AzacytidineNSCLCIINCT01928576
RomidepsinDurvalumab, Azacytidine/PralatrexateLymphomaI/IIaNCT03161223
RicolinostatPomalidomide, DexamethasoneMultiple MyelomaIb/IINCT01997840
CitarinostatPVX-410(cancer vaccine), LenalidomideMultiple MyelomaINCT02886065
Tucidinostatrituximab-CHOPMYC/BCL2 Double-Expressor DLBCLIIINCT04231448 [268]
VorinostatBevacizumab, Temozolomide, RadiotherapyHGGII/IIINCT01236560

Author Contributions

D.K. and T.R.: writing—review and editing. Both authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

We apologize to all authors whose work was not cited due to space limitations or an oversight on our part.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mann, B.S.; Johnson, J.R.; Cohen, M.H.; Justice, R.; Pazdur, R. FDA Approval Summary: Vorinostat for Treatment of Advanced Primary Cutaneous T-Cell Lymphoma. Oncology 2007, 12, 1247–1252. [Google Scholar] [CrossRef] [PubMed]
  2. Lee, H.-Z.; Kwitkowski, V.E.; Valle, P.L.D.; Ricci, M.S.; Saber, H.; Habtemariam, B.A.; Bullock, J.; Bloomquist, E.; Shen, Y.L.; Chen, X.-H.; et al. FDA Approval: Belinostat for the Treatment of Patients with Relapsed or Refractory Peripheral T-Cell Lymphoma. Clin. Cancer Res. 2015, 21, 2666–2670. [Google Scholar] [CrossRef] [Green Version]
  3. Iyer, S.P.; Foss, F.F. Romidepsin for the Treatment of Peripheral T-Cell Lymphoma. Oncology 2015, 20, 1084–1091. [Google Scholar] [CrossRef] [Green Version]
  4. Prince, H.M.; Dickinson, M. Romidepsin for Cutaneous T-Cell Lymphoma. Clin. Cancer Res. 2012, 18, 3509–3515. [Google Scholar] [CrossRef] [Green Version]
  5. Jenke, R.; Reßing, N.; Hansen, F.K.; Aigner, A.; Büch, T. Anticancer Therapy with HDAC Inhibitors: Mechanism-Based Combination Strategies and Future Perspectives. Cancers 2021, 13, 634. [Google Scholar] [CrossRef] [PubMed]
  6. Suraweera, A.; O’Byrne, K.J.; Richard, D.J. Combination Therapy With Histone Deacetylase Inhibitors (HDACi) for the Treatment of Cancer: Achieving the Full Therapeutic Potential of HDACi. Front. Oncol. 2018, 8, 92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Wang, P.; Wang, Z.; Liu, J. Role of HDACs in Normal and Malignant Hematopoiesis. Mol. Cancer 2020, 19, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Jiang, Z.; Li, W.; Hu, X.; Zhang, Q.; Sun, T.; Cui, S.; Wang, S.; Ouyang, Q.; Yin, Y.; Geng, C.; et al. Tucidinostat plus Exemestane for Postmenopausal Patients with Advanced, Hormone Receptor-Positive Breast Cancer (ACE): A Randomised, Double-Blind, Placebo-Controlled, Phase 3 Trial. Lancet Oncol. 2019, 20, 806–815. [Google Scholar] [CrossRef]
  9. Yeruva, S.L.H.; Zhao, F.; Miller, K.D.; Tevaarwerk, A.J.; Wagner, L.I.; Gray, R.J.; Sparano, J.A.; Connolly, R.M. E2112: Randomized Phase Iii Trial of Endocrine Therapy plus Entinostat/Placebo in Patients with Hormone Receptor-Positive Advanced Breast Cancer. NPJ Breast Cancer 2018, 4, 1–5. [Google Scholar] [CrossRef] [Green Version]
  10. Furumai, R.; Matsuyama, A.; Kobashi, N.; Lee, K.-H.; Nishiyama, M.; Nakajima, H.; Tanaka, A.; Komatsu, Y.; Nishino, N.; Yoshida, M.; et al. FK228 (Depsipeptide) as a Natural Prodrug That Inhibits Class I Histone Deacetylases. Cancer Res. 2002, 62, 4916–4921. [Google Scholar]
  11. Bowers, A.; West, N.; Taunton, J.; Schreiber, S.L.; Bradner, J.E.; Williams, R.M. Total Synthesis and Biological Mode of Action of Largazole: A Potent Class I Histone Deacetylase Inhibitor. J. Am. Chem. Soc. 2008, 130, 11219–11222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Khan, N.; Jeffers, M.; Kumar, S.; Hackett, C.; Boldog, F.; Khramtsov, N.; Qian, X.; Mills, E.; Berghs, S.C.; Carey, N.; et al. Determination of the Class and Isoform Selectivity of Small-Molecule Histone Deacetylase Inhibitors. Biochem. J. 2007, 409, 581–589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Arts, J.; King, P.; Mariën, A.; Floren, W.; Beliën, A.; Janssen, L.; Pilatte, I.; Roux, B.; Decrane, L.; Gilissen, R.; et al. JNJ-26481585, a Novel “Second-Generation” Oral Histone Deacetylase Inhibitor, Shows Broad-Spectrum Preclinical Antitumoral Activity. Clin. Cancer Res. 2009, 15, 6841–6851. [Google Scholar] [CrossRef] [Green Version]
  14. FDA Approves Farydak (Panobinostat) for Multiple Myeloma. Oncol. Times 2015, 37, 16–17. [CrossRef]
  15. Beckers, T.; Burkhardt, C.; Wieland, H.; Gimmnich, P.; Ciossek, T.; Maier, T.; Sanders, K. Distinct Pharmacological Properties of Second Generation HDAC Inhibitors with the Benzamide or Hydroxamate Head Group. Int. J. Cancer 2007, 121, 1138–1148. [Google Scholar] [CrossRef] [PubMed]
  16. Ning, Z.-Q.; Li, Z.-B.; Newman, M.J.; Shan, S.; Wang, X.-H.; Pan, D.-S.; Zhang, J.; Dong, M.; Du, X.; Lu, X.-P. Chidamide (CS055/HBI-8000): A New Histone Deacetylase Inhibitor of the Benzamide Class with Antitumor Activity and the Ability to Enhance Immune Cell-Mediated Tumor Cell Cytotoxicity. Cancer Chemother. Pharmacol. 2012, 69, 901–909. [Google Scholar] [CrossRef] [PubMed]
  17. Göttlicher, M.; Minucci, S.; Zhu, P.; Krämer, O.H.; Schimpf, A.; Giavara, S.; Sleeman, J.P.; Lo Coco, F.; Nervi, C.; Pelicci, P.G.; et al. Valproic Acid Defines a Novel Class of HDAC Inhibitors Inducing Differentiation of Transformed Cells. EMBO J. 2001, 20, 6969–6978. [Google Scholar] [CrossRef] [Green Version]
  18. Gurvich, N.; Tsygankova, O.M.; Meinkoth, J.L.; Klein, P.S. Histone Deacetylase Is a Target of Valproic Acid-Mediated Cellular Differentiation. Cancer Res. 2004, 64, 1079–1086. [Google Scholar] [CrossRef] [Green Version]
  19. Milazzo, G.; Mercatelli, D.; Di Muzio, G.; Triboli, L.; De Rosa, P.; Perini, G.; Giorgi, F.M. Histone Deacetylases (HDACs): Evolution, Specificity, Role in Transcriptional Complexes, and Pharmacological Actionability. Genes 2020, 11, 556. [Google Scholar] [CrossRef]
  20. Seto, E.; Yoshida, M. Erasers of Histone Acetylation: The Histone Deacetylase Enzymes. Cold Spring Harb. Perspect. Biol. 2014, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Luo, J.; Su, F.; Chen, D.; Shiloh, A.; Gu, W. Deacetylation of P53 Modulates Its Effect on Cell Growth and Apoptosis. Nature 2000, 408, 377–381. [Google Scholar] [CrossRef]
  22. Nebbioso, A.; Carafa, V.; Conte, M.; Tambaro, F.P.; Abbondanza, C.; Martens, J.; Nees, M.; Benedetti, R.; Pallavicini, I.; Minucci, S.; et al. C-Myc Modulation and Acetylation Is a Key HDAC Inhibitor Target in Cancer. Clin. Cancer Res. 2017, 23, 2542–2555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Vannini, A.; Volpari, C.; Filocamo, G.; Casavola, E.C.; Brunetti, M.; Renzoni, D.; Chakravarty, P.; Paolini, C.; Francesco, R.D.; Gallinari, P.; et al. Crystal Structure of a Eukaryotic Zinc-Dependent Histone Deacetylase, Human HDAC8, Complexed with a Hydroxamic Acid Inhibitor. Proc. Natl. Acad. Sci. USA 2004, 101, 15064–15069. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Jung, M.; Hoffmann, K.; Brosch, G.; Loidl, P. Analogues of Trichosтatin a and Trapoxin B as Histone Deacetylase Inhibitors. Bioorg. Med. Chem. Lett. 1997, 7, 1655–1658. [Google Scholar] [CrossRef]
  25. Roche, J.; Bertrand, P. Inside HDACs with More Selective HDAC Inhibitors. Eur. J. Med. Chem. 2016, 121, 451–483. [Google Scholar] [CrossRef]
  26. Lombardi, P.M.; Cole, K.E.; Dowling, D.P.; Christianson, D.W. Structure, Mechanism, and Inhibition of Histone Deacetylases and Related Metalloenzymes. Curr. Opin. Struct. Biol. 2011, 21, 735–743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Maolanon, A.R.; Madsen, A.S.; Olsen, C.A. Innovative Strategies for Selective Inhibition of Histone Deacetylases. Cell Chem. Biol. 2016, 23, 759–768. [Google Scholar] [CrossRef] [Green Version]
  28. Yoshida, M.; Furumai, R.; Nishiyama, M.; Komatsu, Y.; Nishino, N.; Horinouchi, S. Histone Deacetylase as a New Target for Cancer Chemotherapy. Cancer Chemother. Pharmacol. 2001, 48 (Suppl. 1), S20–S26. [Google Scholar] [CrossRef]
  29. Bertrand, P. Inside HDAC with HDAC Inhibitors. Eur. J. Med. Chem. 2010, 45, 2095–2116. [Google Scholar] [CrossRef]
  30. Melesina, J.; Simoben, C.V.; Praetorius, L.; Bülbül, E.F.; Robaa, D.; Sippl, W. Strategies To Design Selective Histone Deacetylase Inhibitors. ChemMedChem 2021, 16, 1336–1359. [Google Scholar] [CrossRef]
  31. Turajlic, S.; Sottoriva, A.; Graham, T.; Swanton, C. Resolving Genetic Heterogeneity in Cancer. Nat. Rev. Genet. 2019, 20, 404–416. [Google Scholar] [CrossRef]
  32. Rosenthal, R.; McGranahan, N.; Herrero, J.; Swanton, C. Deciphering Genetic Intratumor Heterogeneity and Its Impact on Cancer Evolution. Annu. Rev. Cancer Biol. 2017, 1, 223–240. [Google Scholar] [CrossRef] [Green Version]
  33. Flavahan, W.A.; Gaskell, E.; Bernstein, B.E. Epigenetic Plasticity and the Hallmarks of Cancer. Science 2017, 357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. González-Silva, L.; Quevedo, L.; Varela, I. Tumor Functional Heterogeneity Unraveled by ScRNA-Seq Technologies. Trends Cancer 2020, 6, 13–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Dagogo-Jack, I.; Shaw, A.T. Tumour Heterogeneity and Resistance to Cancer Therapies. Nat. Rev. Clin. Oncol. 2018, 15, 81–94. [Google Scholar] [CrossRef] [PubMed]
  36. Burrell, R.A.; Swanton, C. Tumour Heterogeneity and the Evolution of Polyclonal Drug Resistance. Mol. Oncol. 2014, 8, 1095–1111. [Google Scholar] [CrossRef]
  37. Galluzzi, L.; Senovilla, L.; Vitale, I.; Michels, J.; Martins, I.; Kepp, O.; Castedo, M.; Kroemer, G. Molecular Mechanisms of Cisplatin Resistance. Oncogene 2012, 31, 1869–1883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Chabner, B.A.; Roberts, T.G. Chemotherapy and the War on Cancer. Nat. Rev. Cancer 2005, 5, 65–72. [Google Scholar] [CrossRef]
  39. Deans, A.J.; West, S.C. DNA Interstrand Crosslink Repair and Cancer. Nat. Rev. Cancer 2011, 11, 467–480. [Google Scholar] [CrossRef] [Green Version]
  40. Reisz, J.A.; Bansal, N.; Qian, J.; Zhao, W.; Furdui, C.M. Effects of Ionizing Radiation on Biological Molecules--Mechanisms of Damage and Emerging Methods of Detection. Antioxid. Redox Signal. 2014, 21, 260–292. [Google Scholar] [CrossRef]
  41. Zahreddine, H.; Borden, K. Mechanisms and Insights into Drug Resistance in Cancer. Front. Pharmacol. 2013, 4. [Google Scholar] [CrossRef] [Green Version]
  42. Gatti, L.; Zunino, F. Overview of Tumor Cell Chemoresistance Mechanisms. In Chemosensitivity: Volume II: In VIVO Models, Imaging, and Molecular Regulators; Blumenthal, R.D., Ed.; Methods in Molecular MedicineTM; Humana Press: Totowa, NJ, USA, 2005; pp. 127–148. ISBN 978-1-59259-889-2. [Google Scholar]
  43. Holohan, C.; Van Schaeybroeck, S.; Longley, D.B.; Johnston, P.G. Cancer Drug Resistance: An Evolving Paradigm. Nat. Rev. Cancer 2013, 13, 714–726. [Google Scholar] [CrossRef] [PubMed]
  44. Jansen, G.; Mauritz, R.; Drori, S.; Sprecher, H.; Kathmann, I.; Bunni, M.; Priest, D.G.; Noordhuis, P.; Schornagel, J.H.; Pinedo, H.M.; et al. A Structurally Altered Human Reduced Folate Carrier with Increased Folic Acid Transport Mediates a Novel Mechanism of Antifolate Resistance. J. Biol. Chem. 1998, 273, 30189–30198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hill, B.T.; Bailey, B.D.; White, J.C.; Goldman, I.D. Characteristics of Transport of 4-Amino Antifolates and Folate Compounds by Two Lines of L5178Y Lymphoblasts, One with Impaired Transport of Methotrexate. Cancer Res. 1979, 39, 2440–2446. [Google Scholar] [PubMed]
  46. Banerjee, D.; Mayer-Kuckuk, P.; Capiaux, G.; Budak-Alpdogan, T.; Gorlick, R.; Bertino, J.R. Novel Aspects of Resistance to Drugs Targeted to Dihydrofolate Reductase and Thymidylate Synthase. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2002, 1587, 164–173. [Google Scholar] [CrossRef] [Green Version]
  47. Longley, D.B.; Harkin, D.P.; Johnston, P.G. 5-Fluorouracil: Mechanisms of Action and Clinical Strategies. Nat. Rev. Cancer 2003, 3, 330–338. [Google Scholar] [CrossRef]
  48. Hirakawa, M.; Sato, Y.; Ohnuma, H.; Takayama, T.; Sagawa, T.; Nobuoka, T.; Harada, K.; Miyamoto, H.; Sato, Y.; Takahashi, Y.; et al. A Phase II Study of Neoadjuvant Combination Chemotherapy with Docetaxel, Cisplatin, and S-1 for Locally Advanced Resectable Gastric Cancer: Nucleotide Excision Repair (NER) as Potential Chemoresistance Marker. Cancer Chemother. Pharmacol. 2013, 71, 789–797. [Google Scholar] [CrossRef]
  49. Chaney, S.G.; Sancar, A. DNA Repair: Enzymatic Mechanisms and Relevance to Drug Response. J. Natl. Cancer Inst. 1996, 88, 1346–1360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Buzdar, A.; Howell, A. Advances in Aromatase Inhibition: Clinical Efficacy and Tolerability in the Treatment of Breast Cancer. Clin. Cancer Res. 2001, 7, 2620–2635. [Google Scholar]
  51. Musgrove, E.A.; Sutherland, R.L. Biological Determinants of Endocrine Resistance in Breast Cancer. Nat. Rev. Cancer 2009, 9, 631–643. [Google Scholar] [CrossRef]
  52. Lei, J.T.; Anurag, M.; Haricharan, S.; Gou, X.; Ellis, M.J. Endocrine Therapy Resistance: New Insights. Breast 2019, 48, S26–S30. [Google Scholar] [CrossRef] [Green Version]
  53. Watson, P.A.; Arora, V.K.; Sawyers, C.L. Emerging Mechanisms of Resistance to Androgen Receptor Inhibitors in Prostate Cancer. Nat. Rev. Cancer 2015, 15, 701–711. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Handle, F.; Prekovic, S.; Helsen, C.; Van den Broeck, T.; Smeets, E.; Moris, L.; Eerlings, R.; Kharraz, S.E.; Urbanucci, A.; Mills, I.G.; et al. Drivers of AR Indifferent Anti-Androgen Resistance in Prostate Cancer Cells. Sci. Rep. 2019, 9, 13786. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Tomasello, C.; Baldessari, C.; Napolitano, M.; Orsi, G.; Grizzi, G.; Bertolini, F.; Barbieri, F.; Cascinu, S. Resistance to EGFR Inhibitors in Non-Small Cell Lung Cancer: Clinical Management and Future Perspectives. Crit. Rev. Oncol. Hematol. 2018, 123, 149–161. [Google Scholar] [CrossRef] [PubMed]
  56. Pao, W.; Chmielecki, J. Rational, Biologically Based Treatment of EGFR -Mutant Non-Small-Cell Lung Cancer. Nat. Rev. Cancer 2010, 10, 760–774. [Google Scholar] [CrossRef] [Green Version]
  57. Lord, C.J.; Ashworth, A. PARP Inhibitors: Synthetic Lethality in the Clinic. Science 2017, 355, 1152–1158. [Google Scholar] [CrossRef] [PubMed]
  58. Gogola, E.; Rottenberg, S.; Jonkers, J. Resistance to PARP Inhibitors: Lessons from Preclinical Models of BRCA-Associated Cancer. Annu. Rev. Cancer Biol. 2019, 3, 235–254. [Google Scholar] [CrossRef]
  59. Ben-Porath, I.; Thomson, M.W.; Carey, V.J.; Ge, R.; Bell, G.W.; Regev, A.; Weinberg, R.A. An Embryonic Stem Cell-like Gene Expression Signature in Poorly Differentiated Aggressive Human Tumors. Nat. Genet. 2008, 40, 499–507. [Google Scholar] [CrossRef]
  60. Wong, D.J.; Liu, H.; Ridky, T.W.; Cassarino, D.; Segal, E.; Chang, H.Y. Module Map of Stem Cell Genes Guides Creation of Epithelial Cancer Stem Cells. Cell Stem Cell 2008, 2, 333–344. [Google Scholar] [CrossRef] [Green Version]
  61. Doi, A.; Park, I.-H.; Wen, B.; Murakami, P.; Aryee, M.J.; Irizarry, R.; Herb, B.; Ladd-Acosta, C.; Rho, J.; Loewer, S.; et al. Differential Methylation of Tissue- and Cancer-Specific CpG Island Shores Distinguishes Human Induced Pluripotent Stem Cells, Embryonic Stem Cells and Fibroblasts. Nat. Genet. 2009, 41, 1350–1353. [Google Scholar] [CrossRef] [Green Version]
  62. Sottoriva, A.; Verhoeff, J.J.C.; Borovski, T.; McWeeney, S.K.; Naumov, L.; Medema, J.P.; Sloot, P.M.A.; Vermeulen, L. Cancer Stem Cell Tumor Model Reveals Invasive Morphology and Increased Phenotypical Heterogeneity. Cancer Res. 2010, 70, 46–56. [Google Scholar] [CrossRef] [Green Version]
  63. Patel, A.P.; Tirosh, I.; Trombetta, J.J.; Shalek, A.K.; Gillespie, S.M.; Wakimoto, H.; Cahill, D.P.; Nahed, B.V.; Curry, W.T.; Martuza, R.L.; et al. Single-Cell RNA-Seq Highlights Intratumoral Heterogeneity in Primary Glioblastoma. Science 2014, 344, 1396–1401. [Google Scholar] [CrossRef] [Green Version]
  64. Tirosh, I.; Venteicher, A.S.; Hebert, C.; Escalante, L.E.; Patel, A.P.; Yizhak, K.; Fisher, J.M.; Rodman, C.; Mount, C.; Filbin, M.G.; et al. Single-Cell RNA-Seq Supports a Developmental Hierarchy in Human Oligodendroglioma. Nature 2016, 539, 309–313. [Google Scholar] [CrossRef] [Green Version]
  65. Sharma, S.V.; Lee, D.Y.; Li, B.; Quinlan, M.P.; Takahashi, F.; Maheswaran, S.; McDermott, U.; Azizian, N.; Zou, L.; Fischbach, M.A.; et al. A Chromatin-Mediated Reversible Drug-Tolerant State in Cancer Cell Subpopulations. Cell 2010, 141, 69–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Liau, B.B.; Sievers, C.; Donohue, L.K.; Gillespie, S.M.; Flavahan, W.A.; Miller, T.E.; Venteicher, A.S.; Hebert, C.H.; Carey, C.D.; Rodig, S.J.; et al. Adaptive Chromatin Remodeling Drives Glioblastoma Stem Cell Plasticity and Drug Tolerance. Cell Stem Cell 2017, 20, 233–246.e7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Shi, J.; Wang, Y.; Zeng, L.; Wu, Y.; Deng, J.; Zhang, Q.; Lin, Y.; Li, J.; Kang, T.; Tao, M.; et al. Disrupting the Interaction of BRD4 with Diacetylated Twist Suppresses Tumorigenesis in Basal-like Breast Cancer. Cancer Cell 2014, 25, 210–225. [Google Scholar] [CrossRef] [Green Version]
  68. Stewart, C.A.; Byers, L.A. Altering the Course of Small Cell Lung Cancer: Targeting Cancer Stem Cells via LSD1 Inhibition. Cancer Cell 2015, 28, 4–6. [Google Scholar] [CrossRef] [Green Version]
  69. Daigle, S.R.; Olhava, E.J.; Therkelsen, C.A.; Majer, C.R.; Sneeringer, C.J.; Song, J.; Johnston, L.D.; Scott, M.P.; Smith, J.J.; Xiao, Y.; et al. Selective Killing of Mixed Lineage Leukemia Cells by a Potent Small-Molecule DOT1L Inhibitor. Cancer Cell 2011, 20, 53–65. [Google Scholar] [CrossRef] [Green Version]
  70. Kagohara, L.T.; Zamuner, F.; Davis-Marcisak, E.F.; Sharma, G.; Considine, M.; Allen, J.; Yegnasubramanian, S.; Gaykalova, D.A.; Fertig, E.J. Integrated Single-Cell and Bulk Gene Expression and ATAC-Seq Reveals Heterogeneity and Early Changes in Pathways Associated with Resistance to Cetuximab in HNSCC-Sensitive Cell Lines. Br. J. Cancer 2020, 123, 101–113. [Google Scholar] [CrossRef]
  71. Hinohara, K.; Wu, H.-J.; Vigneau, S.; McDonald, T.O.; Igarashi, K.J.; Yamamoto, K.N.; Madsen, T.; Fassl, A.; Egri, S.B.; Papanastasiou, M.; et al. KDM5 Histone Demethylase Activity Links Cellular Transcriptomic Heterogeneity to Therapeutic Resistance. Cancer Cell 2018, 34, 939–953.e9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Strauss, J.; Figg, W.D. Using Epigenetic Therapy to Overcome Chemotherapy Resistance. Anticancer Res. 2016, 36, 1–4. [Google Scholar]
  73. Morel, D.; Jeffery, D.; Aspeslagh, S.; Almouzni, G.; Postel-Vinay, S. Combining Epigenetic Drugs with Other Therapies for Solid Tumours—Past Lessons and Future Promise. Nat. Rev. Clin. Oncol. 2020, 17, 91–107. [Google Scholar] [CrossRef] [PubMed]
  74. Han, C.Y.; Patten, D.A.; Richardson, R.B.; Harper, M.-E.; Tsang, B.K. Tumor Metabolism Regulating Chemosensitivity in Ovarian Cancer. Genes Cancer 2018, 9, 155–175. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Binnewies, M.; Roberts, E.W.; Kersten, K.; Chan, V.; Fearon, D.F.; Merad, M.; Coussens, L.M.; Gabrilovich, D.I.; Ostrand-Rosenberg, S.; Hedrick, C.C.; et al. Understanding the Tumor Immune Microenvironment (TIME) for Effective Therapy. Nat. Med. 2018, 24, 541–550. [Google Scholar] [CrossRef]
  76. Rada-Iglesias, A.; Enroth, S.; Ameur, A.; Koch, C.M.; Clelland, G.K.; Respuela-Alonso, P.; Wilcox, S.; Dovey, O.M.; Ellis, P.D.; Langford, C.F.; et al. Butyrate Mediates Decrease of Histone Acetylation Centered on Transcription Start Sites and Down-Regulation of Associated Genes. Genome Res. 2007, 17, 708–719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Slaughter, M.J.; Shanle, E.K.; Khan, A.; Chua, K.F.; Hong, T.; Boxer, L.D.; Allis, C.D.; Josefowicz, S.Z.; Garcia, B.A.; Rothbart, S.B.; et al. HDAC Inhibition Results in Widespread Alteration of the Histone Acetylation Landscape and BRD4 Targeting to Gene Bodies. Cell Rep. 2021, 34, 108638. [Google Scholar] [CrossRef] [PubMed]
  78. Hanigan, T.W.; Danes, J.M.; Taha, T.Y.; Frasor, J.; Petukhov, P.A. Histone Deacetylase Inhibitor-Based Chromatin Precipitation for Identification of Targeted Genomic Loci. J. Biol. Methods 2018, 5. [Google Scholar] [CrossRef] [Green Version]
  79. Qu, K.; Zaba, L.C.; Satpathy, A.T.; Giresi, P.G.; Li, R.; Jin, Y.; Armstrong, R.; Jin, C.; Schmitt, N.; Rahbar, Z.; et al. Chromatin Accessibility Landscape of Cutaneous T Cell Lymphoma and Dynamic Response to HDAC Inhibitors. Cancer Cell 2017, 32, 27.e4–41.e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Fiziev, P.; Akdemir, K.C.; Miller, J.P.; Keung, E.Z.; Samant, N.S.; Sharma, S.; Natale, C.A.; Terranova, C.J.; Maitituoheti, M.; Amin, S.B.; et al. Systematic Epigenomic Analysis Reveals Chromatin States Associated with Melanoma Progression. Cell Rep. 2017, 19, 875–889. [Google Scholar] [CrossRef]
  81. Ellis, L.; Pan, Y.; Smyth, G.K.; George, D.J.; McCormack, C.; Williams-Truax, R.; Mita, M.; Beck, J.; Burris, H.; Ryan, G.; et al. Histone Deacetylase Inhibitor Panobinostat Induces Clinical Responses with Associated Alterations in Gene Expression Profiles in Cutaneous T-Cell Lymphoma. Clin. Cancer Res. 2008, 14, 4500–4510. [Google Scholar] [CrossRef] [Green Version]
  82. Peart, M.J.; Smyth, G.K.; van Laar, R.K.; Bowtell, D.D.; Richon, V.M.; Marks, P.A.; Holloway, A.J.; Johnstone, R.W. Identification and Functional Significance of Genes Regulated by Structurally Different Histone Deacetylase Inhibitors. Proc. Natl. Acad. Sci. USA 2005, 102, 3697–3702. [Google Scholar] [CrossRef] [Green Version]
  83. Tang, F.; Yang, Z.; Tan, Y.; Li, Y. Super-Enhancer Function and Its Application in Cancer Targeted Therapy. NPJ Precis. Oncol. 2020, 4, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Gryder, B.E.; Wu, L.; Woldemichael, G.M.; Pomella, S.; Quinn, T.R.; Park, P.M.C.; Cleveland, A.; Stanton, B.Z.; Song, Y.; Rota, R.; et al. Chemical Genomics Reveals Histone Deacetylases Are Required for Core Regulatory Transcription. Nat. Commun. 2019, 10, 3004. [Google Scholar] [CrossRef] [Green Version]
  85. Nagaraja, S.; Vitanza, N.A.; Woo, P.J.; Taylor, K.R.; Liu, F.; Zhang, L.; Li, M.; Meng, W.; Ponnuswami, A.; Sun, W.; et al. Transcriptional Dependencies in Diffuse Intrinsic Pontine Glioma. Cancer Cell 2017, 31, 635–652.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Nguyen, T.T.T.; Zhang, Y.; Shang, E.; Shu, C.; Torrini, C.; Zhao, J.; Bianchetti, E.; Mela, A.; Humala, N.; Mahajan, A.; et al. HDAC Inhibitors Elicit Metabolic Reprogramming by Targeting Super-Enhancers in Glioblastoma Models. J. Clin. Investig. 2020, 130, 3699–3716. [Google Scholar] [CrossRef] [PubMed]
  87. Donati, B.; Lorenzini, E.; Ciarrocchi, A. BRD4 and Cancer: Going beyond Transcriptional Regulation. Mol. Cancer 2018, 17, 164. [Google Scholar] [CrossRef] [PubMed]
  88. Kim, Y.J.; Greer, C.B.; Cecchini, K.R.; Harris, L.N.; Tuck, D.P.; Kim, T.H. HDAC Inhibitors Induce Transcriptional Repression of High Copy Number Genes in Breast Cancer through Elongation Blockade. Oncogene 2013, 32, 2828–2835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Vaid, R.; Wen, J.; Mannervik, M. Release of Promoter–Proximal Paused Pol II in Response to Histone Deacetylase Inhibition. Nucleic Acids Res. 2020, 48, 4877–4890. [Google Scholar] [CrossRef]
  90. Mishra, V.K.; Wegwitz, F.; Kosinsky, R.L.; Sen, M.; Baumgartner, R.; Wulff, T.; Siveke, J.T.; Schildhaus, H.-U.; Najafova, Z.; Kari, V.; et al. Histone Deacetylase Class-I Inhibition Promotes Epithelial Gene Expression in Pancreatic Cancer Cells in a BRD4- and MYC-Dependent Manner. Nucleic Acids Res. 2017, 45, 6334–6349. [Google Scholar] [CrossRef] [PubMed]
  91. Gou, L.-Y.; Li, A.-N.; Yang, J.-J.; Zhang, X.-C.; Su, J.; Yan, H.-H.; Xie, Z.; Lou, N.-N.; Liu, S.-Y.; Dong, Z.-Y.; et al. The Coexistence of MET Over-Expression and an EGFR T790M Mutation Is Related to Acquired Resistance to EGFR Tyrosine Kinase Inhibitors in Advanced Non-Small Cell Lung Cancer. Oncotarget 2016, 7, 51311–51319. [Google Scholar] [CrossRef] [Green Version]
  92. Turke, A.B.; Zejnullahu, K.; Wu, Y.-L.; Song, Y.; Dias-Santagata, D.; Lifshits, E.; Toschi, L.; Rogers, A.; Mok, T.; Sequist, L.; et al. Preexistence and Clonal Selection of MET Amplification in EGFR Mutant NSCLC. Cancer Cell 2010, 17, 77–88. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Ramzy, M.M.; Abdelghany, H.M.; Zenhom, N.M.; El-Tahawy, N.F. Effect of Histone Deacetylase Inhibitor on Epithelial-Mesenchymal Transition of Liver Fibrosis. IUBMB Life 2018, 70, 511–518. [Google Scholar] [CrossRef] [Green Version]
  94. Witta, S.E.; Jotte, R.M.; Konduri, K.; Neubauer, M.A.; Spira, A.I.; Ruxer, R.L.; Varella-Garcia, M.; Bunn, P.A.; Hirsch, F.R. Randomized Phase II Trial of Erlotinib With and Without Entinostat in Patients With Advanced Non–Small-Cell Lung Cancer Who Progressed on Prior Chemotherapy. J. Clin. Oncol. 2012, 30, 2248–2255. [Google Scholar] [CrossRef]
  95. Karakaidos, P.; Karagiannis, D.; Rampias, T. Resolving DNA Damage: Epigenetic Regulation of DNA Repair. Molecules 2020, 25, 2496. [Google Scholar] [CrossRef]
  96. Bhaskara, S.; Knutson, S.K.; Jiang, G.; Chandrasekharan, M.B.; Wilson, A.J.; Zheng, S.; Yenamandra, A.; Locke, K.; Yuan, J.; Bonine-Summers, A.R.; et al. Hdac3 Is Essential for the Maintenance of Chromatin Structure and Genome Stability. Cancer Cell 2010, 18, 436–447. [Google Scholar] [CrossRef] [Green Version]
  97. Bhaskara, S.; Chyla, B.J.; Amann, J.M.; Knutson, S.K.; Cortez, D.; Sun, Z.-W.; Hiebert, S.W. Deletion of Histone Deacetylase 3 Reveals Critical Roles in S Phase Progression and DNA Damage Control. Mol. Cell 2008, 30, 61–72. [Google Scholar] [CrossRef] [Green Version]
  98. Eot-Houllier, G.; Fulcrand, G.; Watanabe, Y.; Magnaghi-Jaulin, L.; Jaulin, C. Histone Deacetylase 3 Is Required for Centromeric H3K4 Deacetylation and Sister Chromatid Cohesion. Genes Dev. 2008, 22, 2639–2644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Miller, K.M.; Tjeertes, J.V.; Coates, J.; Legube, G.; Polo, S.E.; Britton, S.; Jackson, S.P. Human HDAC1 and HDAC2 Function in the DNA-Damage Response to Promote DNA Nonhomologous End-Joining. Nat. Struct. Mol. Biol. 2010, 17, 1144–1151. [Google Scholar] [CrossRef]
  100. Bhaskara, S.; Jacques, V.; Rusche, J.R.; Olson, E.N.; Cairns, B.R.; Chandrasekharan, M.B. Histone Deacetylases 1 and 2 Maintain S-Phase Chromatin and DNA Replication Fork Progression. Epigenet. Chromatin 2013, 6, 27. [Google Scholar] [CrossRef] [Green Version]
  101. Kao, G.D.; McKenna, W.G.; Guenther, M.G.; Muschel, R.J.; Lazar, M.A.; Yen, T.J. Histone Deacetylase 4 Interacts with 53BP1 to Mediate the DNA Damage Response. J. Cell Biol. 2003, 160, 1017–1027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Conti, C.; Leo, E.; Eichler, G.S.; Sordet, O.; Martin, M.M.; Fan, A.; Aladjem, M.I.; Pommier, Y. Inhibition of Histone Deacetylase in Cancer Cells Slows down Replication Forks, Activates Dormant Origins, and Induces DNA Damage. Cancer Res. 2010, 70, 4470–4480. [Google Scholar] [CrossRef] [Green Version]
  103. Munshi, A.; Kurland, J.F.; Nishikawa, T.; Tanaka, T.; Hobbs, M.L.; Tucker, S.L.; Ismail, S.; Stevens, C.; Meyn, R.E. Histone Deacetylase Inhibitors Radiosensitize Human Melanoma Cells by Suppressing DNA Repair Activity. Clin. Cancer Res. 2005, 11, 4912–4922. [Google Scholar] [CrossRef] [Green Version]
  104. Thurn, K.T.; Thomas, S.; Raha, P.; Qureshi, I.; Munster, P.N. Histone Deacetylase Regulation of ATM-Mediated DNA Damage Signaling. Mol. Cancer Ther. 2013, 12, 2078–2087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Kachhap, S.K.; Rosmus, N.; Collis, S.J.; Kortenhorst, M.S.Q.; Wissing, M.D.; Hedayati, M.; Shabbeer, S.; Mendonca, J.; Deangelis, J.; Marchionni, L.; et al. Downregulation of Homologous Recombination DNA Repair Genes by HDAC Inhibition in Prostate Cancer Is Mediated through the E2F1 Transcription Factor. PLoS ONE 2010, 5, e11208. [Google Scholar] [CrossRef] [PubMed]
  106. Robert, C.; Nagaria, P.K.; Pawar, N.; Adewuyi, A.; Gojo, I.; Meyers, D.J.; Cole, P.A.; Rassool, F.V. Histone Deacetylase Inhibitors Decrease NHEJ Both by Acetylation of Repair Factors and Trapping of PARP1 at DNA Double-Strand Breaks in Chromatin. Leuk Res. 2016, 45, 14–23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Gaymes, T.J.; Padua, R.A.; Pla, M.; Orr, S.; Omidvar, N.; Chomienne, C.; Mufti, G.J.; Rassool, F.V. Histone Deacetylase Inhibitors (HDI) Cause DNA Damage in Leukemia Cells: A Mechanism for Leukemia-Specific HDI-Dependent Apoptosis? Mol. Cancer Res. 2006, 4, 563–573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Ruefli, A.A.; Ausserlechner, M.J.; Bernhard, D.; Sutton, V.R.; Tainton, K.M.; Kofler, R.; Smyth, M.J.; Johnstone, R.W. The Histone Deacetylase Inhibitor and Chemotherapeutic Agent Suberoylanilide Hydroxamic Acid (SAHA) Induces a Cell-Death Pathway Characterized by Cleavage of Bid and Production of Reactive Oxygen Species. Proc. Natl. Acad. Sci. USA 2001, 98, 10833–10838. [Google Scholar] [CrossRef] [Green Version]
  109. Petruccelli, L.A.; Dupéré-Richer, D.; Pettersson, F.; Retrouvey, H.; Skoulikas, S.; Jr, W.H.M. Vorinostat Induces Reactive Oxygen Species and DNA Damage in Acute Myeloid Leukemia Cells. PLoS ONE 2011, 6, e20987. [Google Scholar] [CrossRef] [Green Version]
  110. Rosato, R.R.; Almenara, J.A.; Grant, S. The Histone Deacetylase Inhibitor MS-275 Promotes Differentiation or Apoptosis in Human Leukemia Cells through a Process Regulated by Generation of Reactive Oxygen Species and Induction of P21CIP1/WAF1 1. Cancer Res. 2003, 63, 3637–3645. [Google Scholar]
  111. Nikolova, T.; Kiweler, N.; Krämer, O.H. Interstrand Crosslink Repair as a Target for HDAC Inhibition. Trends Pharmacol. Sci. 2017, 38, 822–836. [Google Scholar] [CrossRef]
  112. Bouwman, P.; Jonkers, J. The Effects of Deregulated DNA Damage Signalling on Cancer Chemotherapy Response and Resistance. Nat. Rev. Cancer 2012, 12, 587–598. [Google Scholar] [CrossRef]
  113. Lord, C.J.; Ashworth, A. BRCAness Revisited. Nat. Rev. Cancer 2016, 16, 110–120. [Google Scholar] [CrossRef] [PubMed]
  114. Chen, J. The Cell-Cycle Arrest and Apoptotic Functions of P53 in Tumor Initiation and Progression. Cold Spring Harb. Perspect. Med. 2016, 6. [Google Scholar] [CrossRef] [PubMed]
  115. Richon, V.M.; Sandhoff, T.W.; Rifkind, R.A.; Marks, P.A. Histone Deacetylase Inhibitor Selectively Induces P21WAF1 Expression and Gene-Associated Histone Acetylation. Proc. Natl. Acad. Sci. USA 2000, 97, 10014–10019. [Google Scholar] [CrossRef] [Green Version]
  116. Shio, S.; Kodama, Y.; Ida, H.; Shiokawa, M.; Kitamura, K.; Hatano, E.; Uemoto, S.; Chiba, T. Loss of RUNX3 Expression by Histone Deacetylation Is Associated with Biliary Tract Carcinogenesis. Cancer Sci. 2011, 102, 776–783. [Google Scholar] [CrossRef] [PubMed]
  117. Yamaguchi, T.; Cubizolles, F.; Zhang, Y.; Reichert, N.; Kohler, H.; Seiser, C.; Matthias, P. Histone Deacetylases 1 and 2 Act in Concert to Promote the G1-to-S Progression. Genes Dev. 2010, 24, 455–469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Zupkovitz, G.; Grausenburger, R.; Brunmeir, R.; Senese, S.; Tischler, J.; Jurkin, J.; Rembold, M.; Meunier, D.; Egger, G.; Lagger, S.; et al. The Cyclin-Dependent Kinase Inhibitor P21 Is a Crucial Target for Histone Deacetylase 1 as a Regulator of Cellular Proliferation. Mol. Cell. Biol. 2010, 30, 1171–1181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Zhao, Y.; Lu, S.; Wu, L.; Chai, G.; Wang, H.; Chen, Y.; Sun, J.; Yu, Y.; Zhou, W.; Zheng, Q.; et al. Acetylation of P53 at Lysine 373/382 by the Histone Deacetylase Inhibitor Depsipeptide Induces Expression of P21Waf1/Cip1. Mol. Cell. Biol. 2006, 26, 2782–2790. [Google Scholar] [CrossRef] [Green Version]
  120. Burgess, A.J.; Pavey, S.; Warrener, R.; Hunter, L.-J.K.; Piva, T.J.; Musgrove, E.A.; Saunders, N.; Parsons, P.G.; Gabrielli, B.G. Up-Regulation of P21WAF1/CIP1 by Histone Deacetylase Inhibitors Reduces Their Cytotoxicity. Mol. Pharmacol. 2001, 60, 828–837. [Google Scholar]
  121. Qiu, L.; Burgess, A.; Fairlie, D.P.; Leonard, H.; Parsons, P.G.; Gabrielli, B.G. Histone Deacetylase Inhibitors Trigger a G2 Checkpoint in Normal Cells That Is Defective in Tumor Cells. MBoC 2000, 11, 2069–2083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Warrener, R.; Beamish, H.; Burgess, A.; Waterhouse, N.J.; Giles, N.; Fairlie, D.P.; Gabrielli, B. Tumor Cell-Specific Cytotoxicity by Targeting Cell Cycle Checkpoints. FASEB J. 2003, 17, 1–21. [Google Scholar] [CrossRef]
  123. Vidal-Laliena, M.; Gallastegui, E.; Mateo, F.; Martínez-Balbás, M.; Pujol, M.J.; Bachs, O. Histone Deacetylase 3 Regulates Cyclin A Stability. J. Biol. Chem. 2013, 288, 21096–21104. [Google Scholar] [CrossRef] [Green Version]
  124. Davis, J.E.; Kirk, J.; Ji, Y.; Tang, D.G. Tumor Dormancy and Slow-Cycling Cancer Cells. Adv. Exp. Med. Biol. 2019, 1164, 199–206. [Google Scholar] [CrossRef]
  125. Das, C.M.; Aguilera, D.; Vasquez, H.; Prasad, P.; Zhang, M.; Wolff, J.E.; Gopalakrishnan, V. Valproic Acid Induces P21 and Topoisomerase-II (α/β) Expression and Synergistically Enhances Etoposide Cytotoxicity in Human Glioblastoma Cell Lines. J. Neurooncol 2007, 85, 159–170. [Google Scholar] [CrossRef]
  126. Marchion, D.C.; Bicaku, E.; Turner, J.G.; Daud, A.I.; Sullivan, D.M.; Munster, P.N. Synergistic Interaction between Histone Deacetylase and Topoisomerase II Inhibitors Is Mediated through Topoisomerase IIβ. Clin. Cancer Res. 2005, 11, 8467–8475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Munster, P.; Marchion, D.; Bicaku, E.; Lacevic, M.; Kim, J.; Centeno, B.; Daud, A.; Neuger, A.; Minton, S.; Sullivan, D. Clinical and Biological Effects of Valproic Acid as a Histone Deacetylase Inhibitor on Tumor and Surrogate Tissues: Phase I/II Trial of Valproic Acid and Epirubicin/FEC. Clin. Cancer Res. 2009, 15, 2488–2496. [Google Scholar] [CrossRef] [Green Version]
  128. Munster, P.N.; Marchion, D.; Thomas, S.; Egorin, M.; Minton, S.; Springett, G.; Lee, J.-H.; Simon, G.; Chiappori, A.; Sullivan, D.; et al. Phase I Trial of Vorinostat and Doxorubicin in Solid Tumours: Histone Deacetylase 2 Expression as a Predictive Marker. Br. J. Cancer 2009, 101, 1044–1050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Valentini, A.; Gravina, P.; Federici, G.; Bernardini, S. Valproic Acid Induces Apoptosis, P16INK4A Upregulation and Sensitization to Chemotherapy in Human Melanoma Cells. Cancer Biol. Ther. 2007, 6, 185–191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Ramalingam, S.S.; Maitland, M.L.; Frankel, P.; Argiris, A.E.; Koczywas, M.; Gitlitz, B.; Thomas, S.; Espinoza-Delgado, I.; Vokes, E.E.; Gandara, D.R.; et al. Carboplatin and Paclitaxel in Combination With Either Vorinostat or Placebo for First-Line Therapy of Advanced Non–Small-Cell Lung Cancer. J. Clin. Oncol. 2010, 28, 56–62. [Google Scholar] [CrossRef] [Green Version]
  131. Powathil, G.G.; Gordon, K.E.; Hill, L.A.; Chaplain, M.A.J. Modelling the Effects of Cell-Cycle Heterogeneity on the Response of a Solid Tumour to Chemotherapy: Biological Insights from a Hybrid Multiscale Cellular Automaton Model. J. Theor. Biol. 2012, 308, 1–19. [Google Scholar] [CrossRef]
  132. Gallaher, J.A.; Massey, S.C.; Hawkins-Daarud, A.; Noticewala, S.S.; Rockne, R.C.; Johnston, S.K.; Gonzalez-Cuyar, L.; Juliano, J.; Gil, O.; Swanson, K.R.; et al. From Cells to Tissue: How Cell Scale Heterogeneity Impacts Glioblastoma Growth and Treatment Response. PLoS Comput. Biol. 2020, 16, e1007672. [Google Scholar] [CrossRef] [Green Version]
  133. Alao, J.P.; Lam, E.W.-F.; Ali, S.; Buluwela, L.; Bordogna, W.; Lockey, P.; Varshochi, R.; Stavropoulou, A.V.; Coombes, R.C.; Vigushin, D.M. Histone Deacetylase Inhibitor Trichostatin A Represses Estrogen Receptor α-Dependent Transcription and Promotes Proteasomal Degradation of Cyclin D1 in Human Breast Carcinoma Cell Lines. Clin. Cancer Res. 2004, 10, 8094–8104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Fan, J.; Yin, W.-J.; Lu, J.-S.; Wang, L.; Wu, J.; Wu, F.-Y.; Di, G.-H.; Shen, Z.-Z.; Shao, Z.-M. ERα Negative Breast Cancer Cells Restore Response to Endocrine Therapy by Combination Treatment with Both HDAC Inhibitor and DNMT Inhibitor. J. Cancer Res. Clin. Oncol. 2008, 134, 883–890. [Google Scholar] [CrossRef] [PubMed]
  135. Elmore, S. Apoptosis: A Review of Programmed Cell Death. Toxicol. Pathol. 2007, 35, 495–516. [Google Scholar] [CrossRef]
  136. Riley, J.S.; Hutchinson, R.; McArt, D.G.; Crawford, N.; Holohan, C.; Paul, I.; Van Schaeybroeck, S.; Salto-Tellez, M.; Johnston, P.G.; Fennell, D.A.; et al. Prognostic and Therapeutic Relevance of FLIP and Procaspase-8 Overexpression in Non-Small Cell Lung Cancer. Cell Death Dis. 2013, 4, e951. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Ungerstedt, J.S.; Sowa, Y.; Xu, W.-S.; Shao, Y.; Dokmanovic, M.; Perez, G.; Ngo, L.; Holmgren, A.; Jiang, X.; Marks, P.A. Role of Thioredoxin in the Response of Normal and Transformed Cells to Histone Deacetylase Inhibitors. Proc. Natl. Acad. Sci. USA 2005, 102, 673–678. [Google Scholar] [CrossRef] [Green Version]
  138. Jing, B.; Jin, J.; Xiang, R.; Liu, M.; Yang, L.; Tong, Y.; Xiao, X.; Lei, H.; Liu, W.; Xu, H.; et al. Vorinostat and Quinacrine Have Synergistic Effects in T-Cell Acute Lymphoblastic Leukemia through Reactive Oxygen Species Increase and Mitophagy Inhibition. Cell Death Dis. 2018, 9, 1–11. [Google Scholar] [CrossRef]
  139. Bolden, J.E.; Shi, W.; Jankowski, K.; Kan, C.-Y.; Cluse, L.; Martin, B.P.; MacKenzie, K.L.; Smyth, G.K.; Johnstone, R.W. HDAC Inhibitors Induce Tumor-Cell-Selective pro-Apoptotic Transcriptional Responses. Cell Death Dis. 2013, 4, e519. [Google Scholar] [CrossRef] [Green Version]
  140. Fandy, T.E.; Srivastava, R.K. Trichostatin A Sensitizes TRAIL-Resistant Myeloma Cells by Downregulation of the Antiapoptotic Bcl-2 Proteins. Cancer Chemother. Pharmacol. 2006, 58, 471–477. [Google Scholar] [CrossRef]
  141. Juan, L.J.; Shia, W.J.; Chen, M.H.; Yang, W.M.; Seto, E.; Lin, Y.S.; Wu, C.W. Histone Deacetylases Specifically Down-Regulate P53-Dependent Gene Activation. J. Biol. Chem. 2000, 275, 20436–20443. [Google Scholar] [CrossRef] [Green Version]
  142. Ding, G.; Liu, H.-D.; Huang, Q.; Liang, H.-X.; Ding, Z.-H.; Liao, Z.-J.; Huang, G. HDAC6 Promotes Hepatocellular Carcinoma Progression by Inhibiting P53 Transcriptional Activity. FEBS Lett. 2013, 587, 880–886. [Google Scholar] [CrossRef] [Green Version]
  143. Yan, W.; Liu, S.; Xu, E.; Zhang, J.; Zhang, Y.; Chen, X.; Chen, X. Histone Deacetylase Inhibitors Suppress Mutant P53 Transcription via Histone Deacetylase 8. Oncogene 2013, 32, 599–609. [Google Scholar] [CrossRef] [Green Version]
  144. Yu, X.; Guo, Z.S.; Marcu, M.G.; Neckers, L.; Nguyen, D.M.; Chen, G.A.; Schrump, D.S. Modulation of P53, ErbB1, ErbB2, and Raf-1 Expression in Lung Cancer Cells by Depsipeptide FR901228. J. Natl. Cancer Inst. 2002, 94, 504–513. [Google Scholar] [CrossRef] [Green Version]
  145. Sonnemann, J.; Marx, C.; Becker, S.; Wittig, S.; Palani, C.D.; Krämer, O.H.; Beck, J.F. P53-Dependent and P53-Independent Anticancer Effects of Different Histone Deacetylase Inhibitors. Br. J. Cancer 2014, 110, 656–667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Shao, Y.; Gao, Z.; Marks, P.A.; Jiang, X. Apoptotic and Autophagic Cell Death Induced by Histone Deacetylase Inhibitors. Proc. Natl. Acad. Sci. USA 2004, 101, 18030–18035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Stankov, M.V.; El Khatib, M.; Kumar Thakur, B.; Heitmann, K.; Panayotova-Dimitrova, D.; Schoening, J.; Bourquin, J.P.; Schweitzer, N.; Leverkus, M.; Welte, K.; et al. Histone Deacetylase Inhibitors Induce Apoptosis in Myeloid Leukemia by Suppressing Autophagy. Leukemia 2014, 28, 577–588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Natarajan, U.; Venkatesan, T.; Radhakrishnan, V.; Samuel, S.; Rathinavelu, A. Differential Mechanisms of Cell Death Induced by HDAC Inhibitor SAHA and MDM2 Inhibitor RG7388 in MCF-7 Cells. Cells 2018, 8, 8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Gong, P.; Wang, Y.; Jing, Y. Apoptosis Induction ByHistone Deacetylase Inhibitors in Cancer Cells: Role of Ku70. Int. J. Mol. Sci. 2019, 20, 1601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Meng, J.; Zhang, F.; Zhang, X.-T.; Zhang, T.; Li, Y.-H.; Fan, L.; Sun, Y.; Zhang, H.-L.; Mei, Q.-B. Ku70 Is Essential for Histone Deacetylase Inhibitor Trichostatin A-Induced Apoptosis. Mol. Med. Rep. 2015, 12, 581–586. [Google Scholar] [CrossRef] [Green Version]
  151. Subramanian, C.; Opipari, A.W.; Bian, X.; Castle, V.P.; Kwok, R.P.S. Ku70 Acetylation Mediates Neuroblastoma Cell Death Induced by Histone Deacetylase Inhibitors. Proc. Natl. Acad. Sci. USA 2005, 102, 4842–4847. [Google Scholar] [CrossRef] [Green Version]
  152. Cohen, H.Y.; Lavu, S.; Bitterman, K.J.; Hekking, B.; Imahiyerobo, T.A.; Miller, C.; Frye, R.; Ploegh, H.; Kessler, B.M.; Sinclair, D.A. Acetylation of the C Terminus of Ku70 by CBP and PCAF Controls Bax-Mediated Apoptosis. Mol. Cell 2004, 13, 627–638. [Google Scholar] [CrossRef]
  153. Molife, L.R.; Attard, G.; Fong, P.C.; Karavasilis, V.; Reid, A.H.M.; Patterson, S.; Riggs, C.E.; Higano, C.; Stadler, W.M.; McCulloch, W.; et al. Phase II, Two-Stage, Single-Arm Trial of the Histone Deacetylase Inhibitor (HDACi) Romidepsin in Metastatic Castration-Resistant Prostate Cancer (CRPC). Ann. Oncol. 2010, 21, 109–113. [Google Scholar] [CrossRef]
  154. Mackay, H.J.; Hirte, H.; Colgan, T.; Covens, A.; MacAlpine, K.; Grenci, P.; Wang, L.; Mason, J.; Pham, P.-A.; Tsao, M.-S.; et al. 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 2010, 46, 1573–1579. [Google Scholar] [CrossRef] [Green Version]
  155. Vansteenkiste, J.; Van Cutsem, E.; Dumez, H.; Chen, C.; Ricker, J.L.; Randolph, S.S.; Schöffski, P. Early Phase II Trial of Oral Vorinostat in Relapsed or Refractory Breast, Colorectal, or Non-Small Cell Lung Cancer. Investig. New Drugs 2008, 26, 483–488. [Google Scholar] [CrossRef]
  156. Modesitt, S.C.; Sill, M.; Hoffman, J.S.; Bender, D.P. A Phase II Study of Vorinostat in the Treatment of Persistent or Recurrent Epithelial Ovarian or Primary Peritoneal Carcinoma: A Gynecologic Oncology Group Study. Gynecol. Oncol. 2008, 109, 182–186. [Google Scholar] [CrossRef]
  157. McClure, J.J.; Li, X.; Chou, C.J. Chapter Six—Advances and Challenges of HDAC Inhibitors in Cancer Therapeutics. In Advances in Cancer Research; Tew, K.D., Fisher, P.B., Eds.; Advances in Cancer Research; Academic Press: Cambridge, MA, USA, 2018; Volume 138, pp. 183–211. [Google Scholar]
  158. Chen, I.-C.; Sethy, B.; Liou, J.-P. Recent Update of HDAC Inhibitors in Lymphoma. Front. Cell Dev. Biol. 2020, 8, 576391. [Google Scholar] [CrossRef] [PubMed]
  159. Di Gennaro, E.; Bruzzese, F.; Pepe, S.; Leone, A.; Delrio, P.; Subbarayan, P.R.; Avallone, A.; Budillon, A. Modulation of Thymidilate Synthase and P53 Expression by HDAC Inhibitor Vorinostat Resulted in Synergistic Antitumor Effect in Combination with 5FU or Raltitrexed. Cancer Biol. Ther. 2009, 8, 782–791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Terranova-Barberio, M.; Pecori, B.; Roca, M.S.; Imbimbo, S.; Bruzzese, F.; Leone, A.; Muto, P.; Delrio, P.; Avallone, A.; Budillon, A.; et al. Synergistic Antitumor Interaction between Valproic Acid, Capecitabine and Radiotherapy in Colorectal Cancer: Critical Role of P53. J. Exp. Clin. Cancer Res. 2017, 36, 177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Inde, Z.; Dixon, S.J. The Impact of Non-Genetic Heterogeneity on Cancer Cell Death. Crit. Rev. Biochem. Mol. Biol. 2018, 53, 99–114. [Google Scholar] [CrossRef]
  162. Ogden, A.; Rida, P.C.; Reid, M.D.; Kucuk, O.; Aneja, R. Die-Hard Survivors: Heterogeneity in Apoptotic Thresholds May Underlie Chemoresistance. Expert Rev. Anticancer Ther. 2015, 15, 277–281. [Google Scholar] [CrossRef] [Green Version]
  163. Roux, J.; Hafner, M.; Bandara, S.; Sims, J.J.; Hudson, H.; Chai, D.; Sorger, P.K. Fractional Killing Arises from Cell-to-Cell Variability in Overcoming a Caspase Activity Threshold. Mol. Syst. Biol. 2015, 11, 803. [Google Scholar] [CrossRef]
  164. Alcarraz-Vizán, G.; Boren, J.; Lee, W.-N.P.; Cascante, M. Histone Deacetylase Inhibition Results in a Common Metabolic Profile Associated with HT29 Differentiation. Metabolomics 2010, 6, 229–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Yang, J.; Jin, X.; Yan, Y.; Shao, Y.; Pan, Y.; Roberts, L.R.; Zhang, J.; Huang, H.; Jiang, J. Inhibiting Histone Deacetylases Suppresses Glucose Metabolism and Hepatocellular Carcinoma Growth by Restoring FBP1 Expression. Sci. Rep. 2017, 7, 43864. [Google Scholar] [CrossRef]
  166. Wardell, S.E.; Ilkayeva, O.R.; Wieman, H.L.; Frigo, D.E.; Rathmell, J.C.; Newgard, C.B.; McDonnell, D.P. Glucose Metabolism as a Target of Histone Deacetylase Inhibitors. Mol. Endocrinol. 2009, 23, 388–401. [Google Scholar] [CrossRef] [Green Version]
  167. Ferrari, A.; Longo, R.; Fiorino, E.; Silva, R.; Mitro, N.; Cermenati, G.; Gilardi, F.; Desvergne, B.; Andolfo, A.; Magagnotti, C.; et al. HDAC3 Is a Molecular Brake of the Metabolic Switch Supporting White Adipose Tissue Browning. Nat. Commun. 2017, 8, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Song, S.; Wen, Y.; Tong, H.; Loro, E.; Gong, Y.; Liu, J.; Hong, S.; Li, L.; Khurana, T.S.; Chu, M.; et al. The HDAC3 Enzymatic Activity Regulates Skeletal Muscle Fuel Metabolism. J. Mol. Cell Biol. 2019, 11, 133–143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Egler, V.; Korur, S.; Failly, M.; Boulay, J.-L.; Imber, R.; Lino, M.M.; Merlo, A. Histone Deacetylase Inhibition and Blockade of the Glycolytic Pathway Synergistically Induce Glioblastoma Cell Death. Clin. Cancer Res. 2008, 14, 3132–3140. [Google Scholar] [CrossRef] [Green Version]
  170. Faubert, B.; Solmonson, A.; DeBerardinis, R.J. Metabolic Reprogramming and Cancer Progression. Science 2020, 368. [Google Scholar] [CrossRef] [PubMed]
  171. Lu, C.; Zhang, K.; Zhang, Y.; Tan, M.; Li, Y.; He, X.; Zhang, Y. Preparation and Characterization of Vorinostat-Coated Beads for Profiling of Novel Target Proteins. J. Chromatogr. A 2014, 1372, 34–41. [Google Scholar] [CrossRef]
  172. Luengo, A.; Gui, D.Y.; Vander Heiden, M.G. Targeting Metabolism for Cancer Therapy. Cell Chem. Biol. 2017, 24, 1161–1180. [Google Scholar] [CrossRef] [Green Version]
  173. Kim, J.; DeBerardinis, R.J. Mechanisms and Implications of Metabolic Heterogeneity in Cancer. Cell Metab. 2019, 30, 434–446. [Google Scholar] [CrossRef] [PubMed]
  174. Kondo, H.; Ratcliffe, C.D.H.; Hooper, S.; Ellis, J.; MacRae, J.I.; Hennequart, M.; Dunsby, C.W.; Anderson, K.I.; Sahai, E. Single-Cell Resolved Imaging Reveals Intra-Tumor Heterogeneity in Glycolysis, Transitions between Metabolic States, and Their Regulatory Mechanisms. Cell Rep. 2021, 34, 108750. [Google Scholar] [CrossRef] [PubMed]
  175. Quail, D.F.; Joyce, J.A. Microenvironmental Regulation of Tumor Progression and Metastasis. Nat. Med. 2013, 19, 1423–1437. [Google Scholar] [CrossRef]
  176. Aspeslagh, S.; Morel, D.; Soria, J.-C.; Postel-Vinay, S. Epigenetic Modifiers as New Immunomodulatory Therapies in Solid Tumours. Ann. Oncol. 2018, 29, 812–824. [Google Scholar] [CrossRef] [PubMed]
  177. Conte, M.; De Palma, R.; Altucci, L. HDAC Inhibitors as Epigenetic Regulators for Cancer Immunotherapy. Int. J. Biochem. Cell Biol. 2018, 98, 65–74. [Google Scholar] [CrossRef]
  178. Licciardi, P.V.; Karagiannis, T.C. Regulation of Immune Responses by Histone Deacetylase Inhibitors. ISRN Hematol. 2012, 2012, 690901. [Google Scholar] [CrossRef] [Green Version]
  179. Setiadi, A.F.; Omilusik, K.; David, M.D.; Seipp, R.P.; Hartikainen, J.; Gopaul, R.; Choi, K.B.; Jefferies, W.A. Epigenetic Enhancement of Antigen Processing and Presentation Promotes Immune Recognition of Tumors. Cancer Res. 2008, 68, 9601–9607. [Google Scholar] [CrossRef] [Green Version]
  180. Skov, S.; Pedersen, M.T.; Andresen, L.; Straten, P.T.; Woetmann, A.; Ødum, N. Cancer Cells Become Susceptible to Natural Killer Cell Killing after Exposure to Histone Deacetylase Inhibitors Due to Glycogen Synthase Kinase-3–Dependent Expression of MHC Class I–Related Chain A and B. Cancer Res. 2005, 65, 11136–11145. [Google Scholar] [CrossRef] [Green Version]
  181. Murakami, T.; Sato, A.; Chun, N.A.L.; Hara, M.; Naito, Y.; Kobayashi, Y.; Kano, Y.; Ohtsuki, M.; Furukawa, Y.; Kobayashi, E. Transcriptional Modulation Using HDACi Depsipeptide Promotes Immune Cell-Mediated Tumor Destruction of Murine B16 Melanoma. J. Investig. Dermatol. 2008, 128, 1506–1516. [Google Scholar] [CrossRef] [Green Version]
  182. Ashburner, B.P.; Westerheide, S.D.; Baldwin, A.S. The P65 (RelA) Subunit of NF-KappaB Interacts with the Histone Deacetylase (HDAC) Corepressors HDAC1 and HDAC2 to Negatively Regulate Gene Expression. Mol. Cell. Biol. 2001, 21, 7065–7077. [Google Scholar] [CrossRef] [Green Version]
  183. Chen, L.; Fischle, W.; Verdin, E.; Greene, W.C. Duration of Nuclear NF-ΚB Action Regulated by Reversible Acetylation. Science 2001, 293, 1653–1657. [Google Scholar] [CrossRef] [Green Version]
  184. Brogdon, J.L.; Xu, Y.; Szabo, S.J.; An, S.; Buxton, F.; Cohen, D.; Huang, Q. Histone Deacetylase Activities Are Required for Innate Immune Cell Control of Th1 but Not Th2 Effector Cell Function. Blood 2007, 109, 1123–1130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Moreira, J.M.A.; Scheipers, P.; Sørensen, P. The Histone Deacetylase Inhibitor Trichostatin A Modulates CD4+ T Cell Responses. BMC Cancer 2003, 3, 30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Li, X.; Su, X.; Liu, R.; Pan, Y.; Fang, J.; Cao, L.; Feng, C.; Shang, Q.; Chen, Y.; Shao, C.; et al. HDAC Inhibition Potentiates Anti-Tumor Activity of Macrophages and Enhances Anti-PD-L1-Mediated Tumor Suppression. Oncogene 2021, 40, 1836–1850. [Google Scholar] [CrossRef] [PubMed]
  187. Valkenburg, K.C.; de Groot, A.E.; Pienta, K.J. Targeting the Tumour Stroma to Improve Cancer Therapy. Nat. Rev. Clin. Oncol. 2018, 15, 366–381. [Google Scholar] [CrossRef] [PubMed]
  188. Ping, Q.; Yan, R.; Cheng, X.; Wang, W.; Zhong, Y.; Hou, Z.; Shi, Y.; Wang, C.; Li, R. Cancer-Associated Fibroblasts: Overview, Progress, Challenges, and Directions. Cancer Gene Ther. 2021, 1–16. [Google Scholar] [CrossRef]
  189. Kim, D.J.; Dunleavey, J.M.; Xiao, L.; Ollila, D.W.; Troester, M.A.; Otey, C.A.; Li, W.; Barker, T.H.; Dudley, A.C. Suppression of TGFβ-Mediated Conversion of Endothelial Cells and Fibroblasts into Cancer Associated (Myo)Fibroblasts via HDAC Inhibition. Br. J. Cancer 2018, 118, 1359–1368. [Google Scholar] [CrossRef] [Green Version]
  190. Nguyen, A.H.; Elliott, I.A.; Wu, N.; Matsumura, C.; Vogelauer, M.; Attar, N.; Dann, A.; Ghukasyan, R.; Toste, P.A.; Patel, S.G.; et al. Histone Deacetylase Inhibitors Provoke a Tumor Supportive Phenotype in Pancreatic Cancer Associated Fibroblasts. Oncotarget 2016, 8, 19074–19088. [Google Scholar] [CrossRef] [Green Version]
  191. Wang, Z.; Dabrosin, C.; Yin, X.; Fuster, M.M.; Arreola, A.; Rathmell, W.K.; Generali, D.; Nagaraju, G.P.; El-Rayes, B.; Ribatti, D.; et al. Broad Targeting of Angiogenesis for Cancer Prevention and Therapy. Semin. Cancer Biol. 2015, 35, S224–S243. [Google Scholar] [CrossRef]
  192. Chung, A.S.; Lee, J.; Ferrara, N. Targeting the Tumour Vasculature: Insights from Physiological Angiogenesis. Nat. Rev. Cancer 2010, 10, 505–514. [Google Scholar] [CrossRef]
  193. Teleanu, R.I.; Chircov, C.; Grumezescu, A.M.; Teleanu, D.M. Tumor Angiogenesis and Anti-Angiogenic Strategies for Cancer Treatment. J. Clin. Med. 2019, 9, 84. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Deroanne, C.F.; Bonjean, K.; Servotte, S.; Devy, L.; Colige, A.; Clausse, N.; Blacher, S.; Verdin, E.; Foidart, J.-M.; Nusgens, B.V.; et al. Histone Deacetylases Inhibitors as Anti-Angiogenic Agents Altering Vascular Endothelial Growth Factor Signaling. Oncogene 2002, 21, 427–436. [Google Scholar] [CrossRef] [Green Version]
  195. Kim, M.S.; Kwon, H.J.; Lee, Y.M.; Baek, J.H.; Jang, J.E.; Lee, S.W.; Moon, E.J.; Kim, H.S.; Lee, S.K.; Chung, H.Y.; et al. Histone Deacetylases Induce Angiogenesis by Negative Regulation of Tumor Suppressor Genes. Nat. Med. 2001, 7, 437–443. [Google Scholar] [CrossRef]
  196. Qian, D.Z.; Kato, Y.; Shabbeer, S.; Wei, Y.; Verheul, H.M.W.; Salumbides, B.; Sanni, T.; Atadja, P.; Pili, R. Targeting Tumor Angiogenesis with Histone Deacetylase Inhibitors: The Hydroxamic Acid Derivative LBH589. Clin. Cancer Res. 2006, 12, 634–642. [Google Scholar] [CrossRef] [Green Version]
  197. Deng, B.; Luo, Q.; Halim, A.; Liu, Q.; Zhang, B.; Song, G. The Antiangiogenesis Role of Histone Deacetylase Inhibitors: Their Potential Application to Tumor Therapy and Tissue Repair. DNA Cell Biol. 2019, 39, 167–176. [Google Scholar] [CrossRef]
  198. Ellis, L.; Hammers, H.; Pili, R. Targeting Tumor Angiogenesis with Histone Deacetylase Inhibitors. Cancer Lett. 2009, 280, 145. [Google Scholar] [CrossRef] [Green Version]
  199. Qian, D.Z.; Kachhap, S.K.; Collis, S.J.; Verheul, H.M.W.; Carducci, M.A.; Atadja, P.; Pili, R. Class II Histone Deacetylases Are Associated with VHL-Independent Regulation of Hypoxia-Inducible Factor 1α. Cancer Res. 2006, 66, 8814–8821. [Google Scholar] [CrossRef] [Green Version]
  200. Zhang, Y.; Zhang, Z. The History and Advances in Cancer Immunotherapy: Understanding the Characteristics of Tumor-Infiltrating Immune Cells and Their Therapeutic Implications. Cell. Mol. Immunol. 2020, 17, 807–821. [Google Scholar] [CrossRef] [PubMed]
  201. Cao, K.; Wang, G.; Li, W.; Zhang, L.; Wang, R.; Huang, Y.; Du, L.; Jiang, J.; Wu, C.; He, X.; et al. Histone Deacetylase Inhibitors Prevent Activation-Induced Cell Death and Promote Anti-Tumor Immunity. Oncogene 2015, 34, 5960–5970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Christiansen, A.J.; West, A.; Banks, K.-M.; Haynes, N.M.; Teng, M.W.; Smyth, M.J.; Johnstone, R.W. Eradication of Solid Tumors Using Histone Deacetylase Inhibitors Combined with Immune-Stimulating Antibodies. Proc. Natl. Acad. Sci. USA 2011, 108, 4141–4146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Duvic, M.; Talpur, R.; Ni, X.; Zhang, C.; Hazarika, P.; Kelly, C.; Chiao, J.H.; Reilly, J.F.; Ricker, J.L.; Richon, V.M.; et al. Phase 2 Trial of Oral Vorinostat (Suberoylanilide Hydroxamic Acid, SAHA) for Refractory Cutaneous T-Cell Lymphoma (CTCL). Blood 2007, 109, 31–39. [Google Scholar] [CrossRef] [PubMed]
  204. Pili, R.; Liu, G.; Chintala, S.; Verheul, H.; Rehman, S.; Attwood, K.; Lodge, M.A.; Wahl, R.; Martin, J.I.; Miles, K.M.; et al. Combination of the Histone Deacetylase Inhibitor Vorinostat with Bevacizumab in Patients with Clear-Cell Renal Cell Carcinoma: A Multicentre, Single-Arm Phase I/II Clinical Trial. Br. J. Cancer 2017, 116, 874–883. [Google Scholar] [CrossRef] [Green Version]
  205. Heng, D.Y.C.; Xie, W.; Regan, M.M.; Harshman, L.C.; Bjarnason, G.A.; Vaishampayan, U.N.; Mackenzie, M.; Wood, L.; Donskov, F.; Tan, M.-H.; et al. External Validation and Comparison with Other Models of the International Metastatic Renal-Cell Carcinoma Database Consortium Prognostic Model: A Population-Based Study. Lancet Oncol. 2013, 14, 141–148. [Google Scholar] [CrossRef] [Green Version]
  206. Shen, C.; Kaelin, W.G. The VHL/HIF Axis in Clear Cell Renal Carcinoma. Semin. Cancer Biol. 2013, 23, 18–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Makhov, P.; Joshi, S.; Ghatalia, P.; Kutikov, A.; Uzzo, R.G.; Kolenko, V.M. Resistance to Systemic Therapies in Clear Cell Renal Cell Carcinoma: Mechanisms and Management Strategies. Mol. Cancer Ther. 2018, 17, 1355–1364. [Google Scholar] [CrossRef] [Green Version]
  208. Rini, B.I. Targeted Therapy for Patients with Renal-Cell Carcinoma. Lancet Oncol. 2011, 12, 1085–1087. [Google Scholar] [CrossRef]
  209. Jolly, M.K.; Celià-Terrassa, T. Dynamics of Phenotypic Heterogeneity Associated with EMT and Stemness during Cancer Progression. J. Clin. Med. 2019, 8, 1542. [Google Scholar] [CrossRef] [Green Version]
  210. Craene, B.D.; Berx, G. Regulatory Networks Defining EMT during Cancer Initiation and Progression. Nat. Rev. Cancer 2013, 13, 97–110. [Google Scholar] [CrossRef] [PubMed]
  211. Nieto, M.A.; Huang, R.Y.-J.; Jackson, R.A.; Thiery, J.P. EMT: 2016. Cell 2016, 166, 21–45. [Google Scholar] [CrossRef] [Green Version]
  212. Shibue, T.; Weinberg, R.A. EMT, CSCs, and Drug Resistance: The Mechanistic Link and Clinical Implications. Nat. Rev. Clin. Oncol. 2017, 14, 611–629. [Google Scholar] [CrossRef] [Green Version]
  213. Shah, P.; Gau, Y.; Sabnis, G. Histone Deacetylase Inhibitor Entinostat Reverses Epithelial to Mesenchymal Transition of Breast Cancer Cells by Reversing the Repression of E-Cadherin. Breast Cancer Res. Treat. 2014, 143, 99–111. [Google Scholar] [CrossRef]
  214. Su, Y.; Hopfinger, N.R.; Nguyen, T.D.; Pogash, T.J.; Santucci-Pereira, J.; Russo, J. Epigenetic Reprogramming of Epithelial Mesenchymal Transition in Triple Negative Breast Cancer Cells with DNA Methyltransferase and Histone Deacetylase Inhibitors. J. Exp. Clin. Cancer Res. 2018, 37, 314. [Google Scholar] [CrossRef] [PubMed]
  215. Tang, H.M.; Kuay, K.T.; Koh, P.F.; Asad, M.; Tan, T.Z.; Chung, V.Y.; Lee, S.C.; Thiery, J.P.; Huang, R.-J. An Epithelial Marker Promoter Induction Screen Identifies Histone Deacetylase Inhibitors to Restore Epithelial Differentiation and Abolishes Anchorage Independence Growth in Cancers. Cell Death Discov. 2016, 2, 1–10. [Google Scholar] [CrossRef] [PubMed]
  216. Hu, Y.; Nie, Q.; Dai, M.; Chen, F.; Wu, H. Histone Deacetylases Inhibit the Snail2-Mediated EMT During Metastasis of Hepatocellular Carcinoma Cells. Front. Cell Dev. Biol. 2020, 8, 752. [Google Scholar] [CrossRef] [PubMed]
  217. Rahimian, A.; Barati, G.; Mehrandish, R.; Mellati, A.A. Inhibition of Histone Deacetylases Reverses Epithelial-Mesenchymal Transition in Triple-Negative Breast Cancer Cells through a Slug Mediated Mechanism. Mol. Biol. 2018, 52, 406–413. [Google Scholar] [CrossRef]
  218. Xiao, Q.; Liu, H.; Wang, H.-S.; Cao, M.-T.; Meng, X.-J.; Xiang, Y.-L.; Zhang, Y.-Q.; Shu, F.; Zhang, Q.-G.; Shan, H.; et al. Histone Deacetylase Inhibitors Promote Epithelial-Mesenchymal Transition in Hepatocellular Carcinoma via AMPK-FOXO1-ULK1 Signaling Axis-Mediated Autophagy. Theranostics 2020, 10, 10245–10261. [Google Scholar] [CrossRef]
  219. Jiang, G.-M.; Wang, H.-S.; Zhang, F.; Zhang, K.-S.; Liu, Z.-C.; Fang, R.; Wang, H.; Cai, S.-H.; Du, J. Histone Deacetylase Inhibitor Induction of Epithelial–Mesenchymal Transitions via up-Regulation of Snail Facilitates Cancer Progression. Biochim. Biophys. Acta BBA Mol. Cell Res. 2013, 1833, 663–671. [Google Scholar] [CrossRef] [Green Version]
  220. Li, X.; Lewis, M.T.; Huang, J.; Gutierrez, C.; Osborne, C.K.; Wu, M.-F.; Hilsenbeck, S.G.; Pavlick, A.; Zhang, X.; Chamness, G.C.; et al. Intrinsic Resistance of Tumorigenic Breast Cancer Cells to Chemotherapy. J. Natl. Cancer Inst. 2008, 100, 672–679. [Google Scholar] [CrossRef]
  221. Yu, F.; Yao, H.; Zhu, P.; Zhang, X.; Pan, Q.; Gong, C.; Huang, Y.; Hu, X.; Su, F.; Lieberman, J.; et al. Let-7 Regulates Self Renewal and Tumorigenicity of Breast Cancer Cells. Cell 2007, 131, 1109–1123. [Google Scholar] [CrossRef] [Green Version]
  222. Creighton, C.J.; Li, X.; Landis, M.; Dixon, J.M.; Neumeister, V.M.; Sjolund, A.; Rimm, D.L.; Wong, H.; Rodriguez, A.; Herschkowitz, J.I.; et al. Residual Breast Cancers after Conventional Therapy Display Mesenchymal as Well as Tumor-Initiating Features. Proc. Natl. Acad. Sci. USA 2009, 106, 13820–13825. [Google Scholar] [CrossRef] [Green Version]
  223. Tehranchi, R.; Woll, P.S.; Anderson, K.; Buza-Vidas, N.; Mizukami, T.; Mead, A.J.; Åstrand-Grundström, I.; Strömbeck, B.; Horvat, A.; Ferry, H.; et al. Persistent Malignant Stem Cells in Del(5q) Myelodysplasia in Remission. N. Engl. J. Med. 2010, 363, 1025–1037. [Google Scholar] [CrossRef] [PubMed]
  224. Plaks, V.; Kong, N.; Werb, Z. The Cancer Stem Cell Niche: How Essential Is the Niche in Regulating Stemness of Tumor Cells? Cell Stem Cell 2015, 16, 225–238. [Google Scholar] [CrossRef] [Green Version]
  225. Carnero, A.; Garcia-Mayea, Y.; Mir, C.; Lorente, J.; Rubio, I.T.; LLeonart, M.E. The Cancer Stem-Cell Signaling Network and Resistance to Therapy. Cancer Treat. Rev. 2016, 49, 25–36. [Google Scholar] [CrossRef] [PubMed]
  226. Saygin, C.; Matei, D.; Majeti, R.; Reizes, O.; Lathia, J.D. Targeting Cancer Stemness in the Clinic: From Hype to Hope. Cell Stem Cell 2019, 24, 25–40. [Google Scholar] [CrossRef] [Green Version]
  227. Takebe, N.; Miele, L.; Harris, P.J.; Jeong, W.; Bando, H.; Kahn, M.; Yang, S.X.; Ivy, S.P. Targeting Notch, Hedgehog, and Wnt Pathways in Cancer Stem Cells: Clinical Update. Nat. Rev. Clin. Oncol. 2015, 12, 445–464. [Google Scholar] [CrossRef]
  228. Wainwright, E.N.; Scaffidi, P. Epigenetics and Cancer Stem Cells: Unleashing, Hijacking, and Restricting Cellular Plasticity. Trends Cancer 2017, 3, 372–386. [Google Scholar] [CrossRef] [Green Version]
  229. Dawson, M.A. The Cancer Epigenome: Concepts, Challenges, and Therapeutic Opportunities. Science 2017, 355, 1147–1152. [Google Scholar] [CrossRef]
  230. Baba, T.; Convery, P.A.; Matsumura, N.; Whitaker, R.S.; Kondoh, E.; Perry, T.; Huang, Z.; Bentley, R.C.; Mori, S.; Fujii, S.; et al. Epigenetic Regulation of CD133 and Tumorigenicity of CD133+ Ovarian Cancer Cells. Oncogene 2009, 28, 209–218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  231. Yi, J.M.; Tsai, H.-C.; Glöckner, S.C.; Lin, S.; Ohm, J.E.; Easwaran, H.; James, C.D.; Costello, J.F.; Riggins, G.; Eberhart, C.G.; et al. Abnormal DNA Methylation of CD133 in Colorectal and Glioblastoma Tumors. Cancer Res. 2008, 68, 8094–8103. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Lin, B.; Lee, H.; Yoon, J.-G.; Madan, A.; Wayner, E.; Tonning, S.; Hothi, P.; Schroeder, B.; Ulasov, I.; Foltz, G.; et al. Global Analysis of H3K4me3 and H3K27me3 Profiles in Glioblastoma Stem Cells and Identification of SLC17A7 as a Bivalent Tumor Suppressor Gene. Oncotarget 2015, 6, 5369–5381. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Marampon, F.; Megiorni, F.; Camero, S.; Crescioli, C.; McDowell, H.P.; Sferra, R.; Vetuschi, A.; Pompili, S.; Ventura, L.; De Felice, F.; et al. HDAC4 and HDAC6 Sustain DNA Double Strand Break Repair and Stem-like Phenotype by Promoting Radioresistance in Glioblastoma Cells. Cancer Lett. 2017, 397, 1–11. [Google Scholar] [CrossRef]
  234. Yang, W.; Liu, Y.; Gao, R.; Yu, H.; Sun, T. HDAC6 Inhibition Induces Glioma Stem Cells Differentiation and Enhances Cellular Radiation Sensitivity through the SHH/Gli1 Signaling Pathway. Cancer Lett. 2018, 415, 164–176. [Google Scholar] [CrossRef] [PubMed]
  235. Witt, A.E.; Lee, C.-W.; Lee, T.I.; Azzam, D.J.; Wang, B.; Caslini, C.; Petrocca, F.; Grosso, J.; Jones, M.; Cohick, E.B.; et al. Identification of a Cancer Stem Cell-Specific Function for the Histone Deacetylases, HDAC1 and HDAC7, in Breast and Ovarian Cancer. Oncogene 2017, 36, 1707–1720. [Google Scholar] [CrossRef] [PubMed]
  236. Bora-Singhal, N.; Mohankumar, D.; Saha, B.; Colin, C.M.; Lee, J.Y.; Martin, M.W.; Zheng, X.; Coppola, D.; Chellappan, S. Novel HDAC11 Inhibitors Suppress Lung Adenocarcinoma Stem Cell Self-Renewal and Overcome Drug Resistance by Suppressing Sox2. Sci. Rep. 2020, 10, 4722. [Google Scholar] [CrossRef]
  237. Yamazaki, J.; Estecio, M.R.; Lu, Y.; Long, H.; Malouf, G.G.; Graber, D.; Huo, Y.; Ramagli, L.; Liang, S.; Kornblau, S.M.; et al. The Epigenome of AML Stem and Progenitor Cells. Epigenetics 2013, 8, 92–104. [Google Scholar] [CrossRef] [Green Version]
  238. Easwaran, H.; Tsai, H.-C.; Baylin, S.B. Cancer Epigenetics: Tumor Heterogeneity, Plasticity of Stem-like States, and Drug Resistance. Mol. Cell 2014, 54, 716–727. [Google Scholar] [CrossRef] [Green Version]
  239. Hennessy, B.T.; Gonzalez-Angulo, A.-M.; Stemke-Hale, K.; Gilcrease, M.Z.; Krishnamurthy, S.; Lee, J.-S.; Fridlyand, J.; Sahin, A.; Agarwal, R.; Joy, C.; et al. Characterization of a Naturally Occurring Breast Cancer Subset Enriched in Epithelial-to-Mesenchymal Transition and Stem Cell Characteristics. Cancer Res. 2009, 69, 4116–4124. [Google Scholar] [CrossRef] [Green Version]
  240. Wellner, U.; Schubert, J.; Burk, U.C.; Schmalhofer, O.; Zhu, F.; Sonntag, A.; Waldvogel, B.; Vannier, C.; Darling, D.; zur Hausen, A.; et al. The EMT-Activator ZEB1 Promotes Tumorigenicity by Repressing Stemness-Inhibiting MicroRNAs. Nat. Cell Biol. 2009, 11, 1487–1495. [Google Scholar] [CrossRef] [PubMed]
  241. Keyvani-Ghamsari, S.; Khorsandi, K.; Rasul, A.; Zaman, M.K. Current Understanding of Epigenetics Mechanism as a Novel Target in Reducing Cancer Stem Cells Resistance. Clin. Epigenet. 2021, 13, 120. [Google Scholar] [CrossRef]
  242. Schech, A.; Kazi, A.; Yu, S.; Shah, P.; Sabnis, G. Histone Deacetylase Inhibitor Entinostat Inhibits Tumor-Initiating Cells in Triple-Negative Breast Cancer Cells. Mol. Cancer Ther. 2015, 14, 1848–1857. [Google Scholar] [CrossRef] [Green Version]
  243. Hsu, C.-C.; Chang, W.-C.; Hsu, T.-I.; Liu, J.-J.; Yeh, S.-H.; Wang, J.-Y.; Liou, J.-P.; Ko, C.-Y.; Chang, K.-Y.; Chuang, J.-Y. Suberoylanilide Hydroxamic Acid Represses Glioma Stem-like Cells. J. Biomed. Sci. 2016, 23, 81. [Google Scholar] [CrossRef] [Green Version]
  244. Pastorino, O.; Gentile, M.T.; Mancini, A.; Del Gaudio, N.; Di Costanzo, A.; Bajetto, A.; Franco, P.; Altucci, L.; Florio, T.; Stoppelli, M.P.; et al. Histone Deacetylase Inhibitors Impair Vasculogenic Mimicry from Glioblastoma Cells. Cancers 2019, 11, 747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Kawaguchi, Y.; Kovacs, J.J.; McLaurin, A.; Vance, J.M.; Ito, A.; Yao, T.P. The Deacetylase HDAC6 Regulates Aggresome Formation and Cell Viability in Response to Misfolded Protein Stress. Cell 2003, 115, 727–738. [Google Scholar] [CrossRef] [Green Version]
  246. Kovacs, J.J.; Murphy, P.J.M.; Gaillard, S.; Zhao, X.; Wu, J.-T.; Nicchitta, C.V.; Yoshida, M.; Toft, D.O.; Pratt, W.B.; Yao, T.-P. HDAC6 Regulates Hsp90 Acetylation and Chaperone-Dependent Activation of Glucocorticoid Receptor. Mol. Cell 2005, 18, 601–607. [Google Scholar] [CrossRef] [PubMed]
  247. Krämer, O.H.; Mahboobi, S.; Sellmer, A. Drugging the HDAC6-HSP90 Interplay in Malignant Cells. Trends Pharmacol. Sci. 2014, 35, 501–509. [Google Scholar] [CrossRef] [PubMed]
  248. Ryu, H.-W.; Shin, D.-H.; Lee, D.H.; Choi, J.; Han, G.; Lee, K.Y.; Kwon, S.H. HDAC6 Deacetylates P53 at Lysines 381/382 and Differentially Coordinates P53-Induced Apoptosis. Cancer Lett. 2017, 391, 162–171. [Google Scholar] [CrossRef]
  249. Tien, S.-C.; Chang, Z.-F. Oncogenic Shp2 Disturbs Microtubule Regulation to Cause HDAC6-Dependent ERK Hyperactivation. Oncogene 2014, 33, 2938–2946. [Google Scholar] [CrossRef]
  250. Cosenza, M.; Pozzi, S. The Therapeutic Strategy of HDAC6 Inhibitors in Lymphoproliferative Disease. Int. J. Mol. Sci. 2018, 19, 2337. [Google Scholar] [CrossRef] [Green Version]
  251. Hai, Y.; Christianson, D.W. Histone Deacetylase 6 Structure and Molecular Basis of Catalysis and Inhibition. Nat. Chem. Biol. 2016, 12, 741–747. [Google Scholar] [CrossRef] [Green Version]
  252. Gawel, J.M.; Shouksmith, A.E.; Raouf, Y.S.; Nawar, N.; Toutah, K.; Bukhari, S.; Manaswiyoungkul, P.; Olaoye, O.O.; Israelian, J.; Radu, T.B.; et al. PTG-0861: A Novel HDAC6-Selective Inhibitor as a Therapeutic Strategy in Acute Myeloid Leukaemia. Eur. J. Med. Chem. 2020, 201, 112411. [Google Scholar] [CrossRef]
  253. Santo, L.; Hideshima, T.; Kung, A.L.; Tseng, J.-C.; Tamang, D.; Yang, M.; Jarpe, M.; van Duzer, J.H.; Mazitschek, R.; Ogier, W.C.; et al. Preclinical Activity, Pharmacodynamic, and Pharmacokinetic Properties of a Selective HDAC6 Inhibitor, ACY-1215, in Combination with Bortezomib in Multiple Myeloma. Blood 2012, 119, 2579–2589. [Google Scholar] [CrossRef]
  254. Amengual, J.E.; Johannet, P.; Lombardo, M.; Zullo, K.; Hoehn, D.; Bhagat, G.; Scotto, L.; Jirau-Serrano, X.; Radeski, D.; Heinen, J.; et al. Dual Targeting of Protein Degradation Pathways with the Selective HDAC6 Inhibitor ACY-1215 and Bortezomib Is Synergistic in Lymphoma. Clin. Cancer Res. 2015, 21, 4663–4675. [Google Scholar] [CrossRef] [Green Version]
  255. Yee, A.J.; Bensinger, W.I.; Supko, J.G.; Voorhees, P.M.; Berdeja, J.G.; Richardson, P.G.; Libby, E.N.; Wallace, E.E.; Birrer, N.E.; Burke, J.N.; et al. Ricolinostat plus Lenalidomide, and Dexamethasone in Relapsed or Refractory Multiple Myeloma: A Multicentre Phase 1b Trial. Lancet Oncol. 2016, 17, 1569–1578. [Google Scholar] [CrossRef]
  256. Lin, A.; Giuliano, C.J.; Palladino, A.; John, K.M.; Abramowicz, C.; Yuan, M.L.; Sausville, E.L.; Lukow, D.A.; Liu, L.; Chait, A.R.; et al. Off-Target Toxicity Is a Common Mechanism of Action of Cancer Drugs Undergoing Clinical Trials. Sci. Transl. Med. 2019, 11. [Google Scholar] [CrossRef]
  257. Huang, P.; Almeciga-Pinto, I.; Jarpe, M.; van Duzer, J.H.; Mazitschek, R.; Yang, M.; Jones, S.S.; Quayle, S.N. Selective HDAC Inhibition by ACY-241 Enhances the Activity of Paclitaxel in Solid Tumor Models. Oncotarget 2016, 8, 2694–2707. [Google Scholar] [CrossRef] [Green Version]
  258. Schäfer, C.; Göder, A.; Beyer, M.; Kiweler, N.; Mahendrarajah, N.; Rauch, A.; Nikolova, T.; Stojanovic, N.; Wieczorek, M.; Reich, T.R.; et al. Class I Histone Deacetylases Regulate P53/NF-ΚB Crosstalk in Cancer Cells. Cell Signal. 2017, 29, 218–225. [Google Scholar] [CrossRef] [PubMed]
  259. Tiffon, C.E.; Adams, J.E.; van der Fits, L.; Wen, S.; Townsend, P.A.; Ganesan, A.; Hodges, E.; Vermeer, M.H.; Packham, G. The Histone Deacetylase Inhibitors Vorinostat and Romidepsin Downmodulate IL-10 Expression in Cutaneous T-Cell Lymphoma Cells. Br. J. Pharmacol. 2011, 162, 1590–1602. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Cortiguera, M.G.; García-Gaipo, L.; Wagner, S.D.; León, J.; Batlle-López, A.; Delgado, M.D. Suppression of BCL6 Function by HDAC Inhibitor Mediated Acetylation and Chromatin Modification Enhances BET Inhibitor Effects in B-Cell Lymphoma Cells. Sci. Rep. 2019, 9, 16495. [Google Scholar] [CrossRef] [PubMed]
  261. Hideshima, T.; Bradner, J.E.; Wong, J.; Chauhan, D.; Richardson, P.; Schreiber, S.L.; Anderson, K.C. Small-Molecule Inhibition of Proteasome and Aggresome Function Induces Synergistic Antitumor Activity in Multiple Myeloma. Proc. Natl. Acad. Sci. USA 2005, 102, 8567–8572. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Simms-Waldrip, T.; Rodriguez-Gonzalez, A.; Lin, T.; Ikeda, A.K.; Fu, C.; Sakamoto, K.M. Targeting the Aggresome Pathway in Hematologic Malignancies. Mol. Genet. Metab. 2008, 94, 283–286. [Google Scholar] [CrossRef] [Green Version]
  263. Chun, P. Histone Deacetylase Inhibitors in Hematological Malignancies and Solid Tumors. Arch. Pharm. Res. 2015, 38, 933–949. [Google Scholar] [CrossRef] [PubMed]
  264. Sanchez, G.J.; Richmond, P.A.; Bunker, E.N.; Karman, S.S.; Azofeifa, J.; Garnett, A.T.; Xu, Q.; Wheeler, G.E.; Toomey, C.M.; Zhang, Q.; et al. Genome-Wide Dose-Dependent Inhibition of Histone Deacetylases Studies Reveal Their Roles in Enhancer Remodeling and Suppression of Oncogenic Super-Enhancers. Nucleic Acids Res. 2018, 46, 1756–1776. [Google Scholar] [CrossRef]
  265. Otterson, G.A.; Hodgson, L.; Pang, H.; Vokes, E.E. Phase II Study of the Histone Deacetylase Inhibitor Romidepsin in Relapsed Small Cell Lung Cancer (Cancer and Leukemia Group B 30304). J. Thorac. Oncol. 2010, 5, 1644–1648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Yardley, D.A.; Ismail-Khan, R.R.; Melichar, B.; Lichinitser, M.; Munster, P.N.; Klein, P.M.; Cruickshank, S.; Miller, K.D.; Lee, M.J.; Trepel, J.B. Randomized Phase II, Double-Blind, Placebo-Controlled Study of Exemestane with or without Entinostat in Postmenopausal Women with Locally Recurrent or Metastatic Estrogen Receptor-Positive Breast Cancer Progressing on Treatment with a Nonsteroidal Aromatase Inhibitor. J. Clin. Oncol. 2013, 31, 2128–2135. [Google Scholar] [CrossRef] [Green Version]
  267. Richardson, P.G.; Hungria, V.T.M.; Yoon, S.-S.; Beksac, M.; Dimopoulos, M.A.; Elghandour, A.; Jedrzejczak, W.W.; Guenther, A.; Na Nakorn, T.; Siritanaratkul, N.; et al. Panorama 1: A Randomized, Double-Blind, Phase 3 Study of Panobinostat or Placebo plus Bortezomib and Dexamethasone in Relapsed or Relapsed and Refractory Multiple Myeloma. JCO 2014, 32, 8510. [Google Scholar] [CrossRef]
  268. Zhang, M.-C.; Fang, Y.; Wang, L.; Cheng, S.; Fu, D.; He, Y.; Zhao, Y.; Wang, C.-F.; Jiang, X.-F.; Song, Q.; et al. Clinical Efficacy and Molecular Biomarkers in a Phase II Study of Tucidinostat plus R-CHOP in Elderly Patients with Newly Diagnosed Diffuse Large B-Cell Lymphoma. Clin. Epigenet. 2020, 12, 160. [Google Scholar] [CrossRef]
Figure 1. A schematic depicting how HDAC inhibitors modulate histone acetylation in relation to chromatin structure and acetylation of non-histone proteins. HAT: Histone Acetyltransferase; HDAC: Histone deacetylase; HDACi: Histone deacetylase inhibitors.
Figure 1. A schematic depicting how HDAC inhibitors modulate histone acetylation in relation to chromatin structure and acetylation of non-histone proteins. HAT: Histone Acetyltransferase; HDAC: Histone deacetylase; HDACi: Histone deacetylase inhibitors.
Cancers 13 03575 g001
Figure 2. HDAC inhibitor coordination to the active site metal Zn2+ ion. (A) HDAC8 structures showing the binding sites of substrate (left, PDB: 2V5W), SAHA (center, PDB: 1T69), and trichostatin A (right, PDB: 1T64) (colored small molecules). (B) Stereoviews showing substrate carbonyl (left) and HDAC inhibitors (center and right) coordination to the catalytic Zn2+. (C) Zinc binding group and cap group in SAHA and trichostatin A inhibitors. TSA: Trichostatin A; SAHA: Suberoylanilide Hydroxamic Acid.
Figure 2. HDAC inhibitor coordination to the active site metal Zn2+ ion. (A) HDAC8 structures showing the binding sites of substrate (left, PDB: 2V5W), SAHA (center, PDB: 1T69), and trichostatin A (right, PDB: 1T64) (colored small molecules). (B) Stereoviews showing substrate carbonyl (left) and HDAC inhibitors (center and right) coordination to the catalytic Zn2+. (C) Zinc binding group and cap group in SAHA and trichostatin A inhibitors. TSA: Trichostatin A; SAHA: Suberoylanilide Hydroxamic Acid.
Cancers 13 03575 g002
Figure 3. A schematic illustrating how HDAC inhibitors could be utilized to reduce heterogeneity in cancer. During cancer progression, subclone selection pressures from the microenvironment, intrinsic cellular properties and selection pressures during treatment generate tumor heterogeneity. The effects of HDACi can be utilized rationally to counter these sources of heterogeneity. Through this approach, the tumor becomes more homogeneous in certain aspects of tumor cell biology and the microenvironment, and thus responds better to cytotoxic and/or targeted therapy. EMT: Epithelial-Mesenchymal Transition; ROS: Reactive Oxygen Species; CSCs: Cancer Stem Cells; ECM: Extracellular Matrix.
Figure 3. A schematic illustrating how HDAC inhibitors could be utilized to reduce heterogeneity in cancer. During cancer progression, subclone selection pressures from the microenvironment, intrinsic cellular properties and selection pressures during treatment generate tumor heterogeneity. The effects of HDACi can be utilized rationally to counter these sources of heterogeneity. Through this approach, the tumor becomes more homogeneous in certain aspects of tumor cell biology and the microenvironment, and thus responds better to cytotoxic and/or targeted therapy. EMT: Epithelial-Mesenchymal Transition; ROS: Reactive Oxygen Species; CSCs: Cancer Stem Cells; ECM: Extracellular Matrix.
Cancers 13 03575 g003
Figure 4. Deacetylation of p53 (e.g., at K373/K382) and Ku70 by histone deacetylases (HDACs) regulates tumor cell proliferation and DNA damage repair. DSB: double-strand break; HAT: histone acetyltransferase; c-NHEJ: canonical non-homologous end-joining.
Figure 4. Deacetylation of p53 (e.g., at K373/K382) and Ku70 by histone deacetylases (HDACs) regulates tumor cell proliferation and DNA damage repair. DSB: double-strand break; HAT: histone acetyltransferase; c-NHEJ: canonical non-homologous end-joining.
Cancers 13 03575 g004
Figure 5. A perspective on how prominent effects of HDAC inhibitors take place in a time and dose-dependent manner based on the current literature.
Figure 5. A perspective on how prominent effects of HDAC inhibitors take place in a time and dose-dependent manner based on the current literature.
Cancers 13 03575 g005
Table 1. Classification, targeting selectivity (HDAC class), and applications in cancer therapy for the most-studied HDAC inhibitors.
Table 1. Classification, targeting selectivity (HDAC class), and applications in cancer therapy for the most-studied HDAC inhibitors.
ClassificationChemical NameTargeted HDACsFDA Approval
Cyclic depsipeptideRomidepsinClass I [10,11]CTCL, PTCL [3,4]
LargazoleClass I [11]Under investigation
Hydroxamic acidTrichostatin A(TSA)Class I/II/IV [12,13]Under investigation
Vorinostat/SAHACTCL [1]
BelinostatPTCL [2]
PanobinostatMM [14]
BenzamideEntinostatClass I [12,15]Under investigation
Chidamide/TucidinostatClass I [16]PTCL (China)
Carboxylic acidValproic acidClass I/IIa [17,18]Under investigation
Butyric acidClass I/IIUnder investigation
CTCL: Cutaneous T-cell Lymphoma; PTCL: Peripheral T-cell Lymphoma; TSA: Trichostatin A; SAHA: Suberoylanilide Hydroxamic Acid; MM: Multiple Myeloma.
Table 2. Examples of how HDAC inhibitors can be utilized to mitigate specific resistance mechanisms in several types of treatment and cancer settings.
Table 2. Examples of how HDAC inhibitors can be utilized to mitigate specific resistance mechanisms in several types of treatment and cancer settings.
TreatmentCancer TypeResistance Mechanism That Can Be Suppressed by HDACi
ChemotherapySolid tumor without targeted therapy optionClonal transcriptional heterogeneity
Glycolysis induction [74]
PARP inhibitionHRR-deficient cancerHRR activation [58]
Checkpoint blockade inhibitionLung, Bladder and moreImmune surveillance evasion [75]
Tyrosine kinase inhibitorEGFR+ Lung cancerMET overexpression; EMT [55]
Anti-estrogensER+ Breast cancerTranscriptional Heterogeneity [71]
HRR: homologous recombination repair; EGFR: epidermal growth factor receptor; EMT: epithelial–mesenchymal transition; ER: estrogen receptor.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Karagiannis, D.; Rampias, T. HDAC Inhibitors: Dissecting Mechanisms of Action to Counter Tumor Heterogeneity. Cancers 2021, 13, 3575. https://doi.org/10.3390/cancers13143575

AMA Style

Karagiannis D, Rampias T. HDAC Inhibitors: Dissecting Mechanisms of Action to Counter Tumor Heterogeneity. Cancers. 2021; 13(14):3575. https://doi.org/10.3390/cancers13143575

Chicago/Turabian Style

Karagiannis, Dimitris, and Theodoros Rampias. 2021. "HDAC Inhibitors: Dissecting Mechanisms of Action to Counter Tumor Heterogeneity" Cancers 13, no. 14: 3575. https://doi.org/10.3390/cancers13143575

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

Article Metrics

Back to TopTop