Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Jan 1.
Published in final edited form as: Med Res Rev. 2019 Jun 20;40(1):190–244. doi: 10.1002/med.21600

Epigenetic Polypharmacology: A New Frontier for Epi-Drug Discovery

Daniela Tomaselli 1, Alessia Lucidi 1, Dante Rotili 1,*, Antonello Mai 1,2,*
PMCID: PMC6917854  NIHMSID: NIHMS1031232  PMID: 31218726

Abstract

Recently, despite the great success achieved by the so-called “magic bullets” in the treatment of different diseases through a marked and specific interaction with the target of interest, the pharmacological research is moving toward the development of “molecular network active compounds”, embracing the related polypharmacology approach. This strategy was born to overcome the main limitations of the single target therapy leading to a superior therapeutic effect, a decrease of adverse reactions, and a reduction of potential mechanism(s) of drug resistance caused by robustness and redundancy of biological pathways. It has become clear that multifactorial diseases such as cancer, neurological and inflammatory disorders, may require more complex therapeutic approaches hitting a certain biological system as a whole. Concerning epigenetics, the goal of the multi-epi-target approach consists in the development of small molecules able to simultaneously and (often) reversibly bind different specific epi-targets. To date, two dual HDAC/kinase inhibitors (CUDC-101 and CUDC-907) are in advanced stage of clinical trials. In the last years, the growing interest in polypharmacology encouraged the publication of high quality reviews on combination therapy and hybrid molecules. Hence, in order to update the state-of-art of these therapeutic approaches avoiding redundancy, herein we focused only on multiple medication therapies and multi-targeting compounds exploiting epigenetic plus non-epigenetic drugs reported in literature in 2018. In addition, all the multi-epi-target inhibitors known in literature so far, hitting two or more epigenetic targets, have been included.

Keywords: Polypharmacology, epigenetics, hybrid molecules design, cancer

1. INTRODUCTION

1.1. The polypharmacology strategy

Epigenetics (from the Greek words επί, over, and γεννετικός, genetics) studies the heritable and reversible changes in gene expression that cannot be explained by changes in the DNA sequence of bases. These alterations are governed by at least three main mechanisms: covalent modifications to the cytosine residues of DNA, histone covalent chemical modifications such as acetylation, methylation, phosphorylation, etc., also known on the whole as chromatin remodeling, and noncoding RNAs. Proteins taking part in chromatin remodeling complexes can add, remove, or read such epigenetic modifications, and are classified as “writers”, “erasers” or “readers”, respectively, determining either transcription or repression of specific genes. Since dysfunctional gene regulation is responsible for many human diseases, the modulation of epigenetic processes is presently considered an innovative and very interesting therapeutic strategy.1 In the last few years, the research in drug discovery shifted its attention from the “one drug, one target” to the “network active compounds” approach, which is the cornerstone of the modern polypharmacology.2 In case of failure of the so-called “magic bullets”, three different and alternative strategies may be followed: the multiple-medication therapy (MMT, drug combination), the multi-compound medication (MCM, which consists in the association of multiple active principles in the same pharmaceutical formulation) or the new, promising multi-target-directed ligands (MTDLs, using single multi-targeting compounds).3 The polypharmacology approach was born with the aim to overcome the limitations of the single-targeted therapy, such as the potential mechanism(s) of resistance caused by redundancy and robustness of biological pathways.4 Polypharmacology has the aim to simultaneously hit different targets, all related to the onset and development of a certain disease, although at different levels.5 Notably, polypharmacology is quite different from “compound promiscuity”. In fact, the so-called promiscuous (or “dirty”) drugs are typically characterized by a broad spectrum of biological activities directed against targets often not related to the specific disorder of interest, thus leading to a plethora of adverse reactions. Drug promiscuity is often responsible for the occurrence of drug toxicity at the pre-clinical stage.

During the last years, several research groups questioned the possibility to identify structural and not structural features emerging from statistically significant correlation studies (such as lipophilicity, positive or negative charges, basic centers, binding site similarity, flexibility, etc.) common to these promiscuous molecules, in order to identify them as early as possible during the drug discovery process, thus reducing the risk of failure.4,6,7 Just to give an example, the clinical application of the anti-allergy drug astemizole8 was terminated because, in addition to the desired antagonism of H1 receptors, it inhibits hERG heart potassium channels leading to serious arrhythmia events. For this reason, the dirty drug astemizole was withdrawn from the market.9 Despite “promiscuity” is related to negative connotation (off-target interactions), it can be potentially useful in the hands of a medicinal chemist. Medicinal chemists could try to readdress the drug’s unwanted widespread multi-target effects into a specific and desired activity, potentially discovering a new target and allowing drug repurposing.7,10 This is the case, for example, of the antifungal agent itraconazole, which exerts its pharmacological effects through inhibition of CYP350-mediated ergosterol synthesis. Follow-up studies showed that itraconazole is also able to inhibit Hedgehog signal transduction and angiogenesis, opening the way for its use as anticancer agent. High doses of itraconazole showed moderate antitumor activity in a phase II clinical trial based on patients affected by castration-resistant prostate cancer ().9,11 In cancer, it has become clear that genetic, epigenetic and metabolic factors all contribute to neoplasia, involving significant changes in molecular networks at the base of many cell functions such as growth, differentiation, development, and death. This is the reason why multi-factorial diseases like cancer may require more complex therapeutic approaches, able to simultaneously interact with various signaling pathways and various cross-talks existing between epigenetic and not epigenetic players. Thus, an effective anticancer therapy should not consider a single pharmacological target, but the biological system as a whole.4 Some of the most important challenges of the polypharmacology strategy are the identification of the most appropriate targets and of the relative hit compound(s), as well as the optimization of its (their) structure-activity relationship.

1.2. An overview on single epigenetic targets and related inhibitors in cancer

Recently, a lot of small molecules characterized by high affinity and reversible binding to a specific epigenetic target have been discovered, resulting in seven approved anticancer agents and a number of clinical candidates which are shown in Table 1.

Table 1.

Clinically approved and clinical candidates among epigenetic modulators.

Compound Structure Epi-Target Clinical implication Status Clinical trial number(s)
1
Azacytidine
graphic file with name nihms-1031232-t0014.jpg DNMT acute myeloid leukemia, myelodysplastic syndrome FDA-approved,2004
2
Decitabine
graphic file with name nihms-1031232-t0015.jpg DNMT acute myeloid leukemia, myelodysplastic syndrome FDA-approved, 2006
3
Vorinostat
graphic file with name nihms-1031232-t0016.jpg HDAC cutaneous T-cell lymphoma FDA-approved, 2006
4
Romidepsin
graphic file with name nihms-1031232-t0017.jpg HDAC cutaneous T-cell lymphoma FDA-approved, 2009
5
Belinostat
graphic file with name nihms-1031232-t0018.jpg HDAC peripheral T-cell lymphoma FDA-approved, 2014
6
Panobinostat
graphic file with name nihms-1031232-t0019.jpg HDAC multiple myeloma FDA-approved, 2015
7
Tucidinostat (Chidamide)
graphic file with name nihms-1031232-t0020.jpg HDAC peripheral T-cell lymphoma Chinese FDA-approved, 2015
8
Sodium Valproate
graphic file with name nihms-1031232-t0021.jpg HDAC progressive, non-metastatic prostate cancer breast cancer solid tumors Phase I/II

9
Entinostat
graphic file with name nihms-1031232-t0022.jpg HDAC breast cancer invasive breast cancer ER-negative, PR-negative, HER2-negative (triple negative) breast cancer Phase I/II
10
Mocetinostat
graphic file with name nihms-1031232-t0023.jpg HDAC solid tumors, Hodgkin’s and non-Hodgkin’s lymphoma, leukemia Phase I/II


11
Abexinostat
graphic file with name nihms-1031232-t0024.jpg HDAC Hodgkin’s and non-Hodgkin’s lymphoma, multiple myeloma, leukemia, lymphocytic B-cell lymphoma Phase I/II
12
Pracinostat
graphic file with name nihms-1031232-t0025.jpg HDAC solid tumors, hematologic malignancies, myelodysplastic syndrome Phase I orphan drug for AML
13
Quisinostat
graphic file with name nihms-1031232-t0026.jpg HDAC cutaneous T-cell lymphoma Phase I/II
14
Resminostat
graphic file with name nihms-1031232-t0027.jpg HDAC advanced colorectal carcinoma Phase I/II orphan drug for hepatocellular carcinoma
15
Givinostat
graphic file with name nihms-1031232-t0028.jpg HDAC myeloproliferative diseases, Hodgkin’s lymphoma Phase I/II Orphan drug for Duchenne muscular dystrophy
16
Rocilinostat
graphic file with name nihms-1031232-t0029.jpg HDAC6 lymphoma, lymphoid malignancies, multiple myeloma Phase I/II
17
Nicotinamide
graphic file with name nihms-1031232-t0030.jpg Sirtuins solid tumors Phase II/III
18
GSK2816126
graphic file with name nihms-1031232-t0031.jpg EZH2 acute myeloid leukemia, non-Hodgkin lymphoma, multiple myeloma Phase I/II
19
Tazemetostat
graphic file with name nihms-1031232-t0032.jpg EZH2 different kinds of lymphomas and solid tumors Phase I/II

20
CPI-1205
graphic file with name nihms-1031232-t0033.jpg EZH2 B-cell lymphoma Phase I
21
Pinometostat
graphic file with name nihms-1031232-t0034.jpg DOT1L mixed-lineage leukemia Phase I *
22
JNJ-64619178
graphic file with name nihms-1031232-t0035.jpg PRMT5 relapsed/refractory B cell non-Hodgkin lymphoma, advanced solid tumors Phase I
23
GSK3326595
graphic file with name nihms-1031232-t0036.jpg PRMT5 selected solid tumors and non-Hodgkin’s lymphomas, neoplasms Phase I
24
Tranylcypromine
graphic file with name nihms-1031232-t0037.jpg LSD1 acute myelogenous leukemia Phase I/II
25
ORY-1001
graphic file with name nihms-1031232-t0038.jpg LSD1 myeloid leukemia, small cell lung cancer, relapsed or refractory acute leukemia Phase I/II UDRACT n°
2013-002447-29
2018-000482-36
2018-000469-35
26
GSK2879552
graphic file with name nihms-1031232-t0039.jpg LSD1 myelodysplastic syndrome leukemia small cell lung carcinoma Phase II/I

27
CPI-0610
graphic file with name nihms-1031232-t0040.jpg BRD family multiple myeloma, lymphoma Phase I
28
GSK525762
graphic file with name nihms-1031232-t0041.jpg BRD family relapsed refractory hematological malignancies, midline carcinoma Phase I
29
OTX015
graphic file with name nihms-1031232-t0042.jpg BRD family glioblastoma multiforme, triple negative breast cancer and other solid tumors, acute leukemia Phase II/I
*

Only in combination with azacytidine 1 or cytarabine, daunorubicin hydrochloride and pinometostat 21.

**

Only in combination with tretinoin or all-trans retinoic acid and cytarabine.

The first epigenetic modulators reaching the FDA approval for the treatment of hematological malignancies were the two DNA methyltransferase (DNMT) inhibitors (DNMTi), azacytidine 1 and decitabine 2, in 2004 and 2006, respectively.

DNMT is a class of enzymes able to catalyze the transfer of a methyl unit from S-adenosylmethionine (SAM), the methyl donor co-substrate, to the C5 of cytosine residues at CpG dinucleotide sequences within DNA. 5-Methylcytosines recruit proteins containing methyl binding domains (MBDs), leading to a global silencing of gene expression, especially due to the contribution of histone deacetylases (HDACs) and/or some histone methyltransferases (HMTs), two families of chromatin remodeling enzymes (see below). These enzymes are able to remove acetyl groups (HDACs) or to add methyl units (HMTs) to histone tails, allowing the recognition and binding of chromatin silencers, and leading to a transcriptionally inactive state of chromatin (heterochromatin).6 While in physiological conditions the cytosine methylation is concentrated in non-coding regions of DNA, in cancer cells features an increased methylation at the promoter regions of genes encoding for tumor suppressor factors or other proteins with anti-proliferative effects, joined to a global DNA hypomethylation. The potential utility of DNMTi in the anticancer therapy consists in their ability to reactivate silenced tumor suppressor genes and pathways.7,8 In humans, two kind of DNMTs are known, DNMT1 and DNMT3, while DNMT2 has been recognized as a RNA methyltransferase.12 DNMT1 is the most abundant, and is able to catalyze the methylation of hemi-methylated DNA maintaining the methylation status during cell replication. DNMT3 (including three isoforms: DNMT3A, −3B and −3L, the last catalytically inactive) is defined as “de novo” methyltransferase, acting on both hemi-methylated and non-methylated DNA.13,14

Azacytidine 1 and decitabine 2 (Table 1), characterized by a 1,3,5-triazine ring, are two approved drugs for the treatment of myelodysplastic syndrome (MDS), that often progresses to acute myeloid leukemia (AML). Through a triple phosphorylation, these pro-drugs are metabolically transformed into the active species which are incorporated into DNA and accepted as DNMT substrates. Nevertheless, after the nucleophilic attack of the enzyme at the triazine C6 position and the introduction of the methyl group at N5 from the SAM co-substrate, they do not allow the final elimination step as a result of the lack of a proton at the N5 position, leaving the enzyme irreversibly and covalently inhibited.15 However, despite their high efficacy, such drugs suffer from low selectivity, poor bioavailability, chemical instability and toxic side-effects. Non-nucleoside DNMTi are objects of great interest for the researchers since they do not need to be incorporated into DNA, overcoming the problem of toxicity linked to the use of nucleoside analogues. Unfortunately, many of them display low potency, scarce target selectivity, and unknown mechanisms of inhibition.

Among histone modifications, (de)acetylation and (de)methylation of lysine residues are the most studied. The acetylation reaction is catalyzed by histone acetyltransferases (HATs) using acetyl-CoA as co-substrate. Lysine acetylation decreases the affinity of histones for DNA, allowing the passage from the most compact heterochromatin to the more relaxed euchromatin form. This allows the recruitment of transcription factors, chromatin remodeling complexes and, finally, gene transcription. HDAC enzymes, instead, restore the free amine state of lysines leading to closure of chromatin and gene silencing. Thus, HDAC inhibitors (HDACi) are epi-drugs able to reactivate signaling pathways silenced by deacetylation, in cancer as well as in non-cancer diseases. To date, 11 zinc-dependent human HDAC isoforms have been classified into three classes (I, II and IV), differing in structure, enzymatic function, sub-cellular localization, and tissue expression. In addition to these “classical” HDACs, the mammalian genome encodes another group of deacetylases, known as sirtuins (class III HDACs), which hydrolyze acyl-lysines through a different mechanism involving nicotinamide adenine dinucleotide (NAD+) as co-substrate. Importantly, the HATs/HDACs have also thousands of non-histone protein substrates.16 A pharmacophoric model for HDACi has been described including i) a cap group binding at the enzyme surface and partially exposed to the solvent, ii) a spacer mimicking the lysine side-chain of the substrate, and iii) a moiety able to complex the zinc ion, crucial for the HDAC catalytic action.17

Vorinostat 3 (SAHA, suberoylanilide hydroxamic acid, Table 1) was the first pan-HDAC inhibitor (not selective for any HDAC isoform) approved by FDA in 2006 for the treatment of refractory cutaneous T-cell lymphoma (CTCL). It is characterized by an anilide cap group, a linear spacer of six methylene units and a hydroxamate function as zinc ion binder. Romidepsin 4 (FK-228, Table 1) is a natural prodrug approved by FDA in 2009 for the treatment of refractory CTCL, carrying a disulfide bridge that is reduced in vivo to release the thiol zinc binding moiety. Romidepsin 4 shows mainly inhibitory activity against class I HDACs rather than pan-inhibition. Belinostat 5 and panobinostat 6 (Table 1) are two other hydroxamate-containing pan-HDAC inhibitors approved by FDA, the first in 2014 for the treatment of refractory peripheral T-cell lymphoma (PTCL), the latter in 2015 for the treatment of refractory or relapsed multiple myeloma (MM). Tucidinostat 7 (chidamide, Table 1) is the first benzamide-type HDACi approved for clinical use. It inhibits HDAC1/2/3/10 and was approved by the Chinese FDA in 2014 for the treatment of PTCL. Sodium valproate (VPA) 8 (Table 1) is a known antiepileptic drug belonging to the short-chain fatty acid series of HDACi. It selectively inhibits class I HDACs and reduces tumor growth and metastasis formation in various animal models. Entinostat 9 and mocetinostat 10 (Table 1) are two benzamide-containing, class I-selective HDACi currently in clinical trials for the treatment of numerous solid tumors. Abexinostat 11, pracinostat 12, quisinostat 13, resminostat 14, givinostat 15 and rocilinostat 16 (Table 1) are examples of hydroxamates pan-HDACi (with the exception of 16, quite selective for HDAC6) in clinical trials for the treatment of several hematological (11-13, 15 and 16) and solid tumors (14). Among them, pracinostat 12, resminostat 14 and givinostat 15 granted the status of orphan drugs for AML, hepatocellular carcinoma, and Duchenne muscular dystrophy, respectively.1820 Nicotinamide 17 is the only sirtuin inhibitor currently used in clinics for the treatment of solid tumors (Table 1).

Recently, clinical candidates have been obtained for other epigenetic targets such as lysine methyltransferases (KMTs), arginine methyltransferases (PRMTs), lysine demethylases (KDM) and bromodomains (BRDs). KMTs constitute a large family of enzymes able to catalyze the transfer of one, two and/or three methyl groups to lysine residues using SAM as the methyl donor co-substrate. Similarly, PRMTs perform methylation (single or double, the latter symmetric or asymmetric) at arginine residues of histone and non-histone proteins.21,22

Differently from DNA methylation, lysine methylation can lead to either transcriptional activation or repression, depending on the specific lysine residue modified, and on the extent of methylation (me1, me2, or me3). GSK2816126 18, tazemetostat 19 and CPI-1205 20 (Table 1) are selective, catalytic inhibitors of both wild type (wt) and mutant forms of the methyltransferase EZH2 (enhancer of zeste homolog 2), currently in clinical trials in patients with various lymphomas, multiple myeloma, and solid tumors.23,24 Pinometostat 21 (Table 1) is a picomolar inhibitor of the H3K79 methyltransferase DOT1L (disruptor of telomeric silencing 1-like), with more than 30,000-fold selectivity against other KMTs. When used in rearranged-MLL (mixed-lineage-leukemia) cells and xenograft models, 21 reduced H3K79 methylation level, decreased MLL target gene expression, and induced selective leukemia cell death.25,26

JNJ-64619178 2227 and GSK3326595 2328(Table 1) are two potent and selective PRMT5 inhibitors that induced tumor regression in solid cancers as well as in hematologic malignancies, supporting clinical testing in patients with these kinds of cancer.

To date, two families of KDMs have been identified.27,29 The first is the KDM1 family, including LSD1 and LSD2 (lysine-specific histone demethylase 1 and 2), able to remove methyl units through an oxidative amination process using flavin adenine dinucleotide (FAD) as cofactor. The second KDM family, containing KDM2–7, is known as Jumonji C (Jmj-C) domain-containing protein family and uses an α-ketoglutarate/Fe(II) ion-dependent mechanism to catalyze the hydroxylation of a lysine N-methyl group, followed by its release as formaldehyde. It is known that increased activity or expression of members of both the KDM families are implicated in different types of cancer. Given the similarity of the catalytic mechanism of LSD1 and monoamine oxidases (MAOs), MAO inhibitors (MAOi) were investigated for their potential ability to inhibit LSD1. While the majority of MAOi failed to inhibit LSD1, tranylcypromine 24 (Table 1) is able to covalently inhibit LSD1 in the micromolar range, via radical species that originate from the opening of the cyclopropane ring.30 Two tranylcypromine analogs, ORY-1001 25 and GSK2879552 26 (Table 1), showed potency in the nanomolar range and high selectivity for LSD1 over MAOs, and entered clinical trials for the treatment of leukemia and small cell lung carcinoma.31,32

BRDs are characterized by a hydrophobic pocket that specifically recognizes the acetyl mark of acetylated lysine residues. To date, 61 different BRDs are known, including the BET (bromo- and extra-terminal) domain comprising BRD2, BRD3, BRD4, and BRDT. An enrichment of BRD4 in enhancer regions is connected to the expression of oncogenes in cancers, while translocations fusing BRD3 or BRD4 to the NUT oncogene are responsible for NUT midline carcinoma, a type of cancer with a very low life expectancy. The BET inhibitors CPI-0610 27, GSK525762 28 and OTX015 29 (Table 1) are being tested in clinical trials for a variety of cancers including multiple myeloma (27), NUT midline carcinoma (28), acute leukemia and triple negative breast cancer (29).33,34

One of the main reasons for which a single epi-target approach can be ineffective is the emergence of pharmacological resistance. Phenotypic aberrations at the base of neoplastic diseases are the result of a complicated signaling network system that evolves and changes over time to evade drug-induced events such as cell death, growth inhibition, DNA repair, and metabolic alterations. The goal of drug discovery in the epi-drug research is no longer simply the design of highly potent and selective molecules acting against a known epi-target, but is moving towards the identification of network-active compounds with multi-targeting properties. This latter strategy could be obtained by using the MTDLs approach.4

2. Multiple-Medication Therapy (MMT), Multi Compound Medication (MCM), and Multi-Target-Directed Ligands (MTDLs) approaches

The theoretical concept at the base of polypharmacology is to achieve an improved therapeutic effect hitting different targets, all responsible for the onset of a specific complex disease. All the factors and their relative interactions characterizing a molecular pathway could be considered as potential targets for a polypharmacology-based therapy. In fact, multi-target drugs can hit a cellular pathway at different levels with the goal to get a synergistic effect. This goal can be achieved by all the three aforementioned approaches that, despite having the same purpose, show marked differences. While the MMT and the MCM strategies are based on the combination/association of two or more active principles, each with its own specific target, the MTDLs strategy is focused on the development of only one molecule able to simultaneously interact with different targets.

The MMT could increase or maintain the desired therapeutic efficacy potentially requiring a lower dose of each drug, thus minimizing drug resistance and adverse reactions compared to single-drug treatment.35 One of the main disadvantages of the application of the so-called “drug cocktails” is the negative impact on patient compliance, that could be compensated by the MCM approach, which takes advantage of “polyvalent-pills” whereby two or more agents are co-formulated in a single tablet.36 However, both approaches show potential pharmacodynamic (PD) and pharmacokinetic (PK) limitations, such as drug-drug interactions and the possible different solubility of the active substances that may modify the bloodstream uptake, not guaranteeing the simultaneous presence of the required drugs at the desired concentrations in the target tissue. In addition, when the agents are used in combination or association, regulatory agencies generally require the demonstration of safety of each individual drug before clinical trials. If the drugs of interest are owned by different pharmaceutical companies, this procedure can require an extended time.2,37

Despite the common hallmarks of cancer, each tumor, as well as the therapeutic outcome of each patient, is different from the others. One of the central aspects of MMT consists in the possibility to choose different dosages, establishing a prolonged administration of drugs and increasing the number and type of potential therapeutic schedules, paving the way to a personalized treatment. The MMT approach allows to modulate the dose of each single drug within the cocktail (if necessary), according to the patient’s needs.4,3840 In light of these evidences, we can compare the MMT and MTDL approach uncovering their strengths and weaknesses. To date, despite the MTDL overcame some of the typical limitations of the MMT and MCM strategies, with in principle better results at both PK and PD level (e.g. better patient compliance, no solubility/biodistribution issues, greater efficacy, lower toxicity, etc.), the MMT still proves easier to apply and customize, due to the virtually infinite drug combination and dosage possibilities.4,40

3. The MMT approach

Epigenetic modulators, such as pan-HDAC or DNMT inhibitors, showed efficacy in the treatment of different kinds of tumors such as ovarian cancer,41 CTCL,42 breast cancer,43 and MDS.44 To date, the combination therapy is more extensively explored in the clinic than the use of multi-targeting single drugs. The first clinical success of the MMT approach dates back to 1980 and consists in the treatment of childhood acute lymphoblastic leukemia (ALL), based on the co-administration of methotrexate 30, vincristine 31, prednisone 32 and 6-mercaptopurine 33 (Figure 1).45

Figure 1.

Figure 1.

The first clinical success of the MMT based on the co-administration of methotrexate 30, vincristine 31, prednisone 32 and 6-mercaptopurine 33 for ALL treatment.

Regarding epi-drugs, it has been recently highlighted their capability (mostly HDACi and DNMTi)46,47 to increase chromatin accessibility to conventional DNA-damaging chemotherapeutic agents, markedly enhancing their effects and establishing a rationale for their combination. Several pieces of evidence showed that the CpG island methylation is responsible for the induction of drug resistance, and this effect can be reversed through the treatment with hypomethylating agents4851 including zebularine51 and decitabine 2.51,52 The combination of HDACi with cytostatic agents (such as doxorubicin and paclitaxel) has been widely evaluated in both preclinical and clinical contexts. Here, a reliable increase in the therapeutic outcome was achieved, due to the downregulation of proangiogenic and tumorigenic factors (i.e. HIF-1α and VEGF),53,54 the repression of DNA repair systems55 as well as the enhancement of apoptosis induction.5557 Furthermore, HDACi can also induce immune potentiation responses by increasing the anticancer activity of IL-2.58 The most commonly evaluated epigenetic combinations are those based on HDAC/DNMT inhibitors, since DNMTi incorporation into DNA is restricted to cell cycling cells.59 In addition, this combination could be useful due to the induction of synergic effects in the modulation of gene expression, in particular for IRF4, MYC and several related genes. It is important to underline that the downregulation of these genes was not identified after either HDACi or DNMTi treatment alone.60,61 A lot of preclinical evidences collected in AML or MDS cell lines or in ex vivo cultured patient cells, such as improved cell growth arrest, loss of clonogenic potential and DNA synthesis inhibition, support the combination of HDACi with the nucleoside DNMTi azacytidine 1 or decitabine 2.6264 HDACi were able to sensitize cancer cells to cellular stress-based radiotherapy and hypothermia treatment through induction of apoptosis and arrest of protective cell mechanisms.65,66 The combination of modulators of the histone acetylation and DNA methylation status could be useful also to counteract the onset of cancer resistance.67

3.1. New preclinical studies (2018) based on epigenetic modulators in combined anticancer therapy

Combination therapy based on decitabine 2 and vorinostat 3 (combination A, Figure 2) has been recently evaluated against pancreatic cancer by Han et al.,68 demonstrating improved effects in vitro when compared to the single drug application. Treatment with vorinostat 3 or decitabine 2 alone induced pancreatic cancer cells cycle arrest, as well as inhibition of migration and proliferation. Importantly, while vorinostat 3 induced cell apoptosis also as monotherapy, decitabine 2 alone was unable to cause this effect. Combination A led to stronger effect compared to each epigenetic agent alone, resulting more effective in proliferation and migration inhibition, G2 cell cycle arrest, apoptosis induction, and in enhancing tumor suppressor genes (P16 and TP53) expression levels. Han et al. further reinforced this evidence postulating the inhibition, by both drugs, of the PI3K/AKT/PTEN signaling pathway.68

Figure 2.

Figure 2.

Combinations between HDACi and DNMTi (A, B, and D), and between HDACi, ATRA and doxorubicin (C).

Malignant pleural mesothelioma (MPM) is a rare and strongly aggressive cancer, and many shreds of evidence highlighted that it is mostly associated with asbestos exposure.69 Unfortunately, none of the available treatments is effective, due to developed resistance to all conventional therapies. However, recently epigenetics has been claimed to play a role in this cancer, suggesting the use of epi-drugs for MPM treatment. As a proof of evidence, global hypomethylation and specific promoter hypermethylation (including those of tumor suppressor genes) joined to a decrease in the acetylation status of histone H3 and H4 have been observed.7073

In previous works focused on a mouse model of MPM, either decitabine 2 or HDACi (sodium valproate 8 or vorinostat 3) monotherapy proved to be non-effective exerting only weak anticancer activity, while their combination induced antitumor immune response through the expression of tumor anti-antigen factors, especially cancer testis antigen (CTA).74,75

Starting from the adverse reactions (first of all, hematological toxicity)76 correlated to HDACi, Bensaid et al.,77 in order to identify new suitable HDACi for clinical use, tested four recently developed HDACi,78,79 two hydroxamates (ODH and NODH 34, Figure 2) and two benzamides (ODB and NODB), comparing their toxicity and anticancer effect against MPM with those of vorinostat 3 and sodium valproate 8. The same authors also studied the immunogenic capability of the novel HDACi in combination with decitabine 2, also measuring the mRNA expression of PD-L1 (programmed death-ligand 1), CTA, RIG-I (retinoic-acid-inducible protein I) and MDA5 (melanoma-differentiation-associated gene 5) in MPM cells.

All the tested HDACi were toxic for immune cells, being responsible for the death of 100% CD8 T-cells clones on total lymphocytes at high doses. The combination of the new HDACi NODH 34 and decitabine 2 (Combination B, Figure 2) seemed to be the most promising. In fact, NODH 34 was 100-fold more effective than vorinostat 3 in enhancing histone H3 acetylation in cells, being in addition less toxic. Since this combination induced PD-L1 expression, also the association with anti-PD-L1 agents should be tested with the aim to improve the anti-tumor immune response.77

Compared to other breast cancer subtypes, the triple negative breast cancer (TNBC) is the most aggressive as well as the most difficult to treat. Recently, Merino et al.80 reported a triple combination based on all-trans retinoic acid 35 (ATRA), entinostat 9 and doxorubicin 36 (Combination C, Figure 2), that resulted able to induce differentiation and cell death as well as a significant regression of tumor in TNBC xenograft models. The combination of doxorubicin 36 and entinostat 9 induced the downregulation of cell cycle genes (such E2F, G2M and MYC) and the upregulation of immune checkpoint agonists and cancer testis antigens as well as interferon genes. Moreover, such combination was responsible for an altered expression of genes related to inflammation, differentiation and growth arrest, through topoisomerase II-β inhibition (by doxorubicin 36) and decreased expression (by entinostat 9).81 In particular, entinostat 9 sensitized the cells to the doxorubicin 36 treatment, enhancing G2 cell cycle arrest82 and inducing a more effective up-regulation of cytokines, costimulatory molecules, other inflammatory genes, and tumor antigens with a consequently increased interferon response. The single drug treatment with ATRA 35 led to breast cancer cell death,83 differentiation84,85 and inflammation.86 However, its limited success in solid tumors87 could be due to the frequent epigenetic silencing of the retinoic acid receptor beta (RAR-β).88,89 Different studies evidenced that RAR-β is epigenetically silenced and under-expressed in breast cancer, while the treatment with HDACi typically induces RAR-β re-expression, sensitizing the cells to ATRA treatment.90,91 The triple combination described above, in addition to the effects observed with the doxorubicin 36/entinostat 9 combination, markedly induced cancer stem cell differentiation, cell death and necrosis,81 and increased the expression of different genes playing important roles in the inflammation process, such as DHRS3, IL-1β, CCL2692, ELF3,93,94 IL-1α, TNF-α and CD14.

In TNBC patients, the expression of epithelial-mesenchymal transition (EMT) markers and cancer stem cells with CD44+CD24−/low phenotype are associated with a poor outcome.95,96 Several pieces of evidence support the involvement of epigenetic mechanisms in the activation of EMT and in particular in the modulation of E-cadherin expression. The hypermethylation of the CDH1 promoter plays an essential role in EMT regulation. The E-cadherin expression is also repressed by different EMT inducing factors such as ZEB1, ZEB2, TWIST, SNAIL and SLUG,97100 and this process is regulated by HDACs,98,101104 establishing a rationale for the application of an epigenetic combination therapy for TNBC treatment.105 Exploring EMT modulation in their unique TNBC cell model,106 Su et al.107 decided to test six HDACi (vorinostat 3, entinostat 9, SB939, LBH589, quisinostat 13, tubastatin A) and three DNMTi (azacytidine 1, decitabine 2 and SGI-110 37). These epigenetic modulators were first tested alone in MTT, migration and invasion, three dimensional culture, and colony formation assays. Once selected as the most potent of them, entinostat 9 and SGI-110 37 were used in combination to reprogram TNBC cells, attenuating the EMT process by E-cadherin induction in mesenchymal-like cancer cells. The combination of entinostat 9 and SGI-110 37 (Combination D, Figure 2) showed improved results in cancer cells motility, colony formation, proliferation, apoptosis, and stemness when compared to the use of single agents alone. From a pharmacological point of view, the combination D synergistically suppressed EpCAM cleavage and Wnt signaling, counteracted EMT, reduced the expression of mutant p53, EZH2 and ZEB1, and induced apoptosis as well as H3 hyperacetylation in TNBC.107 In vivo studies showed that entinostat 9 alone, as well as the 9/37 combination, decreased XtMCF xenograft tumor weight in CB17/SCID mice of about 52%. Differently, in NOD/SCID mice the combination D (at higher doses, but with a shorter treatment) reduced tumor weight by 33% compared to the control, while the single target inhibitors 9 and 37 administered alone did not produce significant effects on tumor growth.107

Regular consumption of bioactive dietary compounds may have the capability to modulate different epigenetic mechanisms, thus playing also an active role in cancer prevention and therapy.108111 Genistein 38 is an isoflavone naturally contained in soybean, able to inhibit DNMT and HDAC enzymes and showing anticancer properties in cells.112115 Recently, genistein 38 has been reported to: i) arrest breast cancer cells cycle in G2/M phase inducing the expression of p21WAF1/CIP1;116 ii) re-sensitize ER-negative breast cancer cells to estrogen-targeted chemotherapy through the activation of estrogen receptor (ER);117 iii) increase the acetyl-H3 and acetyl-H4 levels as well as dimethyl- and trimethyl-H3K4 at the promoter region of the BTG3 gene in renal cancer;114 iv) increase the acetylation status of H3K4 at the p21WAF1/CIP1 and p16INK4a transcription start sites, mediated by induction of HATs in prostate cancer.118

Sulforaphane 39, abundant in kale and broccoli sprouts, is a HDACi112,113,119 effective against breast cancer stem cells both in vitro and in vivo suppressing the Wnt/β-catenin pathway.76,120 In addition, Meeran et al. showed that sulforaphane 39 induced hTERT repression in breast cancer121 with chemo preventive effects through the induction of cell cycle arrest via Kruppel like factor (KLF4) transcription mediation.122

Previous studies conducted by Meeran et al. revealed that a diet containing compounds with HDAC and DNMT inhibitory activities altered the epigenome, resulting in the modulation of various aberrant epigenetics patterns in tumor-related genes.76,121,123 Therefore, Paul et al. decided to evaluate the effects deriving from the combination of genistein 38 and sulforaphane 39 (Combination E, Figure 3).124 Genistein 38/sulforaphane 39 combination decreased the HDAC2 and HDAC3 expression both at protein and mRNA levels, downregulating KLF4, essential in stem cells formation, as well as hTERT that is known to be activated when KLF4 binds its promoter region. Combination E resulted more effective than the sinle agents alone in decreasing cell viability, lowering colony formation and inducing apoptosis in breast cancer cell lines (MCF-7 and MDA-MB-231).

Figure 3.

Figure 3.

Combinations between genistein with broccoli sprouts (E), HDACi with platinum complexes (F), BETi with MEKi (G), or with HDACi (H-J), or with BCL2i (J).

The aim of the study conducted by Paul et al.124 was to evaluate the in vivo anticancer effect of 38 and 39 taking advantage of a diet based on food rich in these two bioactive compounds. Concerning genistein 38, the authors could not use soy feeding, because it contains many other active compounds, making it difficult to dissect the role of 38. Thus, the pure 38 was used for the study. Differently, broccoli sprouts contain a 50-fold higher concentration of glucoraphanin, which is converted to the isothiocyanate form (sulforaphane 39) by myrosinase, thus they can be used as 39 source.125,126 Combinatorial 38/broccoli sprouts dietary regimen, at a physiologically accessible concentration by daily consumption,54,127 was orally administered to an ER(−) transgenic breast cancer mouse model and compared with the two bioactive compounds alone. In vivo results supported the in vitro evidence, since the combination E was more effective than the single drug treatment in reducing tumor size and volume and in extending tumor latency. Future plans are focused on the evaluation of this combination in other kind of cancers such as ER(+) and BRCA1 mutated TNBC.124

The combination of platinum-based compounds/HDACi against non-small cell lung carcinoma (NSCLC) founded its rationale mainly in two findings: i) HDACs play a main role in resistance and adaptation to genotoxic treatment in different ways;128135 ii) platinum chemotherapy is the first choice for the treatment of such tumors.136138 A dose escalation phase I clinical study was conducted for panobinostat 6/etoposide/cisplatin triple combination. However, since a dose-limiting febrile neutropenia and thrombocytopenia was observed at the first dose level of panobinostat 6 (10 mg orally three times weekly on a 2 week on/1 week off schedule), it was not possible to determine a phase II starting dose.139 The unacceptable toxicity was probably due to the overlap of adverse reactions induced by each drug, suggesting the potential evaluation of new safer combination therapies. Despite the choice of carboplatin 40 is favored in terms of toxicity, its lower potency requires the combination with other drugs in order to increase its anticancer activity. Recently, Wang et al.140 proposed to combine carboplatin 40 with panobinostat 6 getting a strong improvement in cell proliferation reduction with IC50 values ranging between 4–21 nM in different NSCLC cell lines such as H460, H226, A549, Calu-1 and SKMES-1. These promising outcomes are likely the result of the pleiotropic effects of panobinostat 6 which, in addition to the modulation of histone acetylation, has also hyperacetylating effects on non-histone substrates including proteins that play a fundamental role in carcinogenesis such as STAT3, HSP90, p53, E2F, NF-kB, c-Myc and HIF-1α.136138,141 Most importantly, the hyperacetylation mediated by panobinostat 6 at HSP90 level restores the physiological degradation of different oncoproteins, including epidermal growth factor receptor (EGFR) and its downstream transducers.142 In addition, it has been shown that HDACi reverse the repression of miRNAs facilitating the degradation of EGFR mRNA,143,144 suggesting a potential synergistic effect between panobinostat 6 and EGFR inhibitors against EGFR-mutant NSCLC.145,146 Tumorigenic mouse models also confirm the chemo sensitizer effect induced by panobinostat 6. Indeed, the carboplatin 40/panobinostat 6 combination (Combination F, Figure 3) led to a markedly stalled disease progression by 92% compared to the negative controls, that also resulted greater than that caused by carboplatin 40 (28%) or panobinostat 6 (54%) alone.140

To date, only a few effective therapeutic options for the treatment of NRAS-driven melanoma are available.147 A lot of clinical studies have been conducted focusing on RAS effectors, in particular on inhibitors of phosphatidylinositol-3 kinase (PI3K) signaling pathways148 and of mitogen-activated protein kinase (MAPK) such as MEK inhibitors,148 but their therapeutic potential as single agents was modest. Despite the introduction of new therapeutic approaches for the treatment of BRAF-mutant melanoma (first of all immunotherapy treatment),149 the acquisition of resistance is a significant problem and, in particular, NRAS-mutant (NRASMut) melanoma treatment is one of the hardest challenge in this field.150154 Echevarría-Vargas et al.155 demonstrated that BET proteins are overexpressed in NRASMut melanoma and, in particular, increased levels of BRD4 mRNA are associated with poor patient survival. Starting from these shreds of evidence, they evaluated the potential of the combination of BETi (OTX-015 29 or JQ1) and MEK inhibitors (the FDA-approved trametinib 41 or PD901, in phase II clinical trial) restraining NRASMut melanoma and offsetting drug resistance (Figure 3). The dual targeting of BET/MEK synergistically prolonged NRASMut tumor-bearing mice survival without apparent toxicity, downregulated the expression of the transcription factor TCF19, which is essential for melanoma cells survival, sensitizing it to apoptosis, and mitigated a transcriptional signature associated with innate resistance to already used inhibitors and immune checkpoint agents. The treatment with the BETi/MEKi combination showed the strongest efficacy and tolerability in patient-derived xenograft and immunocompetent syngeneic mouse models. While BETi alone induced mainly cytostatic effects in NRASMut melanoma cells in vitro, minimal effects have been registered in in vivo models. However, the combination of BETi/MEKi was able to induce robust anti-tumor effects in NRAS-mutant melanoma. Considering that the MEK inhibitor trametinib 41 is a FDA-approved drug for BRAF-mutant melanoma patients, and different BETi such as OTX-015 29 are in advanced clinical trials, clinical studies based on the combination of trametinib 41 and OTX-015 29 (Combination G, Figure 3) could be rapidly implemented for therapy-resistant melanoma.155

A panel of HDACi (vorinostat 3 and romidepsin 4) and BETi (OTX015 29, JQ1, I-BET762 and CPI-0610 27) were selected by Zhao et al.156 to evaluate the effectiveness of their combination for the treatment of CTCL. The combinations resulted effective since induced apoptotic events, in addition to G0/G1 cell cycle arrest as well as reduction of the expression levels of fundamental cancer players such as cyclin D1, NF-kB, c-Myc and IL-15Rα. In particular, the combination based on OTX015 29 (125 nM) and romidepsin 4 (1 nM) (Combination H, Figure 3) showed the greatest apoptotic effect (60–80%, while OTX015 29 or romidepsin 4 alone induced only 6% or 31% apoptosis, respectively) at 96 hours of treatment, resulting the most promising combination for clinical use. Ex vivo studies on leukemic CTCL cells obtained from patients with Sezary syndrome resulted in higher levels of apoptosis (about 60–90%) following the combined treatment rather than single agents administration. During the use of Combination H, normal CD4+ T cells were minimally affected (10%).

Recently, other preclinical studies based on the simultaneous administration of BETi and HDACi in melanoma cells as well as in murine models of pancreatic ductal adenocarcinoma have been described recently.175,157159 In 2018, the combination of the BETi JQ1 42 and romidepsin 4 (Combination I, Figure 3) in urothelial carcinoma (UC) inhibited cell cycle progression, suppressed clonogenic growth, and induced apoptosis. In the same system, the treatment with JQ1 42 alone was able to induce cell cycle arrest but only limited apoptosis. At a molecular level, in UC cells the JQ1 42/romidepsin 4 combination increased H3K27 acetylation but decreased H3K4 trimethylation, downregulating anti-apoptotic and oncogenic factors such as survivin, BCL-2, BCL-XL, c-MYC, EZH2 and SKP2 and diminishing AKT phosphorylation.174

Preclinical studies conducted by Kim et al. highlighted the potential of the therapeutic combination between BETi and HDACi or inhibitors of the antiapoptotic protein B-cell lymphoma 2 (BCL2i) in the treatment of advanced CTCL. BCL2 was previously suggested as a targetable pathway on the basis of the common gene alterations that increase BCL2 activity and dependence, including TP53 deletion, STAT3 and STAT5B amplification, and CTLA4 deletions.160164 To date, CTCL is generally considered incurable. Indeed, the overall response rates to single agent systemic therapies (such as the retinoid bexarotene or HDACi) range between 20–45% and relapses are common.165,166 The MYC oncogene is often amplified in CTCL condition, and the capability of BETi to repress its expression has been widely discussed in literature, in particular for AML, multiple myeloma (MM) and Burkitt’s lymphoma.167170 BCL2i are approved for relapsed or refractory chronic lymphocytic leukemia (CLL) with 17p deletion,. They induced apoptosis in patient-derived CTCL cells in vitro, and this effect resulted synergistically potentiated using HDACi in combination.171,172 In addition, drug cocktails based on BETi and other epi- and/or non epi-targeted agents, such as the combination of the BETi JQ1 42 and the BCL2i navitoclax during the treatment of MYCN-amplified SCLC,173 or the combination of vorinostat 3 or romidepsin 4 and venetoclax against CTCL,172 have been reported for the treatment of solid and multiple hematologic tumors.

Kim et al. showed that the cytotoxic effects induced by BETi (such as JQ1 42 or ABBV-075 43)174 resulted synergistically amplified by either HDACi (vorinostat 3 or romidepsin 4) or BCL2i (venetoclax 44) (Combination J, Figure 3) in the majority of both CTCL cell lines (MyLa, Sez4, HH, Hut78) and CTCL tumor samples derived from patients who had previously tried both single and multiple therapies without success. In this study, synergistic effects on the modulation of the genes BCL2 and MYC, and of genes encoding for pro-apoptotic and cell cycle regulation factors (BIM, CDKN1A and p21), as well as a lower level of BCL2L1 expression were observed.175 In addition, a substantial increase in caspase 3/7 activation was registered when the BETi JQ1 42 was combined with either romidepsin 4 or vorinostat 3, confirming that the observed synergies are mainly due to apoptosis induction.175

Preclinical studies conducted by Wang and coworkers176 demonstrated the therapeutic potential of the combination of EZH2i and HDACi against small cell carcinoma of the ovary hypercalcemic type (SCCOHT), a rare but extremely lethal cancer that interests mainly young women. This kind of tumor shows a diploid genome with loss of SMARCA4 and lack of SMARCA2 expression, two mutually exclusive ATPases of the SWI/SNF chromatin-remodeling complex. Previously, preclinical studies showed that a subset of SCCOHT patients took advantage from oncolytic virus177 or c-Met inhibitors.178 In agreement with the known antagonism between the PRC2 and the SWI/SNF complex, it has been recently demonstrated that SCCOHT cells are highly sensitive to EZH2 inhibitors, being EZH2 the catalytic subunit of the PRC2 complex.179,180 However, the EZH2i tazemetostat 19 only induced disease stabilization or partial response in two SCCOHT patients previously treated with a standard chemotherapy (www.epizyme.com). Such failure can be explained by the fact that SCCOHT cells survival depends on additional epigenetic pathways, whose identification and targeting could help in improving the response of SCCOHT cells to EZH2 inhibition. Through a rational epigenetic drug screen, Wang et al. highlighted that pan-HDACi markedly suppress SCCOHT cells growth. In fact, in this cancer type, pan-HDACi such as quisinostat 13181 reversed the expression of a group of deregulated proteins, as a consequence of the SMARCA2 and SMARCA4 deficit, inducing apoptosis, differentiation and growth arrest in vitro, and suppressing growth of xenograft tumors of SCCOHT cells. A combined treatment based on the use of quisinostat 13 and tazemetostat 19 (Combination K, Figure 4) at sub-lethal doses synergistically induced targeted gene expression (such as genes dysregulated by SWI/SNF remodeling complex in SCCOHT) as a result of H3K27 hyperacetylation at the promoter region of PRC2, by simultaneous inhibition of both deacetylase and methyltransferase enzymes. Such H3K27 hyperacetylation led to growth suppression and apoptosis of both SCCOHT cells and xenografted tumors.176

Figure 4.

Figure 4.

Combinations of HDACi and EZH2i (K), DNMTi with a conventional anticancer agent (L), or with monoclonal antibody (N), or with BCL2i and eventually antifungal agent (P, Q), HDACi with monoclonal antibody plus conventional anticancer agents (M), HDACi plus antiandrogen anticancer agent (O).

Recently, we reported a novel quinoline-based non-nucleoside DNMTi, MC3343 45, which showed promising preclinical effects against osteosarcoma. Our compound markedly induced cell proliferation arrest in both patient-derived cell lines and xenograft models. While the antiproliferative activity of MC3343 45 was comparable with that of azacytidine 1, its cell differentiating effect was much more evident. MC3343 45 selectively increased the expression of osteoblastogenesis related-genes (including OCN, ALP and COL1A2) leading to matrix mineralization, an unregistered effect after azacytidine 1 treatment.182 In addition, MC3343 45 has been tested in combination with two of the major chemotherapeutic drugs selected for osteosarcoma treatment, cisplatin and doxorubicin 36, promoting doxorubicin-DNA bond as well as DNA damage and cell death. In particular, the doxorubicin 36/MC3343 45 combination (Combination L, Figure 4) inhibited PDX-OS#1-C4 cells growth by 65% after 24 h, compared to 32% growth inhibition obtained with doxorubicin 36 alone. Considering that none of these differentiating effects has been registered with azacytidine 1, the use of MC3343 45 alone and mostly in combination for osteosarcoma treatment could be very encouraging and should be further explored in the near future.182

Soft tissue sarcomas (STSs) are rare forms of tumor involving different connective tissues originating primarily from mesoderm. Surgery with or without radiation therapy is the first choice for the treatment of localized sarcoma. Sarcoma growth depends on various factors, first of all neoangiogenesis,183 therefore the application of specific antiangiogenic drugs, such as the recombinant human monoclonal antibody against vascular endothelial growth factor bevacizumab,184,185 has been studied in various clinical trials. In particular, the combination therapy based on the two cytotoxic drugs gemcitabine and docetaxel with bevacizumab was evaluated in patients with recurrent or advanced STS (). The treatment resulted safe, and the response rate was up to 31% for chemotherapy naïve patients and the median response duration was about 6 months,186 while treatment with bevacizumab alone resulted only in 13% response rate patients with epithelioid hemangioendothelioma or metastatic or advanced angiosarcoma.187 Monga et al.188 based on the idea that the pro-angiogenetic effect of HDACs joined to tumor microenvironment could block the effect of bevacizumab, added VPA 8 to the combination of bevacizumab, gemcitabine 46, and docetaxel 47 (Combination M, Figure 4), with the aim to increase the therapeutic response suppressing chemo resistance. Among the patients that did not respond to the treatment with gemcitabine/docetaxel alone, 61% responded (completely or partially) to the combination M. Furthermore, the addition of these two agents should be considered for therapeutic applications in patients whose disease progresses after two cycles of gemcitabine/docetaxel. This pilot study reveals the possibility to associate to the standard cytotoxic chemotherapy drugs that influence the tumor microenvironment, and pushes towards the repositioning of old agents. The most common adverse reactions that have been registered were hypertension linked to bevacizumab treatment, neuro and liver toxicity connected with VPA 8, as well as cytopenia related to gemcitabine 46 and docetaxel 47 treatment. This combination appears to be moderately safe and of considerable potential in the next future.188

Hypomethylating agents could increase the cytotoxicity of the anti-CD33 immunoconjugate gemtuzumab ozagamicin (GO)189 against AML through the modulation of SHP-1 and Syk expression,190,191 as well as reducing the expression of p-glycoprotein which mediates the resistance to GO treatment.192194 Based on these evidences, a phase I/II trial based on the combination of GO and azacytidine 1 (Combination N, Figure 4) was conducted ().195 GO was approved in 2000 as monotherapy for the treatment of relapsed AML but was later retired from the US market (in 2011) for its lack of efficacy and toxicity. The aforementioned clinical study has been started before GO market withdrawal, and the dose (6 mg/m2) applied was the same used in therapy at that time. In 2017, GO was approved by the FDA in combination with citarabine 2 and daunorubicin for the treatment of AML, showing encouraging results. Fifty adult patients (29–82 years) affected by refractory or relapsed AML where initially treated with azacytidine 1 followed by GO. No dose-limiting toxicities were registered in phase I, however significant neutropenia and infection were the most common adverse reactions. While the optimal dose of GO remained unclear, the maximum tolerated dose adopted was of 75 mg/m2 of azacytidine 1 (daily for six consecutive days) followed by 6 mg/m2 of GO intravenously on days 7 and 21. After this treatment, 12 patients (24%) obtained complete remission or complete remission with incomplete peripheral blood recovery. These data are very similar to those obtained by administration of GO alone at higher doses, 9 mg/m2 given twice.196 Therefore, additional studies are required to determine the optimal schedule and dose of GO in combination with azacytidine 1.195

Histone methylation, as well as chromatin acetylation, are strongly implicated in androgen receptor (AR) protein functions and in mRNA splicing and transcription.197,198 In particular, the balance between HDACs and HATs activities is very important for tuning transcriptional regulation of AR and its downstream targets.199,200 HDACi modulate the expression of different factors and related pathways that play an essential role in cancer progression and resistance.197,198 Different evidences showed that VPA 8 increased caspase-2/3 activation and decreased proliferation in AR-negative PC3 and DU145 cells.201203 The most widely reported prostate cancer resistance mechanism is the upregulation of the AR.204206 Preclinical shreds of evidence suggested that resistance to bicalutamide 48 treatment could be overcome through the simultaneous administration of panobinostat 6, synergistically inducing apoptosis and growth arrest better than those obtained with panobinostat 6 alone, in both isogenic cell line and xenograft model of castration-resistant prostate cancer (CRPC).203,207209 In 2018, a phase I/II clinical trial () conducted by Ferrari et al.210 highlighted the capability of panobinostat 6 to restore sensitivity to bicalutamide 48 (Combination O, Figure 4) in a CRPC cancer model, as well as the safety and efficacy of its combination with bicalutamide 48 in CRPC second-line androgen resistant condition. The phase I trial was characterized by a 3 × 3 panobinostat 6 dose escalation design and the phase II trial involved randomized 55 patients treated with 20 or 40 mg of panobinostat 6 tri-weekly for 2 weeks in association with bicalutamide 48 50 mg/day in 3-week cycle. The combination of 40 mg panobinostat 6 and bicatulamide 48 showed the best results inducing synergistic antitumor effect and reduced AR activity overcoming resistance as previously postulated.

AML development is highly frequent in elderly subjects, to whom the chemotherapy treatment, widely used for young patients, did not show the desired effects due to the common insurgence of pharmacological resistance211 and comorbidity with other diseases that preclude the correct functioning of different organs, compromising the whole condition.212214 B-cell lymphoma 2 (BCL-2) protein overexpression has been found in leukemia stem cells,215 and it has been associated with resistance to chemotherapy treatment and poor patients outcomes.216 The potent and selective orally bioavailable BCL2 inhibitor (BCL2i) venetoclax 44 showed modest anticancer activity when evaluated as single therapy, since only 16 out of 32 patients affected by refractory and relapsed AML involved in the phase 2 clinical trial achieved an overall response.217 Dose-escalation phase Ib study involving 57 elderly patients (> 65 years) based on venetoclax 44 and decitabine 2 or azacytidine 1 (Combination P, Figure 4) combination resulted well tolerated showing high survival and low early mortality ().218 The dose-escalation stage has been conducted with the aim to evaluate the PK and the safety of this combination. More in detail, this phase involved three groups: group A (venetoclax 44 400 mg (cohort 1) + intravenous decitabine 2 20 mg/m2; days 1–5 of each 18-day cell); group B (venetoclax 44 400 mg (cohorts 2 and 3) or 1200 mg (cohort 4) + intravenous or subcutaneous azacytidine 1 75 mg/m2; days 1–7 of each 18-day cell); group C (venetoclax 44 400 mg + decitabine 2 substudy with oral CYP3A inhibitor posaconazole 49 300 mg twice on cycle – day 1 and 21 - and 300 mg once daily from cycle 1 – days 22–28). AML makes the affected subjects more susceptible to fungal infections, therefore posaconazole 49 resulted effective to prevent this condition.219 Moreover, since venetoclax 44 is a substrate of CYP3A, group C was designed to evaluate the potential interaction of posaconazole 49 on the PK and safety of venetoclax 44 treatment. The collected evidences supported the use of the antifungal agent posaconazole 49 in combination with venetoclax 44 and decitabine 2 (Combination Q, Figure 4). Group A and B showed similar results in terms of safety and no dose-limiting toxicities, and the maximum tolerated doses were individuated but dose escalation was stopped at 1200 mg due to gastrointestinal toxicity. As a result, 60% of patients get complete remission or complete remission with incomplete marrow recovery at all dose levels in group A and B, and 65% in the 400 and 800 mg dose cohorts with both selected hypomethylating drugs. The most common, dose-dependent adverse reaction recorded was the neutropenia due to venetoclax 44.218

4. The MTDL approach.

Multi-targeting single drugs are the result of the conjugation of two or more active molecules that individually show a known activity against specific targets. This can be achieved in three different ways:3,4

  • connecting molecules with a linker that is cleaved in physiological conditions (mutual pro-drugs);

  • connecting molecules with a stable linker, allowing each molecular moiety to interact with its specific target without interfering with the interactions established by the other portion;

  • merging molecules together, preserving and connecting in the final MTDL only functional and reactive groups (warheads).

A MTDL could specifically act on one epi-target and another target not related to epigenetics, or on more epigenetic targets. Moreover, the active components of a MTDL can interact with molecular factors belonging to the same or different cellular pathways, in order to increase the therapeutic efficacy especially in diseases characterized by a frequent insurgence of acquired resistance during the therapy.

Most of the MTDL developed so far are hybrid molecules characterized by the scaffold of HDACi linked to another drug able to hit a related target involved in cancer. Pharmacophore portions of HDACi can be linked to other epi-drugs as well as to conventional cytotoxic agents (or to other drugs) to generate HDAC-based multi-target agents. The preference of HDACi’s structural elements among various epi-drugs to design multi-targeting agents are due to the following reasons: i) in the 2000–2010 years there was a boom of studies on HDACi by medicinal chemists belonging to both industrial and academic institutions, and many HDACi characterized by a wide structural variety have been reported; ii) the cap group that characterizes all HDACi lays out of the catalytic tunnel of the enzymes and is tolerant of a high degree of structural variation without compromising the enzyme inhibition activity.220,221 This allows the introduction in this portion of a specific warhead able to interact with another target different from HDACs.35

4.1. Dual HDAC/kinase inhibitors

Initially, in the tyrosine kinase inhibitors (TKi) drug discovery field, the aim was to synthesize highly selective inhibitors. Today, it is known that better results can be obtained with TKi able to simultaneously hit different kinase targets. In fact, on the one hand it is true that in some types of cancer the deregulation of the activity of a specific kinase seems to be the driving force. On the other hand, the treatment based on the use of a selective TKi may be useful, but after a prolonged treatment it can lose its efficacy because of new mutations and redundancies in biological networks that can compensate the drug effects via the activation of new aberrant cellular pathways. Thus, simultaneous inhibition of several kinases has been recently considered as the best choice for a successful treatment.222,223 Furthermore, since co-administration of HDACi and TKi in in vitro and in vivo models have already shown synergistic effects (see above), the design of dual HDAC/TK inhibitors has become the most popular choice for the design of epigenetic multi-target drugs.224,225 Some of the approved TKi were a source of inspiration for the design of novel dual HDAC/TK inhibitors. These hybrid compounds are characterized by a hydroxamate or a benzamide moiety, that acts as HDAC inhibiting zinc binding group, connected to kinase inhibitors by different linkers.

Two HDAC/TK dual inhibitors entered clinical trials for cancer treatment. One of them, CUDC-101 50 (Figure 5) was obtained by chemical manipulation of erlotinib 51 (Figure 5) through the introduction of the hydroxamate function separated from the quinazoline scaffold by a hexamethylene linker (typical of vorinostat 3). CUDC-101 50 inhibited HDACs as well as EGFR, platelet-derived growth factor receptor (PDGFR), and the human epidermal receptor 2 (HER2) in the nanomolar range.226 Furthermore, it regulated the proliferation and migration state of 51-resistant NSCLC HCC827 cells through the down-regulation of E-cadherin, hepatocyte growth factor receptor (MET), protein kinase B (AKT) and human epidermal growth factor receptor 3 (HER3) compensatory pathways, which enable cancer cells to escape the effects of conventional EGFR/HER2 inhibitors.227 Moreover, CUDC-101 50 induced caspase-dependent apoptosis in anaplastic thyroid cancer (ATC) leading to inhibition of cell proliferation and migration in vitro, through inhibition of MAPK signaling and histone deacetylation.228 CUDC-101 50 is currently in phase I clinical trials to evaluate its safety and oral tolerability in cancer patients (), and in phase Ib open label study in patients with advanced head and neck (alone or in combination with radiotherapy and cisplatin, ),229 liver, gastric, and breast cancer (), as well as NSCLC ().228230

Figure 5.

Figure 5.

Rational design of the HDAC/TKi CUDC-101 and CUDC-907, in clinical trial for the treatment of cancer. IC50 values for the different targets are indicated, when available. The HDAC inhibitory portions are depicted in red.

The phosphatidylinositol 3-kinase (PI3K) plays an essential role in physiological functions such as proliferation, differentiation, protein synthesis, motility, and apoptosis.231 When activated, it is also responsible for the onset of numerous diseases such as indolent BCL and aggressive BCL. Single-drug therapies with either PI3Ki or HDACi were often ineffective for the treatment of diffuse large BCL, because of concurrent activation of other survival- and growth-related pathways, while simultaneous inhibition of the two targets, PI3Ki and HDACi, displayed a synergistic effect. CUDC-907 52 (Figure 5), the dual PI3K/HDAC inhibitor based on the structure of the PI3K inhibitor pictilisib 53 (Figure 5), is an orally active dual inhibitor potent at nanomolar level against both PI3Ks and HDACs [IC50 values = 19 (PI3Kα), 54 (PI3Kβ), 1.7 (HDAC1), 5 (HDAC2), 1.8 (HDAC3), 2.8 (HDAC10) nM]. CUDC-907 52 was effective against chronic lymphocytic leukemia (CLL) through PI3K/HDAC inhibition joined to suppression of STAT3 and RAF/MEK/ERK signaling, thus reducing the expression of anti-apoptotic BCL2-family members MCL-1, BCL-XL and BCL-2. In addition, cytokines such as APRIL and BAFF as well as their receptors BMCA, TACI and BAFFR resulted downregulated after the treatment, blocking, by this way, NF-κB and BAFF-induced signaling. Thus, the pleiotropic effects induced by CUDC-907 52 abrogate different protective and pro-survival signals, rescinding microenvironment protection and highlighting the possibility of sensitizing CLL cells to other drugs. Accordingly, a low concentration combination therapy based on CUDC-907 52 and different inhibitors of NF-κB (IMD-0354), BTK (ibrutinib) and BCL2 (ABT-199) signaling pathways leads to a potent synergistic effect.232 Different pre-clinical data in hematologic and solid cancer cell lines have shown that the effects obtained with CUDC-907 52 in terms of tumor growth inhibition and multiple pro apoptotic activities are more potent than those achieved using single PI3K or HDAC inhibitors. In particular, a treatment with CUDC-907 52 gave a reduction in MYC gene expression in MYC-driven tumor model, whose overexpression has been reported to be among the worst prognostic factors in relapsed refractory diffuse large B-cell lymphoma (DLBCL).233 Dose escalation and expansion of CUDC-907 52 in DLBCL and MYC-altered disease, with and without the monoclonal antibody rituximab (25 patients received monotherapy treatment while 12 received the combination) () showed a median duration of the overall response of 11.2 months (of 13.6 months in MYC-altered patients, 7.8 months in those with unknown MYC status, and 6.0 months in MYC unaltered patient), a durable antitumor activity (in particular in MYC-altered patients) as well as a safety and tolerable profile.234 CUDC-907 52 has been reported to downregulate FLT3 expression in AML cells. Recent preclinical studies conducted by Knight et al.235 confirmed that 52 synergistically increased the apoptosis induced by the novel dual FMS-like tyrosine kinase 3 (FLT3) and AXL receptor tyrosine kinase (AXL) inhibitor gilteritinib in both FLT3-ITD AML cell lines and FLT3-ITD patient samples. Such synergistic antileukemic activity occurred thanks to the downregulation of gileritinib-induced FLT3 expression in AML cells mediated by CUDC-907 52. In addition, this combination cooperatively inhibits the JAK-STAT, PI3K-AKT and RAS-RAF signaling pathways, preventing the escape via alternative pathways and providing a strong foundation for subsequent in vivo murine studies. To date, CUDC-907 52 is in phase I clinical trials for the treatment of solid tumors and lymphomas (, , , ).236

Activation of phosphatidylinositol-3-kinase-Akt-mammalian target of rapamycin (PI3K/Akt/mTOR) signaling pathway is responsible for the onset of hepatocellular carcinoma (HCC), and of the acquired resistance to sorafenib, the first and the only approved systemic TKi drug237,238 for the treatment of this type of aggressive cancer. In April 2017 a fluorinated analog of sorafenib, known as regorafenib has been approved for the treatment of sorafenib-resistant HCC patients,239,240 but no drugs able to overcome both types of drug-resistance are known today.241

In literature it has been reported that the upregulation of class I HDACs is a common factor in HCC contributing to spread, aggressiveness and increased mortality.238241 Recently, Cheng et al.242 reported a series of purine-based hydroxamates able to interact with both HDAC and PI3K/Akt/mTOR signaling pathway for the treatment of HCC. As highlighted through previous structure-activity relationship (SAR) studies focused on the PI3K/mTOR inhibitor 54 (Figure 6),243 the introduction of a reasonably sized substituent, specifically of a hydroxamate portion at the N9-position, resulted well tolerated reaching the desired dual inhibitory capability. Among the several analogues obtained through different manipulations at the positions C2, C6, N7, and N9 of the bicyclic system, the most potent compound 55 (Figure 6) showed lower IC50 values against both HDAC1 (IC50 = 1.1 nM) and PI3Kδ (IC50 = 37 nM) than the relative single targeting compounds [vorinostat 3: IC50 = 12 nM (HDAC1), 54: IC50 = 340 nM (PI3Kβ)]. Accordingly, 55 displayed higher or similar antiproliferative activity in MV4–11 leukemia cell lines [IC50 = 0.047 μM vs 1.46 μM (pictilisib 53) and 0.043 μM (vorinostat 3)] and in the hepatic carcinoma cell lines HuH-7 [IC50 = 0.48 μM vs 1.72 μM (pictilisib 53) and 1.96 μM (vorinostat 3)]. Compound 55 showed excellent oral anti-antitumor activity as a single agent in HCC and 4T1 mouse metastatic breast cancer models and these results, together with the good PK characteristics, suggested the potential use of 55 either as monotherapy or, preferably, in combination with well-known approved (immuno)drugs for treatment of non-Hodgkin lymphomas (i.e., diffuse large B-cell lymphoma) and subgroups of leukemia (i.e., AML). Furthermore, compound 55 proved to be safe as it did not inhibit hERG channel protein (IC50 >100 μM), and no significant inhibition of CPY3A4 (IC50 >10 μM) has been registered.242

Figure 6.

Figure 6.

Figure 6.

Structures of novel HDAC/TK inhibitors based on structures of PI3K, FGFR, CDK (ribociclib 59 and abemaciclib 61), VEGFR (pazopanib 63), JAK (ruxolitinib 66)/HSP90 (BEP800, 67) inhibitors. IC50 values are indicated, when available. The HDAC inhibitory portions are depicted in red. The HSP90 inhibitory portion is depicted in blue.

Fibroblast growth factor (FGF) family and their four receptor tyrosine kinases (FGFR1,2,3,4) cover an essential role in many physiologic processes such as tissue homeostasis and repair, embryogenesis, inflammation, and wound healing. However, the amplification of their expression was found in different kind of cancers such as breast, gastric, cervical, oral or bladder cancer.244252 Recently, the combination composed by HDAC6 and TK inhibitors was proven to be effective against breast cancer, suggesting the possible development of hybrid compounds able to interact with both targets.253,254 Compound 56 (IC50 FGFR1= 2.9 nM) (Figure 6) has been previously reported by Liu et al.255 as novel and selective second generation TKi, and resulted the starting point for the design of the first series of HDAC6/FGFR1 dual inhibitors.256 The 56/FGFR1 co-crystal structure revealed that the N-ethyl-4-phenylpiperazine moiety extended out of the solvent exposed region. Thus, the N-ethyl-4-phenylpiperazine portion was replaced with the Zn2+ binding group of the HDAC6 selective inhibitor nexturastat A 57 (IC50 HDAC6 = 6 nM, with >190-fold selectivity over other HDACs) (Figure 6),257 obtaining several potential hybrid molecules among which the most potent compound was 58 (IC50 HDAC6= 34 nM, 64% inhibition of FGFR1 at 1 μM) (Figure 6). Antiproliferative data in breast cancer MCF-7 cells collected with 58 (IC50 value = 9 μM) were comparable with those of vorinostat 3 (IC50 = 2.7 μM) and nexturastat A 57 (IC50 = 1.4 μM).

Cyclin-dependent kinases (CDKs) play a major role in modulating cell cycle regulation. In particular, CDK4 controls the G1 phase through the formation of a complex with cyclin D1. Overexpression of CDKs and in particular of CDK4, as well as dysregulation of the CDK4-cyclinD-Rb pathway, have been registered in cancer. CDK4 inhibitors (CDK4i) could induce arrest in G1 phase and in cell cycle progression controlling tumor growth.258264 The first generation of CDK4i showed poor specificity and severe toxic side effects that limited their use in clinical trials.265,266 Therefore, highly selective CDK4i were developed such as ribociclib and palbociclib, which entered clinical trials.267271 The combination of CDKi with the aromatase inhibitor letrozole resulted essential for the treatment of ER+/HER2 metastatic breast cancer, suggesting that their use as single agents may be insufficient.272,273 The synergistic effect resulting from the combination of HDACi and CDKi in melanoma and neuroblastoma274 induced Li et al.275 to develop hybrid molecules based on the pyrrolo[2,3-d]pyrimidin-2-ylamine scaffold of the CDK4/9 inhibitor ribociclib 59 (IC50 CDK4 = 13 nM). By the introduction of the anilide-polymethylene-hydroxamate motif (typical of vorinostat 3) at the C2-amino group of ribociclib 59, a series of novel, highly selective (in a kinase profiling assay against 375 kinases) and potent inhibitors of both CDK4/9 and HDAC1 has been obtained. Compound 60 (Figure 6) was the most effective of the series, with IC50 values of 8.8 (CDK4), 12 (CDK9), and 2.2 (HDAC1) nM. In cells, 60 induced cytotoxic effects and apoptosis in different cancer cell lines (liver Hep3B and HepG2, breast MDA-MB-231, 4T1, and T47D, and lung H460 and A549 cancer cell lines) through CDK4/9 and HDAC1 inhibition and phosphorylation of p53 in addition to G0/G1 (at low concentration) or G2/M (at high concentration) arrest, thus preventing cell differentiation and proliferation. Oral or intraperitoneal administration of 60 led to a potent tumor regression in mice bared-breast cancer 4T1 xenograft models. Additional studies showed that 60 downregulated the activity of CDK4 as well as the phosphorylation of protein Rb and significantly upregulated the acetylation state of H3 and the relative levels of antiapoptotic BCL-2 expression and cleaved-caspase-3.275

Huang and coworkers demonstrated that the activity of vorinostat 3 resulted synergistic with that of the CDK4/8 inhibitor abemaciclib 61 (Figure 6) against different solid tumors.276,277 X-ray co-crystal structure analysis of abemaciclib 61 and CDK4/CDK6 revealed that the 2-aminopyrimidine group which interacts with the protein hinge region was important for the 61 inhibitory activity. Therefore, a series of dual CDK/HDAC inhibitors was developed merging the pharmacophoric scaffold of abemaciclib 61 with that of vorinostat 3 or similar.278 Among the synthetized compounds, Roxyl-zhc-84 62 (Figure 6) showed the most potent antitumor activity against a broad spectrum of solid tumors in both in vivo and in vitro models, showing improved HDAC isoform selectivity profile compared to the pan-HDACi vorinostat 3. In cells, Roxyl-zhc-84 62 showed high cytotoxic activity (IC50 values in the 10−40 nM range in breast 4T1, MDA-MB-468 and MDA-MB-231 and ovarian SK-OV-3, OVCAR-5 and H450 cancer cell lines, 30 times lower respect to vorinostat 3. In addition, 62 was well-tolerated in in vivo assays (in both 4T1 and MDA-MB-468 mouse models), with no significant loss of body weight. At molecular level, Roxyl-zhc-84 62 reduced STAT3 and Janus kinase 1 (JAK1) phosphorylation markedly downregulating BCL-2, which is known to be a STAT3 target gene, thus sensitizing cells to HDACi activity. In addition, due to its HDAC inhibitory capability Roxyl-zhc-84 62 increased the expression level of the cell cycle regulator p21. Importantly, Roxyl-zhc-84 62 showed a stronger antitumor activity than the combination of abemaciclib 61 and vorinostat 3, or than that of JAK1i (filgotinib) and HDACi in both in vitro and in vivo models.278

The vascular endothelial growth factor (VEGF)/vascular endothelial growth factor receptor (VEGFR) signaling pathway is one of the most important targets for anti-angiogenesis and counts over ten small molecule inhibitors and monoclonal antibodies adopted for the treatment of solid tumors.279281 Pazopanib 63 (Figure 6) is a multiple VEGFR inhibitor approved by the FDA for the treatment of advanced soft tissue sarcoma and advanced renal cell carcinoma.282 Unfortunately, the clinical application of pazopanib 63 is limited by the insurgence of resistance and tumor relapse.283 Several phase I clinical studies supported the combination therapy focused on the co-administration of pazopanib 63 and different HDACi such as vorinostat 3 or abexinostat 11.284 Hence, based on these evidences, Zang et al. designed and synthetized a series of dual VEGFR/HDAC inhibitors by replacing the 63 benzene sulfonamide moiety with either a ortho-aminoanilide or hydroxamate group.285 The most potent analogs 64 and 65 (Figure 6) were evaluated for their antiproliferative activity against four hematological and seven solid tumor cell lines. In this assay, they showed potent arrest of cell proliferation while pazopanib 63 was not effective, highlighting the need for dual inhibition to obtain the desired effect with these compounds. In HUVECs tube formation assay and rat thoracic aorta rings assay, 64 and 65 strongly inhibited the microvessels outgrowth showing comparable potencies to pazopanib 63. Among the two hybrid compounds, only 64 progressed to in vivo PK studies showing 72% oral bioavailability in rats, while 65 suffered from metabolic liability due to the hydroxamate function. When tested in human colorectal adenocarcinoma (HT-29) xenograft model, 64 displayed considerable in vivo antitumor efficacies with no significant toxicity or body weight loss.286

Yao et al.287 developed a triple inhibitor able to hit distinct targets showing related functions in cancer development: JAK2 kinase, heat-shock protein 90 (HSP90), and HDACs. HSP90 is a molecular chaperone that assists a set of proteins (involved in signal transduction such as kinases and steroid hormone receptors as well as oncoproteins) during their maturation ensuring the correct folding.142 In addition, HSP90 resulted strongly upregulated in cancer inducing the activation of STAT3, an oncogenic inflammatory mediator.288,289 Since also JAK kinase activate STATs,288 and HDACs (in particular HDAC6) play an important role in modulating the HSP90 complex,290 the development of a triple JAK/HSP90/HDAC inhibitor able to interact simultaneously with all the three targets has a strong rationale. It is known that in the structure of the JAK inhibitor ruxolitinib 66 (Figure 6)291,292 (IC50 values = 3.3 (JAK1) and 2.8 (JAK2) nM) the C2 position of the 7-deazapurine, as well as the substituent at the C6 pyrazole group, can be modified without losing the inhibitory activity. Hence, these positions have been functionalized introducing the pharmacophoric elements typical of the HSP90 inhibitor BEP800 67 (IC50 HSP90β = 58 nM), that does not suffer modifications on the meta position of the phenyl ring for its inhibiting activity (Figure 6),293,294 and vorinostat 3, for which the possibility to functionalize the phenyl C4-position is well known.

Among the synthetized derivatives, 68 (Figure 6) showed the highest potency, inhibiting all the desired targets in the low micromolar range (IC50 values: 3.76 (JAK2), 20.2 (HSP90), and 6.3 (HDAC6) μM). Since this is a very recent and early state project, further chemical optimizations and biological analyses are required.287

4.2. Dual LSD1/kinase inhibitors

Osimertinib 69 (Figure 7) is an orally administrated, FDA-approved irreversible third generation EFGR inhibitor for the treatment of EGFR-mutated NSCLC. Recently, Li et al.295 reported that this well-known TKi is a dual EGFR/LSD1 inhibitor, displaying a IC50 against the KDM LSD1 of 3.98 μM due to its capability to act as a FAD competitor or FAD ejector (although not structurally similar to FAD). In NCI-H1975 cells, osimertinib 69 enhanced in a dose-dependent way the H3K4me1/2 levels as well as the expression of CD86, known to be biomarkers to evaluate LSD1 inhibition in cells, and arrested cells migration confirming its dual inhibition capability. Considering the important role that LSD1 covers during the onset of NSCLC, osimertinib 69 could be a new promising strategy for NSCLC treatment.

Figure 7.

Figure 7.

Structure of osimertinib 69, a dual EGFR/LSD1 inhibitor.

4.3. Dual BET/kinase inhibitors

Several kinase subtypes, as well as some members of the BET family, are implicated in the same pathways that characterize specific cancer pathologies. For example, FMS-related tyrosine kinase 3 (FLT3) and BRD4 are key targets in AML,296 and both BRD4 and JAK kinases play an important role in multiple myeloma models.168,297 The co-administration of a known TKi such as the FLT3 inhibitor quizartinib and JQ1 exhibited synergistic effects in cancer cell lines expressing the FLT3-ITD mutation (AML cells and AML steam/blast progenitor cells).298,299 SB284851-BT 70 (Figure 8) belongs to the 1,4,5-trisubstituted imidazoles mainly known as p38α kinase inhibitors. For these compounds, the ability to bind also BET bromodomain has been reported (IC50 values = 18 [BRDT (domain 1)], 3.7 [BRD4 (domain 1)], >250 [BRDT (domain 2)] μM, and not determined [BRD4 (domain 2)]).300302 Starting from SAR studies focused on 70 (Figure 8), in 2018 Divakaran and coworkers303 identified a new series of selective BRD4 (domain 1) over BRD4 (domain 2) (>55 fold) 1,4,5-trisubstituted imidazoles as dual p38α-bromodomain inhibitors. The properly substituted imidazole scaffold binds to bromodomain proteins in a similar manner respect to a histone acetyl group, taking part in a hydrogen bond interaction with conserved asparagine residues (N140) through its nitrogen atom. The selectivity is due to the flexibility of Lys141 typical of BRD4 (domain 1), compared to the steric limitation of the histidine typical of the C-terminal BET bromodomain and BRD4 (domain 2), that allows the displacement of structured water molecules (carried out by the fluorophenyl ring able to penetrate deeper into the binding pocket respect to other BETi), in addition to the effects induced by the imidazole ring itself. The obtained compounds, in particular 71 (Figure 8), showed, in addition to a potent and selective inhibition for domain 1 of BRDT, BRD4, BRD2, and BRD3 as well as for p38α (IC50 values = 3.5 [BRDT (domain 1)], 1.7 [BRD4 (domain 1)], 29 [BRD2 (domain 1)], 11 [BRD3 (domain 1)], >250 [BRDT (domain 2)], >250 [BRD4 (domain 2)], 67 [BRD2 (domain 2)], >100 [BRD3 (domain 2)] μM; kd p38α = 0.47 nM), the capability to efficiently bind BRD4 in different cell lines, and an excellent cell permeability. The A459 lung cancer cell line showed improved sensibility to the treatment, due to the capability of these molecules to target both p38α and BRD4 (demonstrated by decreased cytokine secretion). In contrast, the effect on the multiple myeloma MM.1S cell line resulted largely BRD4-driven, due to the prevalent sensibility of this cell line to the modulation of c-Myc expression, a characteristic hallmark of BETi. Early evidences also showed that such cooperative effects are the consequence not only of the dual inhibitory activity of these molecules, but also of the selectivity of their action among the domain 1 of BET proteins.303

Figure 8.

Figure 8.

Structure of dual BET/TK inhibitors.

4.4. Hybrids compounds able to simultaneously inhibit epigenetic and not epigenetic/not kinase targets

Some dual compounds have been designed by coupling the HDACi pharmacophore group with structures of inhibitors of various enzymes, apart from kinases. Two main factors have been identified at the base of the onset of acute promyelocytic leukemia (APL): the fusion of the RARα protein with the promyelocytic leukemia gene (PML), and the presence of blasts blocked at the promyelocytic stage. The effectiveness of the treatment based on ATRA is due to the capability of this drug to release the HDAC-containing repressive complex bound to PML-RARα with the recruitment of the multi-subunit HAT complex on specific DNA responsive elements (RARE). Based on the evidence that ATRA induces differentiation and cell death in APL cell lines, and assuming the remarkable context-specific physical proximity of a retinoid and HDACi-binding protein in the repressive PML-RARα-HDAC complex, we pioneered the epi-polypharmacology field with design and synthesis of the hydroxamate 72304 (Figure 9) and the benzamide 73305 (MC2392, Figure 9), strictly related to ATRA, to be tested in leukemia. While 72 did not show any significant activity in leukemia cells, 73, despite its weak retinoid activity and general no HDAC inhibition in vitro and in vivo, was able to induce changes in H3 acetylation at a small subset of PML-RARα–binding sites, to alter expression of stress-responsive and apoptotic genes, and to induce massive cell death specifically in PML-RARα expressing APL cells, while solid and other leukemic tumors were not affected.305

Figure 9.

Figure 9.

Figure 9.

Structures of ATRA-based HDACi, dual HDAC/IDO1 inhibitors, dual HDAC/NAMPT inhibitors, lefloxacin- and iso-combretastain A-4-based HDACi, dual HDAC/HSP90 inhibitors, ebselen-based HDACi, dual HDAC/PDE5A inhibitors, HDAC/TG2 inhibitors, dual HDACi/NO donor compound. IC50 values are indicated, where available. The HDAC inhibitory portions are depicted in red. The HSP90 inhibitory portion is depicted in blue.

The activation of T cells in order to obtain a native immune response through immune check point therapy is a very interesting strategy that has achieved significant advances in clinical trials for cancer therapy.306 Indoleamine-2,3-dioxygenase (IDO1) is an extrahepatic heme-containing dioxygenase able to catalyze the conversion of tryptophan (Trp) to N-formylkynurenine (NFK), that is then converted to L-kynurenine (Kyn) and different bioactive metabolites.307 The reaction in which IDO1 is involved is the rate-limiting step of the kynurenine pathway (KP).307,308 T lymphocytes proliferation is dependent on the Trp levels, so Trp depletion induces its inhibition. In addition, the metabolites deriving from KP can activate the aryl hydrocarbon receptor (AhR) inducing immune tolerance. Both of these conditions contribute to the immunosuppressed state at the base of tumor pathogenesis,309,310 and high levels of IDO1 (in both antigen presenting and tumor cells) are often correlated with reduced survival.311,312 IDO1 inhibitors (such as epacadostat, IC50 = 10 nM) eradicate and control cancer cells growth due to their capability to induce antitumor immune response, but a lot of preclinical studies showed that their activity is reduced when these molecules are used as single agents.311,313 Conversely, IDO1 inhibitors improved their efficacy when administered in combination with radiopharmaceutical drugs, cytotoxic anticancer agents, etc.311,313,314 Recent evidences showed that HDACi could be able to reverse immune suppression inducing an improvement in tumor recognition.315,316 Fang et al. reported the first dual HDAC/IDO1 inhibitors, designed by insertion of the benzamide anti-HDAC moiety of entinostat 9 or mocetinostat 10 into the structure of epacadostat 74, through replacement in its molecule of the aminoethyl-sulfamide function, exposed to the solvent in the binding with the enzyme.317,318 The dual agent 75 (Figure 9) inhibited both HDACs and IDO1 at nanomolar range, decreased proliferation, and induced G2/M cell cycle arrest and apoptosis in HCT-116 colon, LLC lewis lung, and A549 lung cancer cell lines. Moreover, 75 showed an acceptable PK profile after oral administration reducing plasma Kyn levels over 8 h with an interesting anticancer activity and low toxicity in vivo in a murine LLC tumor model.318

Considering the high request of NAD+ by cancer cells, the inhibition of nicotinamide phosphoribosyltransferase (NAMPT), the rate-limiting enzyme in NAD+ biosynthesis, has been considered a potential strategy to hit cancer.319,320 The NAMPT inhibitor (NAMPTi) FK866, which entered clinical trials but whose development has been limited by dose-limiting thrombocytopenia,321324 has been reported to synergistically improve the activity of HDACi. Starting from these evidences, Chen et al.325 developed dual NAMPT/HDAC inhibitors, by connecting the N-(3-(1H-imidazol-1-yl)propyl)-5-thiazolecarboxamide moiety contained in the novel NAMPT inhibitor 76 (Figure 9)326 with an amidopolimethylene-hydroxamate tail, able to inhibit HDACs.327 Among them, 77 was the most potent compound in term of enzymatic and cellular activity (IC50 values: 15 (NAMPT), 2 (HDAC1), 6.8 (HDAC2), 0.67 (HDAC3), 12000 (HDAC4), 0.64 (HDAC6) and 204 (HDAC8) nM) (Figure 9). Specifically, 77 was slightly less potent than the NAMPTi FK866 and strongly inhibited better than vorinostat 3 all the tested isoforms of HDACs except HDAC4. Accordingly, 77 markedly increased the acetylation levels of histones H3 and H4 in a dose-dependent manner, and efficiently decreased the cellular level of NAD+ after 24 h of incubation in human colon cancer cells HCT116. In cells, 77 induced moderate cytotoxicity against HCT116, HepG2 liver cancer, and A549 lung cancer cell lines. However, when evaluated in vivo in HCT116 tumor xenografts in nude mice through intraperitoneal infusion at 25 mg/kg twice a day for 21 consecutive days, 77 determined a strong reduction of tumor growth, and higher or comparable antineoplastic activity (TGI = 42%, T/C = 58%) respect to FK866 (TGI = 39%, T/C = 61%) and vorinostat 3 (TGI = 33%, T/C = 67%), without significant body weight loss.325

Recent evidences highlighted the capability of fluoroquinolones to carry effective broad-spectrum antitumor activity against different types of cancer such as SLCL, colorectal carcinoma, and bladder cancer.328,329 In particular, levofloxacin 78 (Figure 9) showed antiproliferative activity against a different kinds of tumors.330 In addition, many fluoroquinolones inhibited tubulin polymerization exhibiting selective activity against some tumor cell types330,331 and favorable PK characteristics.332

Wang et al.333 developed a series of hybrid HDAC/tubulin polymerization inhibitors based on the levofloxacin 78 scaffold. Across this series, 79 (Figure 9) was more potent than vorinostat 3 against HDAC1 and HDAC6, and more effective than levofloxacin 78 and colchicine in inhibition of tubulin polymerization (ITP). Furthermore, 79 exhibited higher antitumor activity and cell type selectivity respect to the corresponding single target inhibitors when tested against different kinds of tumor cell lines, and in particular against MCF-7 breast cancer cells (IC50 79 = 0.3 μM, IC50 vorinostat 3 = 4.4 μM, IC50 levofloxacin 78 = 64.2 μM).

Recently, the combination of vorinostat 3 plus vincristine has been shown to induce synergistic effects in both in vitro and in vivo models, suggesting that vorinostat 3 could alter microtubule dynamics through HDAC inhibition.334 Thus, another series of dual HDAC/tubulin polymerization inhibitors has been successfully developed by Lamaa and coworkers by merging the structure of belinostat 5 with the 1,1-diarylethylene scaffold of iso-combretastatin A-4 80 (IC50 for ITP = 2.0 μM) (Figure 9).335 Among these compounds, 81 (Figure 9) showed the best antiproliferative results (IC50 = 0.4–2 nM) in different tumor cell lines (K562, PC3, U87, CA-4, HT-29, BXPC3), inducing a strong inhibition of both HDAC8 (IC50 = 340 nM) and tubulin polymerization (IC50 = 1.6 μM) as well as cancer cells cycle arrest at the G2/M phase by disruption of the microtubule organization.335

As previously reported, recent evidences highlighted the synergism deriving from the combination of HDAC6-selective and HSP90 inhibitors. The 4-isopropylresorcinol scaffold is known to fit into the hydrophobic and hydrophilic region of HSP90 playing a key role as binder at the ATP binding site.336338 Merging such evidence with those that indole and non-planar indoline/tetrahydroquinoline-based scaffold act as a selective HDAC isoform inhibitor, a series of 1-aroylindoline-hydroxamic acids differing in the length of the linker have been developed by Oiha et al. as dual HDAC/HSP90 inhibitors.339 Among these compounds, 82 (six methylene linker) (Figure 9) was the most potent with IC50 values of 46.3 nM for HSP90 and 1.15 nM for HDAC6 (with a 113, 139 and 246-fold selectivity over HDAC1, HDAC3 and HDAC8 isoforms, respectively). In colorectal HCT116, lung A549, EGFR T90M mutant lung H1975 and leukemia HL60 cell lines, 82 displayed cytotoxic effects with GI50 = 1.04–1.61 μM, and the maximum cytotoxic effect was registered towards H1975 NSCLC cell lines with IC50 = 0.26 μM.

As reported in literature, HDAC6 and class I HDACs cover an essential role in Alzheimer disease (AD), since their overexpression were detected in the cortex and hippocampus of AD patients.2630 It has been postulated that at the basis of the onset and progression of AD there is the aberrant aggregation of the insoluble hyper-phosphorylated Tau protein (pTau). In particular, the acetylation of the KXGS motif of this protein (at the level of Lys280) prevents tau aggregation as a result of the inhibition of the phosphorylation process. Recent evidences showed that HDAC6 is responsible for deacetylation of these residues, therefore blood brain barrier (BBB)-permeable HDAC6 inhibitors could be very useful for the enhancement of this acetylation so preventing tau aggregation.3133 Also the oxidative stress contributed to the onset of different neurodegenerative disorders including AD.34 HDAC6 plays a central role also in redox regulation and in cellular stress response: in fact, the overexpression of HDAC6 is responsible for ROS generation in response to the upregulation of the expression of NADPH oxidase as well as pro-inflammatory cytokines and α-tubulin deacetylation.35 Selenium is a mineral nutrient physiologically present in essential traces in our organism, providing protection from free radicals. Its levels decrease during aging, covering different possible roles in AD.36,37 Starting from these evidences, Hu et al.38 developed a series of compounds based on the structures of vorinostat 3 and of the antioxidant and anti-inflammatory agent ebselen 83 (Figure 9).39 Among the synthetized compounds, 84 (Figure 9) showed the highest anti-HDAC6 activity (IC50 = 0.037 μM), joined to comparable or better glutathione peroxidase-like activity than ebselen 83, providing high free oxygen radical scavenging capability (ORAC = 2.2) in addition to marked protective activity against ROS accumulation in Rat adrenal phaeochromocytoma PC12 cells.

The inhibition of phosphodiesterase 5 (PDE5) leads to an enrichment in the available concentration of the second messenger cyclic guanosine monophosphate (cGMP). It resulted critical for neuronal signaling, promoting memory-related gene transcription through the activation of the cAMP response element (CREB) and the consequent recruitment of CREB binding proteins (showing HAT activity) due to the increased phosphorylation state. Several studies reported a level of cerebrospinal fluid cGMP lower in patients affected by AD than in age-matched controls.340 Clinical evidences of enhanced synaptic plasticity and memory improvement following PDE5 inhibitors administration (in particular, by using compounds able to cross the BBB such as sildenafil, tadalafil and vardenafil) in AD patients were collected during the last years.341343 Combinations of tadalafil (IC50 PDE5A = 9.4 nM), sildenafil (IC50 PDE5A = 8.5 nM) (Figure 9)344 or vardenafil (IC50 PDE5A = 0.89 nM),345 and vorinostat 3 were pre-clinically validated in a mouse model of AD,346 showing a synergistic effect for increased levels of histone acetylation via recruitment of CBP, similarly to what observed by the treatment with previously discovered hybrid PDE5/pan-HDAC inhibitors in mice.347 The effect of the HDAC6-selective inhibitor tubastatin A in combination with PDE5Ai has been evaluated against SH-SY5Y neuroblastoma cell lines, with the aim to highlight the role of this specific HDAC isoform in such disease. A marked increase of H3K9 acetylation has been registered compared to cells treated with tubastatin A alone.348 From these findings, Rabal et al.348 recently identified new sildenafil 85-based dual PDE5A/HDAC6 inhibitors. Among them, 86 (Figure 9) displayed the best enzyme inhibitory activity (IC50 HDAC6 = 15 nM; IC50 PDE5A = 11 nM), was able to cross the BBB and to achieve its functional response, reducing the levels of AD-related markers such as pTau and human amyloid precursor protein in Tg2576 neurons, and increasing the markers of the physiological neuron response in mouse hippocampus following the administration of 40 mg/kg dose. However, this treatment did not induce a significant improvement in memory after 2 weeks.348

Both HDACs and transglutaminases (TGs) play a pivotal role in neurodegeneration, and several evidences showed their induction in the presence of toxic stimuli. Moreover, the inhibition of both targets induces a protective effect in different neurodegenerative diseases restoring the mRNA and protein levels of brain-derived neurotrophic factor (BDNF), a well-known neurotrophin implicated in neuronal vitality, development and synaptic plasticity.349 In particular, the most expressed and studied TG isoform is TG2, which is mostly involved in AD and Huntington’s disease (HD). A drug combination based on the TG2 inhibitor (TGi) B001 and the HDACi sodium butyrate, demonstrated that this co-treatment synergistically protects cells from glutamate toxic effects at the tested concentration (% viable cells: 28 (glutamate alone), 37.5 (glutamate plus B001), 44.3 (glutamate plus sodium butyrate), 67.5 (glutamate plus B001 and sodium butyrate). On the same line, the combination based on cysteamine and vorinostat 3 markedly enhanced cell survival (from 28% for cystamine alone and 36.7% for vorinostat alone to 59.8% for combination). To date, only a few TG2i have been reported in literature and, in particular, some 3-(substituted cinnamoyl)pyridines including compound 87 (IC50 TG2 = 21 μM) (Figure 9) are worthy of note. In the 87 structure, the nitro group at the para position of the cinnamoyl moiety could be replaced by more complex substituents without loss of enzyme inhibitory activity. For the first time, Basso et al.350 developed a series of dual TG2/HDAC inhibitors based on the central scaffold of 87. To support previously collected data about the synergistic effect induced by co-treatment of TG2i and HDACi, the best molecule among those synthetized, compound 88 (IC50 TG2 = 13.3 μM, IC50 HDAC1 = 3.38 μM, IC50 HDAC6 = 4.10 μM) (Figure 9) was evaluated in cortical neurons to test its capability to protect neurons from death. Cells were treated with increasing concentration of 88 without further toxic stimuli. Compound 88 did not show any toxic effects at the highest tested concentration (50 μM), showing a neuroprotective effect at concentration as low as 6.3 μM, where 80% of the neurons were preserved, and at higher concentrations a complete protective effect was achieved (at 25 μM: 95% viable cells, at 50 μM: 100% viable cells).350

Dystrophic mice treated with either NO donors or selective HDACi showed partial rescue of the disease, suggesting to develop dual NO donors/HDACi agents. This goal was reached by introducing the dinitrate NO donor moiety at the C6-position of the pyridine ring of entinostat 9. Compound 89 (Figure 9) was the most potent among the series of hybrid compounds synthetized by Atlante et al.,351 showing inhibitory values against HDACs comparable with those of entinostat 9. In particular, the highest inhibitory activity of 89 was towards HDAC2 (IC50 = 0.39 μM), that is inhibited also by S-nitrosylation.352 Compound 89 furnished also the best results in terms of NO donor activity according to cytometry and fluorescence-activated cell sorting (FACS) analysis, skeletal muscle differentiation, ex vivo (from male Wistar rats) aortic vasodilation, and formation of large myotubes characterized by a high number of multinucleated fibers. Compound 89 has been tested for its cytotoxic effect in immortalized mouse myoblast C2C12 cells and immortalized human keratinocyte HaCaT cells, showing lower cytotoxicity than entinostat 9.351

4.5. PROteolysis TArgeting Chimera (PROTAC) approach applied to epigenetic targets

Thalidomide 90 (Figure 10) was originally used as a sedative and antiemetic drug even for gravidic nausea before being banned due to the teratogenic side effects of one of its two enantiomers. Currently, thalidomide 90 is mainly used for the treatment of certain cancers (multiple myeloma) due to its immunomodulatory properties. The mechanism of action of 90 was fully characterized only recently, demonstrating that its primary target is cereblon, a component of the cullin-dependent ubiquitin ligase responsible for the proteasome-mediated degradation of proteins.353355 Compound 91 (Figure 10) is the result of the conjugation of thalidomide 90 with the pan-BET inhibitor JQ1. Thanks to the capability of its thalidomide portion to recruit cereblon, 91 selectively increased the degradation of BRD2, BRD3 and BRD4 inducing higher apoptosis than JQ1 in leukemia cell lines.356 Compound 92 (Figure 10) displayed a longer linker between thalidomide and JQ1 portions and led to fast, efficient, and prolonged degradation of BRD4 in all the tested Burkitt’s lymphoma cell lines.357

Figure 10.

Figure 10.

Structures of BET/PROTAC compounds.

Many other examples of PROTACs built with epigenetic ligands have been reviewed very recently.358

4.6. Dual Inhibitors Hitting Two or More Epigenetic Targets

All the examples discussed so far are the results of conjugation of two active substances in which only one portion has an epigenetic target. Here, we are going to consider series of multifunctional compounds in which both fragments interact with two (or more) epigenetic targets.

LSD1 and JmjC KDM4A/C enzymes play a key role in human prostate cancer because they are co-expressed and co-localized with AR.359 We designed and synthesized some LSD1/JmjC hybrid molecules360 by coupling the skeleton of the well-known LSD1 inhibitor tranylcypromine 24, to the carboxylate group of either the 4-carboxy-4′-carbomethoxy-2,2′-bipyridine 93361 (Figure 11) or the 5-carboxy-8-hydroxyquinoline (5-carboxy-8-HQ) 94362 (Figure 11), two 2-oxoglutarate competitive scaffolds developed as JmjC inhibitors. In the two hybrid compounds 95 and 96 (Figure 11), the common tranylcypromine portion makes a covalent adduct with FAD to inhibit LSD1, while the carboxylic acid portion released from the bipyridyl prodrug 95, as well as the hydroxyquinoline group of 96, can chelate the JmjC protein-bound iron leading to enzyme inhibition. In cellular assays, 95 and 96 increased H3K4 and H3K9 methylation levels producing growth arrest and apoptosis in LNCaP prostate and HCT116 colon cancer cells. Most importantly, when tested in mesenchymal progenitor (MePR) non-cancer cells, 95 and 96 induced little or no apoptosis, showing a characteristic cancer-selective action.

Figure 11.

Figure 11.

Structure of pan-KDMs inhibitors, dual G9a/LSD1 inhibitors, dual G9a/DNMT inhibitors, dual G9a/HDAC inhibitors, dual G9a/EZH2 inhibitors, multi-target inhibitors, dual HAT/EZH2, dual DNMT/HDAC inhibitors, DNMT/HDAC/SIRT inhibitor, dual DNMT/HDAC, dual LSD1/HDAC inhibitors and dual BRD4/HDAC inhibitors. IC50 values are indicated, where available. The HDAC inhibitory portion is depicted in red.

The KMT G9a is a H3K9 methyltransferase which leads to gene silencing and is overexpressed in different types of tumors. Recent studies demonstrated that the inhibition of G9a decreases cancer cell proliferation, blocks metastases and delays tumor progression.363,364 Our researches on properly substituted quinazoline structures allowed us to identify novel H3K9 methyltransferase/demethylase as well as DNMT3A inhibitors bearing a quinazoline structure.365,366 In particular, structural manipulations of BIX01294 97367 (Figure 11) and UNC0638 98368 (Figure 11), two G9a inhibitors largely known in literature, led us to the identification of MC3774 99 (Figure 11), a Lys-mimicking derivative displaying dual G9a/LSD1 inhibition.369 MC3774 99 is characterized by a dimethylaminopropylamine chain at C2, a 4-aminobenzylpiperidine moiety at C4 and two methoxy groups at C6 and C7 positions of the quinazoline scaffold. We designed 99 supposing that the C2-propanamine moiety could mimic the H3K4me2 skeleton of the substrate bound within the LSD1 active site. Surprisingly, the compound does not enter into the LSD1 catalytic tunnel but binds the enzyme placing five copies of its molecule in a stacked way obstructing the entrance of the LSD1 active site. The orientation of the molecules can be “face to face” or “head to tail”, and in both modes, they interact with a cluster of negatively charged amino acidic residues.369 This kind of non-covalent and reversible inhibition is typical for the interaction with LSD1, because 99 does not inhibit G9a using the same stacking mode. In MV4–11 leukemia cells, 99 showed anti-proliferative activity with IC50 = 0.89 μM, comparable with the potency observed against LSD1 and G9a at the enzyme level.369

G9a is known to physically interact with DNMT1 to coordinate histone and DNA methylation during cell division,370 promoting repression of target gene expression. From these findings, a global reduction of DNA as well as H3K9 methylation can lead to reactivation of tumor suppressor genes, inhibition of cancer cell proliferation, arrest of cell cycle and induction of apoptosis.371,372 The crystal structure of the UNC0638/G9a complex revealed that the C4 aminopiperidine moiety of the ligand as well as the C7 pyrrolidinopropyloxy chain are important for specific interactions in the active site of the enzyme, and that, most importantly, while the protonated (at physiological pH) N1 of the quinazoline interacts, through a salt bridge, with Asp1088, the quinazoline N3 did not take part in any direct and crucial interaction with G9a.368 Thus, Rabal et al. decided to design and synthetize a quinoline analog of UNC0638 98, compound 100 (Figure 11), to be tested against both DNMT1 and G9a through time-resolved fluorescence resonance energy transfer (TR-FRET) technique. Competition experiments confirmed that 100 displayed substrate-competitive inhibition of both targets, with global reduction in H3K9me2 and 5-methyl cytosine (5mC) levels.373,374 Unfortunately, 100 did not show any anti-proliferative response against ALL CEMO-1 (GI50 >10 μM), DCLBL OCI-Ly3 and OCI-Ly10 (0% inhibition at 1 μM) and AML MV4–11 (GI50 >10 μM) cell lines. SAR studies focused on the replacement of the C2 substituent and on the iso-propyl to methyl change at the piperidine moiety led to CM-272 101 (Figure 11), with improved potency against DNMT1 and displaying antiproliferative activities in cells (GI50 <1 μM against the above reported cell lines) joined to low CYP450 isoforms inhibition, good stability in human microsomes, acceptable exposure and good half-life after intravenously administration. Importantly, treatment with 101 induced anticancer effects in vivo against ALL, AML and diffuse large B-cell lymphoma (DLBCL) xenogeneic models.375

Hepatocellular, pulmonary, prostatic carcinoma and leukemia are associated with a global histone deacetylation state and with the double methylation of H3K9.376378 From these evidences, it emerged the need to design inhibitors capable to interact with both G9a and HDAC enzymes. Merging the quinazoline scaffold typical of BIX01294 97 with polymethylene hydroxamate chains allowed the identification of dual G9a/HDAC inhibitors. Among them, 102 (Figure 11) showed low toxicity, good physico-chemical properties, and antiproliferative activity in the micromolar range in HeLa and K562 cell lines.379,380

As previously discussed, G9a plays a key role in melanoma, breast and lung cancer mediating cellular growth and regulating cell cycle and metabolism pathways.381 Another KMT, the H3K27 methyltransferase EZH2, has a repressive role in gene expression, and has been found mutated in B-cell lymphoma and upregulated in breast and prostate carcinoma.382 Simultaneous inhibition of G9a and EZH2 exerted synergistic effects in MDA-MB-231 TNBC.383 Compounds like 103 (Figure 11), strictly related to BIX-01294 97, showed dual G9a/EZH2 inhibition with decrease of both H3K9me2 and H3K27me3 levels and upregulation of genes typically repressed by EZH2 (JMJD3, KRT17, FBX032). Importantly, these dual G9a/EZH2 inhibitors displayed growth arrest in a panel of breast cancer and lymphoma cell lines with low to sub-micromolar IC50 values.383

Curcumin 104 (Figure 11) exerts its powerful anticancer activities with several mechanisms of action, some of them involving epigenetic targets, such as the regulation of HDAC, HAT, DNMT1 and miRNA activities.384 Structurally related to curcumin 104, some unsaturated (di)ketones were reported as epigenetic multiple ligands.385 With appropriate substitutions at the phenyl rings, these compounds (such as 105-108, Figure 11) could act as simultaneous inhibitors of SIRT, HAT, KMT and PRMT.385 When tested in human leukemia U937 cells, 105–108 displayed higher apoptosis, and 105 and 106 increased the cell differentiation percentage respect to the corresponding single-target inhibitors.385

A screening of related chromatin modulating compounds allowed Petraglia et al. to identify MC2884 109 (Figure 11), a novel dual HAT/EZH2 micromolar inhibitor (IC50 p300 = 45 μM, % inhibition of EZH2 40% and 80% at 5–25 μM, respectively) able to induce the decrease of H3K27me3, H3K9/14ac and H3K27ac levels as well as the induction of caspase-dependent apoptosis associated with a strong, dose-dependent and specific anticancer activity in several cancer cell lines (in particular leukemia cells) and a good safety profile (mesenchymal progenitor model MePR2B and endometrial stromal cells were non-responsive to treatment). In addition, cellular analysis revealed that MC2884 109 displayed a synergistic effect on cancer cells death, while no effect was registered when the p300 inhibitor C646 or the EZH2 inhibitor GSK126 were used alone or in combination.

The marked apoptotic effect is due to its capability to induce the modulation of specific mitochondrial pathways of cell death, in addition to the reduction of acetylation of gene promoter regions, downregulating the expression of different anti-apoptotic factors such as BCLX(L) and BCL2. The high anticancer activity of MC2884 109 against hematological, solid and aggressive models of cancer (TET2−/− and p53−/−) and in vivo xenograft models (colon cancer, AML and APL), as well as its massive apoptosis induced in ex vivo human primary leukemia blasts with poor prognosis in vivo, affords a new model of personalized precision medicine by coupling the use of epi-drugs with epigenome analysis.

UVI5008 111 (Figure 11) was obtained by chemical substitution of the aromatic portion of the dimeric natural product psammaplin A 110 (Figure 11), known as HDACs (in particular HDAC1–4) and DNMT inhibitor.386 UVI5008 111 showed inhibition on the same epi-targets as 110 plus sirtuins,387 and evoked apoptotic effects in several human cancers such as leukemia, osteosarcoma, melanoma, breast, prostate and colon carcinoma.

DNMT and HDAC are known to work synergistically to induce tumor suppressor gene silencing. This cross talk between DNA methylation and histone (de)acetylation is essential for the control of gene expression, and its dysregulation contributes to the development of cancer.388390 Interesting evidences showed that the simultaneous inhibition of HDAC and DNMT had a synergistic effect in some types of cancers, encouraging the design of dual inhibitors for these two classes of enzymes.388,389,391393 A series of inhibitors based on the structure of the compound NSC-319745394 have been synthesized, by changing the substitutions at the benzamide portion and by converting the carboxylic acid function into the corresponding hydroxamate. Compound 112 (Figure 11) displayed higher DNMT1 inhibition than NSC-319745 and inhibited HDAC1 and HDAC6 at nanomolar levels. In cellular assays, 112 (Figure 11) showed effective cytotoxicity against human K562 and U937 cancer cells, with low micromolar IC50 values. In particular, in U937 cells, 112 increased the H3K9 and H4K8 acetylation levels, upregulated p16 expression by p16 CpG islands demethylation, and induced remarkable apoptosis.395

Since HDACs and LSD1 are overexpressed in several types of human cancers contributing to the silencing of tumor suppressor genes, also the simultaneous inhibition of these two enzymes can furnish synergistic activity against cancer cell growth, migration and invasion. The LSD1 inhibitors ORY-1001 25 and GSK2879552 26, currently in clinical trials for the treatment of leukemia, demonstrated that the amino group of tranylcypromine 24 can be substituted without loss of LSD1 inhibiting activity. Thus, compound 113 (Figure 11) was designed, prepared and tested by introducing the vorinostat 3 structure into the amino group of tranylcypromine 24. Compound 113 inhibited LSD1 at micromolar and HDAC1/2 at nanomolar levels and showed antiproliferative activities higher than vorinostat 3 against A549 lung, MCF-7 breast, MGC-803 gastric, and SW-620 colorectal cancer cells. Target engagement was demonstrated in MGC-803 cells, where 113 increased methyl-H3K4/methyl-H3K9 together with acetyl-H3 levels, decreased the mitochondrial membrane potential, and induced apoptosis.396

Our contribution in this field was to prepare novel LSD1/HDAC dual inhibitors by introduction of the HDAC inhibiting hydroxamate or benzamide side chain at the phenyl ring of tranylcypromine 24, where we previously inserted various amide or amino acid substituents that highly improved the LSD1 inhibitory potency of 24 itself.30,397401 Size exclusion chromatography revealed stable association of the HDAC1:LSD1:CoREST ternary complex with 1:1:1 stoichiometry. The most potent compound, 114 (Figure 11), bound the CoREST ternary complex showing both LSD1 and HDAC1 inhibition at submicromolar level, and exhibited higher antiproliferative activity than entinostat 9 against cutaneous squamous carcinoma and several melanoma cell lines, being less toxic than 9 towards keratinocytes and melanocytes. Importantly, 114 slowed down tumor growth in a melanoma mouse xenograft model.402

Milelli et al.403 developed a series of polyamine-based hybrid compounds, from which emerged 115 (Figure 11), showing interesting inhibitory activity against LSD1 (IC50 LSD1-CoREST3= 3.8 μM) and HDAC1 (Ki HDAC1-CoREST3= 42.5 nM) in vitro as well as in cellular assays. In 115, the two anti-HDAC and −LSD1 pharmacophoric portions were connected using a polyamine-type chain, since the polyamine moiety can interact with both LSD1 and HDAC enzymes404406 due to protonated nitrogen atoms allowing electrostatic interactions with negatively charge amino acid residues.407,408 Among the different polyamine linkers used, the spermine (3-4-3), typical of 115, produced the best enzymatic results. When tested against MCF7 breast cancer cell line, 115 resulted more effective (at high concentration) than vorinostat 3 in inducing cytotoxicity (68.6% vs 56.6% of dead cells).403

Design and synthesis of dual inhibitors of BRDs and HDACs arises from the evidence that both these families of proteins take part in the acetylation process, and the independent inhibition of each single target is connected with similar genetic and biological effects.409,410 The 9H-purine scaffold has been reported to bind with different strength the BRD proteins,411 and some N6-benzoyladenine derivatives have been described as BRD4 inhibitors.412 In these last compounds, the N6-benzamido group mimics the acetyl-lysine substrate in the binding with BRD4, and electron-donating groups at the C2’-benzamide position are preferred to obtain high inhibition. Thus, compound 116 (Figure 11) was designed, among others, as a dual BRD4/HDAC inhibitor by incorporating the polymethylene hydroxamate chain typical of vorinostat 3 into the C2’-benzamide position of the N6-benzoyladenine. In HL-60 cells, 116 arrested cell proliferation, induced apoptosis, enhanced the ATRA-induced differentiation and inhibited c-MYC production. Importantly, 116 exhibited growth arrest in cells resistant to BRD4 inhibitors.413 The 3,5-dimethylisoxazole is another substructure able to mimic acetyl-lysine in the binding with BRDs.414 The introduction of the N-hydroxyeptanamide substituent at the 3-benzoyl-5-(3,5-dimethylisoxazol-4-yl)phenoxy moiety of a BRD4 inhibitor furnished to the new molecule 117 (Figure 11) dual activity against both BRD4 and HDAC1 at submicromolar level. Compound 117 showed antiproliferative activities against AML, MV4–11, and chronic myelogenous leukemia (CML) K562 cell lines. Docking studies performed on 117 confirmed what previously explored by X-ray studies on the 3,5-dimethylisoxazole, revealing an interesting complementarity of the compound with the binding pocket of BRD4, and showing a binding into the HDAC1 pocket similar as that of vorinostat 3.415

5. Conclusions and future perspectives

MMT, MCM and MTDLs seem to be very favorable approaches for the treatment of complex diseases. Indeed, to date, it is well known that quite often single agent-based therapy characterized by a selective mechanism of action due to the engagement of a specific target is not enough, especially in multifactorial diseases. Despite the success of MMT at pre-clinical and clinical stage, the potential formulation problems, drug-drug interactions and the consequence toxicity lead us and others to hypothesize that the MTDL approach could be, even if not always, a true promise of polypharmacology.

In fact, the potential development of drugs that can be defined as “intelligent” or rather able to induce the inhibition or the degradation (as it happens for PROTACs) of carefully chosen targets, could turn out to be the true frontier of epigenetic and non-epigenetic research. Specifically, the role of the epigenetic cross-talk mechanisms associated with different types of dysregulation is well-accepted in different complex diseases such as cancer, neurodegenerative and metabolic disorders. Therefore, despite the frequent difficulty in structural optimization in order to effectively merge the pharmacophoric portions that characterized the entities of interest, the development of hybrid drugs able to simultaneously hit two or more targets in which one (or more) is (are) epigenetic and the other(s) belongs to a non-epigenetic category(ies) results very promising, as seen for dual HDAC/TK inhibitors, for which the two compounds CUDC-101 88 and CUDC-907 89 have already entered clinical trials. In addition, another important challenge which can be won through the polypharmacology approach is to extend the scope of epigenetic drugs beyond the limited spectrum of hematologic cancers, the only ones for which the epi-drugs have been approved so far.

ACKNOWLEDGMENTS

This work was supported by PRIN 2015 (prot. 20152TE5PK) (A.M.), AIRC 2016 (n. 19162) (A.M.), Ricerca Finalizzata PE–2013–02355271 (A.M.), and NIH (n. R01GM114306) (A.M.) funds. A special thanks to Dr. Giulia Stazi for the critical reading of the whole manuscript.

ABBREVATIONS USED

AD

Alzheimer disease

AhR

aryl hydrocarbon receptor

AKT

protein kinase B

ALL

acute lymphoblastic leukemia

AML

acute myeloid leukemia

APL

acute promyelocytic leukemia

AR

androgen receptor

ATC

anaplastic thyroid cancer

ATRA

all-trans retinoic acid

BBB

blood brain barrier

BCL2

B-cell lymphoma 2

BCL2i

B-cell lymphoma 2 inhibitors

BDNF

brain derived neurotropic factor

BET

bromo and extra terminal

BRD

bromodomain

CDCK4i

CDK4 inhibitors

CDCKS

cyclin-dependent kinases

cGMP

cyclic guanosine monophosphate

CLL

chronic lymphocytic leukemia

CREB

cAMP response element

CRPC

castration-resistant prostate cancer

CTA

cancer testis antigen

CTCL

cutaneous T-cell lymphoma

DNMTi

DNMT inhibitors

DNMTs

DNA methyltransferases

DOT1L

disruptor of telomeric silencing 1-like

EGFR

epidermal growth factor receptor

EMT

epithelial-mesenchymal transition

ER

estrogen receptor

EZH2

enhancer of zeste homolog 2

FACS

fluorescence-activated cell sorting

FAD

flavin adenine dinucleotide

FDA

food and drug administration

FGF

fibroblast growth factor

FLT3

FMS-related tyrosine kinase 3

GO

gentuzumab ozagamicin

HATs

histone acetyltransferases

HCC

hepatocellular carcinoma

HDACi

HDAC inhibitors

HDACs

histone deacetylase

HER2

human epidermal growth factor receptor 2

HER3

human epidermal growth factor receptor 3

HMT

histone methyltransferases

8-HQ

8-hydroxyquinoline

HSP90

heat-shock protein 90

IDO1

indoleamine-2,3-dioxigenase

ITP

inhibition of tubulin polymerization

JAK1

janus kinase 1

JmjC

Jumonji C

KDM

lysine demethylase

KLF4

kruppel like factor

KMTs

lysine methyltransferases

KP

kynurenine pathway

LSD1

lysine-specific histone demethylase 1

LSD2

lysine-specific histone demethylase 2

MAOs

monoamine oxidases

MAOi

MAO inhibitors

MAPK

mitogen activated protein kinase

MBD

methyl binding domains

MCM

multi-compound medication

MDS

myelodysplastic syndrome

MePR

mesenchymal progenitor

MET

hepatocyte growth factor receptor

MLL

mixed-lineage-leukemia

MM

multiple myeloma

MMT

multiple-medication therapy

MPM

malignant pleural mesothelioma

MTDL

multi-target-directed ligand

mTOR

mammalian target of rapamycin

NAD+

nicotinamide adenine dinucleotide

NAMPT

nicotinamide phosphoribosyltransferase

NAMPTi

nicotinamide phosphoribosyltransferase inhibitors

NFK

N-formylkynureine

NRASMut

NRAS-mutant

NSCLC

small cell lung carcinoma

PD

pharmacodynamics

PDE5

phosphodiesterase 5

PDGFR

platelet-derived growth factor receptor

PI3K

phosphatidylinositol-3-kinase

PI3K/Akt/mTOR

phosphatidylinositol-3-kinase

PK

pharmacokinetics

PML

promyelocytic leukemia gene

PRMT

arginine methyltransferase

PROTAC

proteolysis targeting chimera

pTau

tau protein

PTCL

peripheral T-cell lymphoma

RAR-β

retinoic acid receptor beta

RARE

retinoic acid responsive elements

SAHA

suberoylanilide hydroxamic acid

SAM

S-adenosylmethionine

SAR

structure-activity relationship

STS

soft tissue sarcoma

TG

transglutaminase

TG2i

TG2 inhibitors

TKi

tyrosine kinase inhibitors

TNBC

triple negative breast cancer

TR-FRET

fluorescence resonance energy transfer

UR

urothelial carcinoma

VEGF

vascular endothelial growth facto

VEGFR

vascular endothelial growth factor receptor

VPA

sodium valproate

wt

wild type

BIO-SKETCHES OF AUTHORS

Daniela Tomaselli graduated in Medicinal Chemistry at the University of Rome “La Sapienza”, Italy, in 2016, and in the same year she started her PhD in Life Sciences in the research group of Prof. A. Mai at Sapienza University. Her current research interest is focused on the epigenetic field and in particular in the design and synthesis of epigenetic modulators.

Alessia Lucidi graduated in Medicinal Chemistry at the University of Rome “La Sapienza”, Italy, in 2014. She received her Ph.D. in Pharmaceutical Sciences at the same University in 2017, with a thesis entitled “The Quinazoline Ring as Privileged Scaffold in Epigenetic Medicinal Chemistry”, advisor Prof. A. Mai. Her research activity has been focusing mainly on the design and synthesis of epigenetic modulators with potential application mainly in cancer.

Dante Rotili graduated in Medicinal Chemistry at the University of Rome “‘La Sapienza”‘ (Italy) in 2003. He received his Ph.D. in Pharmaceutical Sciences at the same University in 2007. In 2009/2010 he was research associate at the Department of Chemistry of the University of Oxford, where he worked on the development of chemoproteomic probes for the characterization of 2-oxoglutarate-dependent enzymes. Since 2011, he has been a tenured Assistant Professor of Medicinal Chemistry at the University of Rome “La Sapienza”. Since 2014, he has got the Italian National Habilitation to Associate Professor of Medicinal Chemistry and since 2017 to Full Professor of Medicinal Chemistry. His research activity has been focusing mainly on the development of modulators of epigenetic enzymes with potential applications in cancer, neurodegenerative, metabolic, and infectious diseases.

Antonello Mai graduated in Pharmacy at the University of Rome “La Sapienza”, Italy, in 1984. He received his Ph.D. in 1992 in Pharmaceutical Sciences, with a thesis entitled “Research on New Polycyclic Benzodizepines Active on Central Nervous System”, advisor Prof. M. Artico. In 1998, he was appointed Associate Professor of Medicinal Chemistry at the same University. In 2011, Prof. Mai was appointed Full Professor of Medicinal Chemistry at the Faculty of Pharmacy and Medicine, Sapienza University of Rome. He published more than 270 papers on peer-review high-impact factors journals. His research interests include the synthesis and biological evaluation of new bioactive compounds, in particular small molecule modulators of epigenetic targets.

REFERENCES

  • 1.Finley A, Copeland RA. Small molecule control of chromatin remodeling. Chem Biol. 2014;21(9):1196–1210. [DOI] [PubMed] [Google Scholar]
  • 2.Anighoro A, Bajorath J, Rastelli G. Polypharmacology: challenges and opportunities in drug discovery. J Med Chem. 2014;57(19):7874–7887. [DOI] [PubMed] [Google Scholar]
  • 3.de Lera AR, Ganesan A. Epigenetic Polypharmacology: from combination therapy to multitarget drugs. Clin Epigenetics. 2016;8:1–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Benedetti R, Conte M, Iside C, Altucci L. Epigenetic-based therapy: From single- to multi-target approaches. Int J Biochem Cell Biol. 2015;69:121–131. [DOI] [PubMed] [Google Scholar]
  • 5.Ganesan A Multitarget Drugs: an Epigenetic Epiphany. ChemMedChem. 2016;11(12):1227–1241. [DOI] [PubMed] [Google Scholar]
  • 6.Bannister AJ, Zegerman P, Partridge JF, et al. Selective recognition of methylated lysine 9 on histone H3 by the HP1 chromo domain. Nature. 2001;410(6824):120–124. [DOI] [PubMed] [Google Scholar]
  • 7.Yoo J, Choi S, Medina-Franco JL. Molecular modeling studies of the novel inhibitors of DNA methyltransferases SGI-1027 and CBC12: implications for the mechanism of inhibition of DNMTs. PLoS One 2013;8(4):e62152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Fahy J, Jeltsch A, Arimondo PB. DNA methyltransferase inhibitors in cancer: a chemical and therapeutic patent overview and selected clinical studies. Expert Opin Ther Pat. 2012;22(12):1427–1442. [DOI] [PubMed] [Google Scholar]
  • 9.Mei Y, Yang B. Rational application of drug promiscuity in medicinal chemistry. Future Med Chem. 2018;10(15):1835–1851. [DOI] [PubMed] [Google Scholar]
  • 10.Oscar Méndez-Lucio J Jesús Naveja HV-C, Fernando Daniel Prieto-Martínez, Medina-Franco aJL. Review. One Drug for Multiple Targets: A Computational Perspective. J Mex Chem Soc 2016;60(3):168–181. [Google Scholar]
  • 11.Ashburn TT, Thor KB. Drug repositioning: identifying and developing new uses for existing drugs. Nat Rev Drug Discov. 2004;3(8):673–683. [DOI] [PubMed] [Google Scholar]
  • 12.Lyko F The DNA methyltransferase family: a versatile toolkit for epigenetic regulation. Nat Rev Genet. 2018;19(2):81–92. [DOI] [PubMed] [Google Scholar]
  • 13.Pastor WA, Aravind L, Rao A. TETonic shift: biological roles of TET proteins in DNA demethylation and transcription. Nat Rev Mol Cell Biol. 2013;14(6):341–356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Lu X, Zhao BS, He C. TET family proteins: oxidation activity, interacting molecules, and functions in diseases. Chem Rev. 2015;115(6):2225–2239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Santi DV, Garrett CE, Barr PJ. On the mechanism of inhibition of DNA-cytosine methyltransferases by cytosine analogs. Cell. 1983;33(1):9–10. [DOI] [PubMed] [Google Scholar]
  • 16.Glozak MA, Sengupta N, Zhang X, Seto E. Acetylation and deacetylation of non-histone proteins. Gene. 2005;363:15–23. [DOI] [PubMed] [Google Scholar]
  • 17.Mai A, Massa S, Rotili D, et al. Histone deacetylation in epigenetics: an attractive target for anticancer therapy. Med Res Rev. 2005;25(3):261–309. [DOI] [PubMed] [Google Scholar]
  • 18.Blanquart C, Linot C, Cartron PF, Tomaselli D, Mai A, Bertrand P. Epigenetic metalloenzymes. Curr Med Chem. 2018. [DOI] [PubMed] [Google Scholar]
  • 19.Zwergel C, Valente S, Jacob C, Mai A. Emerging approaches for histone deacetylase inhibitor drug discovery. Expert Opin Drug Discov. 2015;10(6):599–613. [DOI] [PubMed] [Google Scholar]
  • 20.Faria Freitas M, Cuendet M, Bertrand P. HDAC inhibitors: a 2013–2017 patent survey. Expert Opin Ther Pat. 2018:1–17. [DOI] [PubMed] [Google Scholar]
  • 21.Kaniskan HU, Konze KD, Jin J. Selective inhibitors of protein methyltransferases. J Med Chem. 2015;58(4):1596–1629. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Schapira M, Ferreira de Freitas R. Structural biology and chemistry of protein arginine methyltransferases. MedChemComm. 2014;5(12):1779–1788. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Stazi G, Zwergel C, Mai A, Valente S. EZH2 inhibitors: a patent review (2014–2016). Expert Opin Ther Pat. 2017;27(7):797–813. [DOI] [PubMed] [Google Scholar]
  • 24.Fioravanti R, Stazi G, Zwergel C, Valente S, Mai A. Six Years (2012–2018) of Researches on Catalytic EZH2 Inhibitors: The Boom of the 2-Pyridone Compounds. Chem Rec 2018;18(12):1818–1832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Stein EM, Garcia-Manero G, Rizzieri DA, et al. The DOT1L inhibitor pinometostat reduces H3K79 methylation and has modest clinical activity in adult acute leukemia. Blood. 2018;131(24):2661–2669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Campbell CT, Haladyna JN, Drubin DA, et al. Mechanisms of Pinometostat (EPZ-5676) Treatment-Emergent Resistance in MLL-Rearranged Leukemia. Mol Cancer Ther. 2017;16(8):1669–1679. [DOI] [PubMed] [Google Scholar]
  • 27.Rotili D, Mai A. Targeting Histone Demethylases: A New Avenue for the Fight against Cancer. Genes Cancer. 2011;2(6):663–679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Gerhart SV, Kellner WA, Thompson C, et al. Activation of the p53-MDM4 regulatory axis defines the anti-tumour response to PRMT5 inhibition through its role in regulating cellular splicing. Sci Rep. 2018;8(1):9711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Heightman TD. Chemical biology of lysine demethylases. Curr Chem Genomics. 2011;5(Suppl 1):62–71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Binda C, Valente S, Romanenghi M, et al. Biochemical, structural, and biological evaluation of tranylcypromine derivatives as inhibitors of histone demethylases LSD1 and LSD2. J Am Chem Soc. 2010;132(19):6827–6833. [DOI] [PubMed] [Google Scholar]
  • 31.Stazi G, Zwergel C, Valente S, Mai A. LSD1 inhibitors: a patent review (2010–2015). Expert Opin Ther Pat. 2016;26(5):565–580. [DOI] [PubMed] [Google Scholar]
  • 32.Zheng YC, Ma J, Wang Z, et al. A Systematic Review of Histone Lysine-Specific Demethylase 1 and Its Inhibitors. Med Res Rev. 2015;35(5):1032–1071. [DOI] [PubMed] [Google Scholar]
  • 33.Filippakopoulos P, Knapp S. Targeting bromodomains: epigenetic readers of lysine acetylation. Nat Rev Drug Discov. 2014;13(5):337–356. [DOI] [PubMed] [Google Scholar]
  • 34.Brand M, Measures AR, Wilson BG, et al. Small molecule inhibitors of bromodomain-acetyl-lysine interactions. ACS Chem Biol. 2015;10(1):22–39. [DOI] [PubMed] [Google Scholar]
  • 35.Chou TC. Theoretical basis, experimental design, and computerized simulation of synergism and antagonism in drug combination studies. Pharmacol Rev. 2006;58(3):621–681. [DOI] [PubMed] [Google Scholar]
  • 36.Webster R, Castellano JM, Onuma OK. Putting polypills into practice: challenges and lessons learned. Lancet. 2017;389(10073):1066–1074. [DOI] [PubMed] [Google Scholar]
  • 37.de Lera AR, Ganesan A. Epigenetic Polypharmacology: from combination therapy to multitarget drugs. Clin Epigenetics. 2016;8:1–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Mroz EA, Rocco JW. The challenges of tumor genetic diversity. Cancer. 2017;123(6):917–927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Dumbrava EI, Meric-Bernstam F. Personalized cancer therapy-leveraging a knowledge base for clinical decision-making. Cold Spring Harb Mol Case Stud. 2018;4(2). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Morphy R, Rankovic Z. Designed multiple ligands. An emerging drug discovery paradigm. J Med Chem. 2005;48(21):6523–6543. [DOI] [PubMed] [Google Scholar]
  • 41.Yang XF, Zhao ZJ, Liu JJ, et al. SAHA and/or MG132 reverse the aggressive phenotypes of glioma cells: An in vitro and vivo study. Oncotarget. 2017;8(2):3156–3169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Glaser KB. HDAC inhibitors: clinical update and mechanism-based potential. Biochem Pharmacol. 2007;74(5):659–671. [DOI] [PubMed] [Google Scholar]
  • 43.Munster PN, Thurn KT, Thomas S, et al. A phase II study of the histone deacetylase inhibitor vorinostat combined with tamoxifen for the treatment of patients with hormone therapy-resistant breast cancer. Br J Cancer. 2011;104(12):1828–1835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Gore SD. Combination therapy with DNA methyltransferase inhibitors in hematologic malignancies. Nat Clin Pract Oncol. 2005;2 Suppl 1:S30–35. [DOI] [PubMed] [Google Scholar]
  • 45.Berry DH, Fernbach DJ, Herson J, Pullen J, Sullivan MP, Vietti TJ. Comparison of prednisolone, vincristine, methotrexate and 6-mercaptopurine vs. 6-mercaptopurine and prednisone maintenance therapy in childhood acute leukemia: a Southwest Oncology Group Study. Cancer. 1980;46(5):1098–1103. [DOI] [PubMed] [Google Scholar]
  • 46.Marchion DC, Bicaku E, Daud AI, Sullivan DM, Munster PN. Valproic acid alters chromatin structure by regulation of chromatin modulation proteins. Cancer Res. 2005;65(9):3815–3822. [DOI] [PubMed] [Google Scholar]
  • 47.Li J, Hao D, Wang L, et al. Epigenetic targeting drugs potentiate chemotherapeutic effects in solid tumor therapy. Sci Rep. 2017;7(1):4035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Christman JK. 5-Azacytidine and 5-aza-2’-deoxycytidine as inhibitors of DNA methylation: mechanistic studies and their implications for cancer therapy. Oncogene. 2002;21(35):5483–5495. [DOI] [PubMed] [Google Scholar]
  • 49.Negrotto S, Ng KP, Jankowska AM, et al. CpG methylation patterns and decitabine treatment response in acute myeloid leukemia cells and normal hematopoietic precursors. Leukemia. 2012;26(2):244–254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Link PA, Baer MR, James SR, Jones DA, Karpf AR. p53-inducible ribonucleotide reductase (p53R2/RRM2B) is a DNA hypomethylation-independent decitabine gene target that correlates with clinical response in myelodysplastic syndrome/acute myelogenous leukemia. Cancer Res. 2008;68(22):9358–9366. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Ikehata M, Ogawa M, Yamada Y, Tanaka S, Ueda K, Iwakawa S. Different effects of epigenetic modifiers on the cytotoxicity induced by 5-fluorouracil, irinotecan or oxaliplatin in colon cancer cells. Biol Pharm Bull. 2014;37(1):67–73. [DOI] [PubMed] [Google Scholar]
  • 52.Strathdee G, MacKean MJ, Illand M, Brown R. A role for methylation of the hMLH1 promoter in loss of hMLH1 expression and drug resistance in ovarian cancer. Oncogene. 1999;18(14):2335–2341. [DOI] [PubMed] [Google Scholar]
  • 53.Deroanne CF, Bonjean K, Servotte S, et al. Histone deacetylases inhibitors as anti-angiogenic agents altering vascular endothelial growth factor signaling. Oncogene. 2002;21(3):427–436. [DOI] [PubMed] [Google Scholar]
  • 54.Shen G, Khor TO, Hu R, et al. Chemoprevention of familial adenomatous polyposis by natural dietary compounds sulforaphane and dibenzoylmethane alone and in combination in ApcMin/+ mouse. Cancer Res. 2007;67(20):9937–9944. [DOI] [PubMed] [Google Scholar]
  • 55.Huang WJ, Tang YA, Chen MY, et al. A histone deacetylase inhibitor YCW1 with antitumor and antimetastasis properties enhances cisplatin activity against non-small cell lung cancer in preclinical studies. Cancer Lett. 2014;346(1):84–93. [DOI] [PubMed] [Google Scholar]
  • 56.Zhang X, Jiang SJ, Shang B, Jiang HJ. Effects of histone deacetylase inhibitor trichostatin A combined with cisplatin on apoptosis of A549 cell line. Thorac Cancer. 2015;6(2):202–208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Rikiishi H, Shinohara F, Sato T, Sato Y, Suzuki M, Echigo S. Chemosensitization of oral squamous cell carcinoma cells to cisplatin by histone deacetylase inhibitor, suberoylanilide hydroxamic acid. Int J Oncol. 2007;30(5):1181–1188. [PubMed] [Google Scholar]
  • 58.Kato Y, Yoshimura K, Shin T, et al. Synergistic in vivo antitumor effect of the histone deacetylase inhibitor MS-275 in combination with interleukin 2 in a murine model of renal cell carcinoma. Clin Cancer Res. 2007;13(15 Pt 1):4538–4546. [DOI] [PubMed] [Google Scholar]
  • 59.Hollenbach PW, Nguyen AN, Brady H, et al. A comparison of azacitidine and decitabine activities in acute myeloid leukemia cell lines. PLoS One. 2010;5(2):e9001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Moreaux J, Reme T, Leonard W, et al. Development of gene expression-based score to predict sensitivity of multiple myeloma cells to DNA methylation inhibitors. Mol Cancer Ther. 2012;11(12):2685–2692. [DOI] [PubMed] [Google Scholar]
  • 61.Moreaux J, Reme T, Leonard W, et al. Gene expression-based prediction of myeloma cell sensitivity to histone deacetylase inhibitors. Br J Cancer. 2013;109(3):676–685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Cameron EE, Bachman KE, Myohanen S, Herman JG, Baylin SB. Synergy of demethylation and histone deacetylase inhibition in the re-expression of genes silenced in cancer. Nat Genet. 1999;21(1):103–107. [DOI] [PubMed] [Google Scholar]
  • 63.Liu HB, Urbanavicius D, Tan P, Spencer A, Dear AE. Mechanisms and potential molecular markers of early response to combination epigenetic therapy in patients with myeloid malignancies. Int J Oncol. 2014;45(4):1742–1748. [DOI] [PubMed] [Google Scholar]
  • 64.Shaker S, Bernstein M, Momparler LF, Momparler RL. Preclinical evaluation of antineoplastic activity of inhibitors of DNA methylation (5-aza-2’-deoxycytidine) and histone deacetylation (trichostatin A, depsipeptide) in combination against myeloid leukemic cells. Leuk Res. 2003;27(5):437–444. [DOI] [PubMed] [Google Scholar]
  • 65.Narita N, Fujieda S, Tokuriki M, et al. Inhibition of histone deacetylase 3 stimulates apoptosis induced by heat shock under acidic conditions in human maxillary cancer. Oncogene. 2005;24(49):7346–7354. [DOI] [PubMed] [Google Scholar]
  • 66.Gressette M, Verillaud B, Jimenez-Pailhes AS, et al. Treatment of nasopharyngeal carcinoma cells with the histone-deacetylase inhibitor abexinostat: cooperative effects with cis-platin and radiotherapy on patient-derived xenografts. PLoS One. 2014;9(3):e91325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Kejik Z, Jakubek M, Kaplanek R, et al. Epigenetic agents in combined anticancer therapy. Future Med Chem. 2018;10(9):1113–1130. [DOI] [PubMed] [Google Scholar]
  • 68.Han T, Zhuo M, Hu H, Jiao F, Wang LW. Synergistic effects of the combination of 5-AzaCdR and suberoylanilide hydroxamic acid on the anticancer property of pancreatic cancer. Oncol Rep. 2018;39(1):264–270. [DOI] [PubMed] [Google Scholar]
  • 69.Wagner JC, Sleggs CA, Marchand P. Diffuse pleural mesothelioma and asbestos exposure in the North Western Cape Province. Br J Ind Med. 1960;17:260–271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Pan Y, Liu G, Zhou F, Su B, Li Y. DNA methylation profiles in cancer diagnosis and therapeutics. Clin Exp Med. 2018;18(1):1–14. [DOI] [PubMed] [Google Scholar]
  • 71.Zhang X, Tang N, Rishi AK, Pass HI, Wali A. Methylation profile landscape in mesothelioma: possible implications in early detection, disease progression, and therapeutic options. Methods Mol Biol. 2015;1238:235–247. [DOI] [PubMed] [Google Scholar]
  • 72.Fraga MF, Ballestar E, Villar-Garea A, et al. Loss of acetylation at Lys16 and trimethylation at Lys20 of histone H4 is a common hallmark of human cancer. Nat Genet. 2005;37(4):391–400. [DOI] [PubMed] [Google Scholar]
  • 73.Chi P, Allis CD, Wang GG. Covalent histone modifications--miswritten, misinterpreted and mis-erased in human cancers. Nat Rev Cancer. 2010;10(7):457–469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Leclercq S, Gueugnon F, Boutin B, et al. A 5-aza-2’-deoxycytidine/valproate combination induces cytotoxic T-cell response against mesothelioma. Eur Respir J. 2011;38(5):1105–1116. [DOI] [PubMed] [Google Scholar]
  • 75.Weiser TS, Guo ZS, Ohnmacht GA, et al. Sequential 5-Aza-2 deoxycytidine-depsipeptide FR901228 treatment induces apoptosis preferentially in cancer cells and facilitates their recognition by cytolytic T lymphocytes specific for NY-ESO-1. J Immunother. 2001;24(2):151–161. [DOI] [PubMed] [Google Scholar]
  • 76.Li Y, Yuan YY, Meeran SM, Tollefsbol TO. Synergistic epigenetic reactivation of estrogen receptor-alpha (ERalpha) by combined green tea polyphenol and histone deacetylase inhibitor in ERalpha-negative breast cancer cells. Mol Cancer. 2010;9:274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Bensaid D, Blondy T, Deshayes S, et al. Assessment of new HDAC inhibitors for immunotherapy of malignant pleural mesothelioma. Clin Epigenetics. 2018;10:79. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Charrier C, Clarhaut J, Gesson JP, et al. Synthesis and modeling of new benzofuranone histone deacetylase inhibitors that stimulate tumor suppressor gene expression. J Med Chem. 2009;52(9):3112–3115. [DOI] [PubMed] [Google Scholar]
  • 79.Gueugnon F, Cartron PF, Charrier C, et al. New histone deacetylase inhibitors improve cisplatin antitumor properties against thoracic cancer cells. Oncotarget. 2014;5(12):4504–4515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Merino VF, Cho S, Nguyen N, et al. Induction of cell cycle arrest and inflammatory genes by combined treatment with epigenetic, differentiating, and chemotherapeutic agents in triple-negative breast cancer. Breast Cancer Res. 2018;20(1):145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Merino VF, Nguyen N, Jin K, et al. Combined Treatment with Epigenetic, Differentiating, and Chemotherapeutic Agents Cooperatively Targets Tumor-Initiating Cells in Triple-Negative Breast Cancer. Cancer Res. 2016;76(7):2013–2024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Topper MJ, Vaz M, Chiappinelli KB, et al. Epigenetic Therapy Ties MYC Depletion to Reversing Immune Evasion and Treating Lung Cancer. Cell. 2017;171(6):1284–1300 e1221. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Seewaldt VL, Johnson BS, Parker MB, Collins SJ, Swisshelm K. Expression of retinoic acid receptor beta mediates retinoic acid-induced growth arrest and apoptosis in breast cancer cells. Cell Growth Differ. 1995;6(9):1077–1088. [PubMed] [Google Scholar]
  • 84.Ginestier C, Wicinski J, Cervera N, et al. Retinoid signaling regulates breast cancer stem cell differentiation. Cell Cycle. 2009;8(20):3297–3302. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Bhat-Nakshatri P, Goswami CP, Badve S, Sledge GW Jr., Nakshatri H. Identification of FDA-approved drugs targeting breast cancer stem cells along with biomarkers of sensitivity. Sci Rep. 2013;3:2530. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Hall JA, Grainger JR, Spencer SP, Belkaid Y. The role of retinoic acid in tolerance and immunity. Immunity. 2011;35(1):13–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Connolly RM, Nguyen NK, Sukumar S. Molecular pathways: current role and future directions of the retinoic acid pathway in cancer prevention and treatment. Clin Cancer Res. 2013;19(7):1651–1659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Anso E, Mullen AR, Felsher DW, Mates JM, Deberardinis RJ, Chandel NS. Metabolic changes in cancer cells upon suppression of MYC. Cancer Metab. 2013;1(1):7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Tang XH, Gudas LJ. Retinoids, retinoic acid receptors, and cancer. Annu Rev Pathol. 2011;6:345–364. [DOI] [PubMed] [Google Scholar]
  • 90.Sirchia SM, Ferguson AT, Sironi E, et al. Evidence of epigenetic changes affecting the chromatin state of the retinoic acid receptor beta2 promoter in breast cancer cells. Oncogene. 2000;19(12):1556–1563. [DOI] [PubMed] [Google Scholar]
  • 91.Nacht M, Ferguson AT, Zhang W, et al. Combining serial analysis of gene expression and array technologies to identify genes differentially expressed in breast cancer. Cancer Res. 1999;59(21):5464–5470. [PubMed] [Google Scholar]
  • 92.Kitaura M, Suzuki N, Imai T, et al. Molecular cloning of a novel human CC chemokine (Eotaxin-3) that is a functional ligand of CC chemokine receptor 3. J Biol Chem. 1999;274(39):27975–27980. [DOI] [PubMed] [Google Scholar]
  • 93.Oliver JR, Kushwah R, Hu J. Multiple roles of the epithelium-specific ETS transcription factor, ESE-1, in development and disease. Lab Invest. 2012;92(3):320–330. [DOI] [PubMed] [Google Scholar]
  • 94.Kushwah R, Oliver JR, Wu J, Chang Z, Hu J. Elf3 regulates allergic airway inflammation by controlling dendritic cell-driven T cell differentiation. J Immunol. 2011;187(9):4639–4653. [DOI] [PubMed] [Google Scholar]
  • 95.Jang MH, Kim HJ, Kim EJ, Chung YR, Park SY. Expression of epithelial-mesenchymal transition-related markers in triple-negative breast cancer: ZEB1 as a potential biomarker for poor clinical outcome. Hum Pathol. 2015;46(9):1267–1274. [DOI] [PubMed] [Google Scholar]
  • 96.Idowu MO, Kmieciak M, Dumur C, et al. CD44(+)/CD24(-/low) cancer stem/progenitor cells are more abundant in triple-negative invasive breast carcinoma phenotype and are associated with poor outcome. Hum Pathol. 2012;43(3):364–373. [DOI] [PubMed] [Google Scholar]
  • 97.Peinado H, Ballestar E, Esteller M, Cano A. Snail mediates E-cadherin repression by the recruitment of the Sin3A/histone deacetylase 1 (HDAC1)/HDAC2 complex. Mol Cell Biol. 2004;24(1):306–319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Hajra KM, Chen DY, Fearon ER. The SLUG zinc-finger protein represses E-cadherin in breast cancer. Cancer Res. 2002;62(6):1613–1618. [PubMed] [Google Scholar]
  • 99.Park SM, Gaur AB, Lengyel E, Peter ME. The miR-200 family determines the epithelial phenotype of cancer cells by targeting the E-cadherin repressors ZEB1 and ZEB2. Genes Dev. 2008;22(7):894–907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Vesuna F, van Diest P, Chen JH, Raman V. Twist is a transcriptional repressor of E-cadherin gene expression in breast cancer. Biochem Biophys Res Commun. 2008;367(2):235–241. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Hemavathy K, Guru SC, Harris J, Chen JD, Ip YT. Human Slug is a repressor that localizes to sites of active transcription. Mol Cell Biol. 2000;20(14):5087–5095. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Shi Y, Sawada J, Sui G, et al. Coordinated histone modifications mediated by a CtBP co-repressor complex. Nature. 2003;422(6933):735–738. [DOI] [PubMed] [Google Scholar]
  • 103.Aghdassi A, Sendler M, Guenther A, et al. Recruitment of histone deacetylases HDAC1 and HDAC2 by the transcriptional repressor ZEB1 downregulates E-cadherin expression in pancreatic cancer. Gut. 2012;61(3):439–448. [DOI] [PubMed] [Google Scholar]
  • 104.Fu J, Qin L, He T, et al. The TWIST/Mi2/NuRD protein complex and its essential role in cancer metastasis. Cell Res. 2011;21(2):275–289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Ramadoss S, Chen X, Wang CY. Histone demethylase KDM6B promotes epithelial-mesenchymal transition. J Biol Chem. 2012;287(53):44508–44517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Su Y, Pogash TJ, Nguyen TD, Russo J. Development and characterization of two human triple-negative breast cancer cell lines with highly tumorigenic and metastatic capabilities. Cancer Med. 2016;5(3):558–573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Su Y, Hopfinger NR, Nguyen TD, Pogash TJ, 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(1):314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Hardy TM, Tollefsbol TO. Epigenetic diet: impact on the epigenome and cancer. Epigenomics. 2011;3(4):503–518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Herceg Z. Epigenetics and cancer: towards an evaluation of the impact of environmental and dietary factors. Mutagenesis. 2007;22(2):91–103. [DOI] [PubMed] [Google Scholar]
  • 110.Russo GL, Vastolo V, Ciccarelli M, Albano L, Macchia PE, Ungaro P. Dietary polyphenols and chromatin remodeling. Crit Rev Food Sci Nutr. 2017;57(12):2589–2599. [DOI] [PubMed] [Google Scholar]
  • 111.Li Y, Meeran SM, Tollefsbol TO. Combinatorial bioactive botanicals re-sensitize tamoxifen treatment in ER-negative breast cancer via epigenetic reactivation of ERalpha expression. Sci Rep. 2017;7(1):9345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Xie Q, Bai Q, Zou LY, et al. Genistein inhibits DNA methylation and increases expression of tumor suppressor genes in human breast cancer cells. Genes Chromosomes Cancer. 2014;53(5):422–431. [DOI] [PubMed] [Google Scholar]
  • 113.Tortorella SM, Royce SG, Licciardi PV, Karagiannis TC. Dietary Sulforaphane in Cancer Chemoprevention: The Role of Epigenetic Regulation and HDAC Inhibition. Antioxid Redox Signal. 2015;22(16):1382–1424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Majid S, Dar AA, Ahmad AE, et al. BTG3 tumor suppressor gene promoter demethylation, histone modification and cell cycle arrest by genistein in renal cancer. Carcinogenesis. 2009;30(4):662–670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Sasamura H, Takahashi A, Yuan J, et al. Antiproliferative and antiangiogenic activities of genistein in human renal cell carcinoma. Urology. 2004;64(2):389–393. [DOI] [PubMed] [Google Scholar]
  • 116.Sinha S, Shukla S, Khan S, Tollefsbol TO, Meeran SM. Epigenetic reactivation of p21CIP1/WAF1 and KLOTHO by a combination of bioactive dietary supplements is partially ERalpha-dependent in ERalpha-negative human breast cancer cells. Mol Cell Endocrinol. 2015;406:102–114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Meeran SM, Patel SN, Li Y, Shukla S, Tollefsbol TO. Bioactive dietary supplements reactivate ER expression in ER-negative breast cancer cells by active chromatin modifications. PLoS One. 2012;7(5):e37748. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Kikuno N, Shiina H, Urakami S, et al. Genistein mediated histone acetylation and demethylation activates tumor suppressor genes in prostate cancer cells. Int J Cancer. 2008;123(3):552–560. [DOI] [PubMed] [Google Scholar]
  • 119.Ganai SA. Histone deacetylase inhibitor sulforaphane: The phytochemical with vibrant activity against prostate cancer. Biomed Pharmacother. 2016;81:250–257. [DOI] [PubMed] [Google Scholar]
  • 120.Tarapore RS, Siddiqui IA, Mukhtar H. Modulation of Wnt/beta-catenin signaling pathway by bioactive food components. Carcinogenesis. 2012;33(3):483–491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Meeran SM, Patel SN, Tollefsbol TO. Sulforaphane causes epigenetic repression of hTERT expression in human breast cancer cell lines. PLoS One. 2010;5(7):e11457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Traka MH, Chambers KF, Lund EK, Goodlad RA, Johnson IT, Mithen RF. Involvement of KLF4 in sulforaphane- and iberin-mediated induction of p21(waf1/cip1). Nutr Cancer. 2009;61(1):137–145. [DOI] [PubMed] [Google Scholar]
  • 123.Royston KJ, Udayakumar N, Lewis K, Tollefsbol TO. A Novel Combination of Withaferin A and Sulforaphane Inhibits Epigenetic Machinery, Cellular Viability and Induces Apoptosis of Breast Cancer Cells. Int J Mol Sci. 2017;18(5). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Paul B, Li Y, Tollefsbol TO. The Effects of Combinatorial Genistein and Sulforaphane in Breast Tumor Inhibition: Role in Epigenetic Regulation. Int J Mol Sci. 2018;19(6). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Fahey JW, Zhang Y, Talalay P. Broccoli sprouts: an exceptionally rich source of inducers of enzymes that protect against chemical carcinogens. Proc Natl Acad Sci U S A. 1997;94(19):10367–10372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Atwell LL, Hsu A, Wong CP, et al. Absorption and chemopreventive targets of sulforaphane in humans following consumption of broccoli sprouts or a myrosinase-treated broccoli sprout extract. Mol Nutr Food Res. 2015;59(3):424–433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Lee AV, Oesterreich S, Davidson NE. MCF-7 cells--changing the course of breast cancer research and care for 45 years. J Natl Cancer Inst. 2015;107(7). [DOI] [PubMed] [Google Scholar]
  • 128.Easwaran H, Tsai HC, Baylin SB. Cancer epigenetics: tumor heterogeneity, plasticity of stem-like states, and drug resistance. Mol Cell. 2014;54(5):716–727. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Seo SK, Jin HO, Woo SH, et al. Histone deacetylase inhibitors sensitize human non-small cell lung cancer cells to ionizing radiation through acetyl p53-mediated c-myc down-regulation. J Thorac Oncol. 2011;6(8):1313–1319. [DOI] [PubMed] [Google Scholar]
  • 130.Miller KM, Tjeertes JV, Coates J, et al. Human HDAC1 and HDAC2 function in the DNA-damage response to promote DNA nonhomologous end-joining. Nat Struct Mol Biol. 2010;17(9):1144–1151. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Roos WP, Krumm A. The multifaceted influence of histone deacetylases on DNA damage signalling and DNA repair. Nucleic Acids Res. 2016;44(21):10017–10030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Lee JH, Choy ML, Ngo L, Foster SS, Marks PA. Histone deacetylase inhibitor induces DNA damage, which normal but not transformed cells can repair. Proc Natl Acad Sci U S A. 2010;107(33):14639–14644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Nikolova T, Kiweler N, Kramer OH. Interstrand Crosslink Repair as a Target for HDAC Inhibition. Trends Pharmacol Sci. 2017;38(9):822–836. [DOI] [PubMed] [Google Scholar]
  • 134.Marchion DC, Bicaku E, Turner JG, Schmitt ML, Morelli DR, Munster PN. HDAC2 regulates chromatin plasticity and enhances DNA vulnerability. Mol Cancer Ther. 2009;8(4):794–801. [DOI] [PubMed] [Google Scholar]
  • 135.Storch K, Eke I, Borgmann K, et al. Three-dimensional cell growth confers radioresistance by chromatin density modification. Cancer Res. 2010;70(10):3925–3934. [DOI] [PubMed] [Google Scholar]
  • 136.Schiller JH, Harrington D, Belani CP, et al. Comparison of four chemotherapy regimens for advanced non-small-cell lung cancer. N Engl J Med. 2002;346(2):92–98. [DOI] [PubMed] [Google Scholar]
  • 137.Group NM-AC. Chemotherapy in addition to supportive care improves survival in advanced non-small-cell lung cancer: a systematic review and meta-analysis of individual patient data from 16 randomized controlled trials. J Clin Oncol. 2008;26(28):4617–4625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.D’Addario G, Pintilie M, Leighl NB, Feld R, Cerny T, Shepherd FA. Platinum-based versus non-platinum-based chemotherapy in advanced non-small-cell lung cancer: a meta-analysis of the published literature. J Clin Oncol. 2005;23(13):2926–2936. [DOI] [PubMed] [Google Scholar]
  • 139.Tarhini AA, Zahoor H, McLaughlin B, et al. Phase I trial of carboplatin and etoposide in combination with panobinostat in patients with lung cancer. Anticancer Res. 2013;33(10):4475–4481. [PMC free article] [PubMed] [Google Scholar]
  • 140.Wang L, Syn NL, Subhash VV, et al. Pan-HDAC inhibition by panobinostat mediates chemosensitization to carboplatin in non-small cell lung cancer via attenuation of EGFR signaling. Cancer Lett. 2018;417:152–160. [DOI] [PubMed] [Google Scholar]
  • 141.Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–674. [DOI] [PubMed] [Google Scholar]
  • 142.Whitesell L, Lindquist SL. HSP90 and the chaperoning of cancer. Nat Rev Cancer. 2005;5(10):761–772. [DOI] [PubMed] [Google Scholar]
  • 143.Zhang M, Zhang X, Bai CX, Chen J, Wei MQ. Inhibition of epidermal growth factor receptor expression by RNA interference in A549 cells. Acta Pharmacol Sin. 2004;25(1):61–67. [PubMed] [Google Scholar]
  • 144.Tu CY, Chen CH, Hsia TC, et al. Trichostatin A suppresses EGFR expression through induction of microRNA-7 in an HDAC-independent manner in lapatinib-treated cells. Biomed Res Int. 2014;2014:168949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Edwards A, Li J, Atadja P, Bhalla K, Haura EB. Effect of the histone deacetylase inhibitor LBH589 against epidermal growth factor receptor-dependent human lung cancer cells. Mol Cancer Ther. 2007;6(9):2515–2524. [DOI] [PubMed] [Google Scholar]
  • 146.Gray JE, Haura E, Chiappori A, et al. A phase I, pharmacokinetic, and pharmacodynamic study of panobinostat, an HDAC inhibitor, combined with erlotinib in patients with advanced aerodigestive tract tumors. Clin Cancer Res. 2014;20(6):1644–1655. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Johnson DB, Puzanov I. Treatment of NRAS-mutant melanoma. Curr Treat Options Oncol. 2015;16(4):15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Kwong LN, Davies MA. Targeted therapy for melanoma: rational combinatorial approaches. Oncogene. 2014;33(1):1–9. [DOI] [PubMed] [Google Scholar]
  • 149.Luke JJ, Flaherty KT, Ribas A, Long GV. Targeted agents and immunotherapies: optimizing outcomes in melanoma. Nat Rev Clin Oncol. 2017;14(8):463–482. [DOI] [PubMed] [Google Scholar]
  • 150.Sullivan RJ, Flaherty K. MAP kinase signaling and inhibition in melanoma. Oncogene. 2013;32(19):2373–2379. [DOI] [PubMed] [Google Scholar]
  • 151.Johnson DB, Smalley KS, Sosman JA. Molecular pathways: targeting NRAS in melanoma and acute myelogenous leukemia. Clin Cancer Res. 2014;20(16):4186–4192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Posch C, Vujic I, Monshi B, et al. Searching for the Chokehold of NRAS Mutant Melanoma. J Invest Dermatol. 2016;136(7):1330–1336. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Vu HL, Aplin AE. Targeting mutant NRAS signaling pathways in melanoma. Pharmacol Res. 2016;107:111–116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Wong DJ, Ribas A. Targeted Therapy for Melanoma. Cancer Treat Res. 2016;167:251–262. [DOI] [PubMed] [Google Scholar]
  • 155.Echevarria-Vargas IM, Reyes-Uribe PI, Guterres AN, et al. Co-targeting BET and MEK as salvage therapy for MAPK and checkpoint inhibitor-resistant melanoma. EMBO Mol Med. 2018;10(5). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Zhao L, Okhovat JP, Hong EK, Kim YH, Wood GS. Preclinical Studies Support Combined Inhibition of BET Family Proteins and Histone Deacetylases as Epigenetic Therapy for Cutaneous T-Cell Lymphoma. Neoplasia. 2018;21(1):82–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Holscher AS, Schulz WA, Pinkerneil M, Niegisch G, Hoffmann MJ. Combined inhibition of BET proteins and class I HDACs synergistically induces apoptosis in urothelial carcinoma cell lines. Clin Epigenetics. 2018;10:1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Heinemann A, Cullinane C, De Paoli-Iseppi R, et al. Combining BET and HDAC inhibitors synergistically induces apoptosis of melanoma and suppresses AKT and YAP signaling. Oncotarget. 2015;6(25):21507–21521. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Mazur PK, Herner A, Mello SS, et al. Combined inhibition of BET family proteins and histone deacetylases as a potential epigenetics-based therapy for pancreatic ductal adenocarcinoma. Nat Med. 2015;21(10):1163-+. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Bhattacharya S, Ray RM, Johnson LR. STAT3-mediated transcription of Bcl-2, Mcl-1 and c-IAP2 prevents apoptosis in polyamine-depleted cells. Biochem J. 2005;392(Pt 2):335–344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Sekulic A, Liang WS, Tembe W, et al. Personalized treatment of Sezary syndrome by targeting a novel CTLA4:CD28 fusion. Mol Genet Genomic Med. 2015;3(2):130–136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Jung JT, Kim DH, Kwak EK, et al. Clinical role of Bcl-2, Bax, or p53 overexpression in peripheral T-cell lymphomas. Ann Hematol. 2006;85(9):575–581. [DOI] [PubMed] [Google Scholar]
  • 163.Lindahl LM, Fredholm S, Joseph C, et al. STAT5 induces miR-21 expression in cutaneous T cell lymphoma. Oncotarget. 2016;7(29):45730–45744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Nielsen M, Kaestel CG, Eriksen KW, et al. Inhibition of constitutively activated Stat3 correlates with altered Bcl-2/Bax expression and induction of apoptosis in mycosis fungoides tumor cells. Leukemia. 1999;13(5):735–738. [DOI] [PubMed] [Google Scholar]
  • 165.Janiga J, Kentley J, Nabhan C, Abdulla F. Current systemic therapeutic options for advanced mycosis fungoides and Sezary syndrome. Leuk Lymphoma. 2018;59(3):562–577. [DOI] [PubMed] [Google Scholar]
  • 166.Rozati S, Cheng PF, Widmer DS, Fujii K, Levesque MP, Dummer R. Romidepsin and Azacitidine Synergize in their Epigenetic Modulatory Effects to Induce Apoptosis in CTCL. Clin Cancer Res. 2016;22(8):2020–2031. [DOI] [PubMed] [Google Scholar]
  • 167.Zuber J, Shi J, Wang E, et al. RNAi screen identifies Brd4 as a therapeutic target in acute myeloid leukaemia. Nature. 2011;478(7370):524–528. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Delmore JE, Issa GC, Lemieux ME, et al. BET bromodomain inhibition as a therapeutic strategy to target c-Myc. Cell. 2011;146(6):904–917. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Mertz JA, Conery AR, Bryant BM, et al. Targeting MYC dependence in cancer by inhibiting BET bromodomains. Proc Natl Acad Sci U S A. 2011;108(40):16669–16674. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Loven J, Hoke HA, Lin CY, et al. Selective inhibition of tumor oncogenes by disruption of super-enhancers. Cell. 2013;153(2):320–334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Souers AJ, Leverson JD, Boghaert ER, et al. ABT-199, a potent and selective BCL-2 inhibitor, achieves antitumor activity while sparing platelets. Nat Med. 2013;19(2):202–208. [DOI] [PubMed] [Google Scholar]
  • 172.Cyrenne BM, Lewis JM, Weed JG, et al. Synergy of BCL2 and histone deacetylase inhibition against leukemic cells from cutaneous T-cell lymphoma patients. Blood. 2017;130(19):2073–2083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Wang H, Hong B, Li X, et al. JQ1 synergizes with the Bcl-2 inhibitor ABT-263 against MYCN-amplified small cell lung cancer. Oncotarget. 2017;8(49):86312–86324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Kamijo H, Sugaya M, Takahashi N, et al. BET bromodomain inhibitor JQ1 decreases CD30 and CCR4 expression and proliferation of cutaneous T-cell lymphoma cell lines. Arch Dermatol Res. 2017;309(6):491–497. [DOI] [PubMed] [Google Scholar]
  • 175.Kim SR, Lewis JM, Cyrenne BM, et al. BET inhibition in advanced cutaneous T cell lymphoma is synergistically potentiated by BCL2 inhibition or HDAC inhibition. Oncotarget. 2018;9(49):29193–29207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Wang Y, Chen SY, Colborne S, et al. Histone Deacetylase Inhibitors Synergize with Catalytic Inhibitors of EZH2 to Exhibit Antitumor Activity in Small Cell Carcinoma of the Ovary, Hypercalcemic Type. Mol Cancer Ther. 2018;17(12):2767–2779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Gamwell LF, Gambaro K, Merziotis M, et al. Small cell ovarian carcinoma: genomic stability and responsiveness to therapeutics. Orphanet J Rare Dis. 2013;8:33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Otte A, Rauprich F, von der Ohe J, et al. c-Met inhibitors attenuate tumor growth of small cell hypercalcemic ovarian carcinoma (SCCOHT) populations. Oncotarget. 2015;6(31):31640–31658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Chan-Penebre E, Armstrong K, Drew A, et al. Selective Killing of SMARCA2- and SMARCA4-deficient Small Cell Carcinoma of the Ovary, Hypercalcemic Type Cells by Inhibition of EZH2: In Vitro and In Vivo Preclinical Models. Mol Cancer Ther. 2017;16(5):850–860. [DOI] [PubMed] [Google Scholar]
  • 180.Wang Y, Chen SY, Karnezis AN, et al. The histone methyltransferase EZH2 is a therapeutic target in small cell carcinoma of the ovary, hypercalcaemic type. J Pathol. 2017;242(3):371–383. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Arts J, King P, Marien A, et al. JNJ-26481585, a novel “second-generation” oral histone deacetylase inhibitor, shows broad-spectrum preclinical antitumoral activity. Clin Cancer Res. 2009;15(22):6841–6851. [DOI] [PubMed] [Google Scholar]
  • 182.Manara MC, Valente S, Cristalli C, et al. A Quinoline-Based DNA Methyltransferase Inhibitor as a Possible Adjuvant in Osteosarcoma Therapy. Mol Cancer Ther. 2018;17(9):1881–1892. [DOI] [PubMed] [Google Scholar]
  • 183.Folkman J Tumor angiogenesis: therapeutic implications. N Engl J Med. 1971;285(21):1182–1186. [DOI] [PubMed] [Google Scholar]
  • 184.Kuhnen C, Lehnhardt M, Tolnay E, Muehlberger T, Vogt PM, Muller KM. Patterns of expression and secretion of vascular endothelial growth factor in malignant soft-tissue tumours. J Cancer Res Clin Oncol. 2000;126(4):219–225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Potti A, Ganti AK, Tendulkar K, et al. Determination of vascular endothelial growth factor (VEGF) overexpression in soft tissue sarcomas and the role of overexpression in leiomyosarcoma. J Cancer Res Clin Oncol. 2004;130(1):52–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Verschraegen CF, Arias-Pulido H, Lee SJ, et al. Phase IB study of the combination of docetaxel, gemcitabine, and bevacizumab in patients with advanced or recurrent soft tissue sarcoma: the Axtell regimen. Ann Oncol. 2012;23(3):785–790. [DOI] [PubMed] [Google Scholar]
  • 187.Agulnik M, Yarber JL, Okuno SH, et al. An open-label, multicenter, phase II study of bevacizumab for the treatment of angiosarcoma and epithelioid hemangioendotheliomas. Ann Oncol. 2013;24(1):257–263. [DOI] [PubMed] [Google Scholar]
  • 188.Monga V, Swami U, Tanas M, et al. A Phase I/II Study Targeting Angiogenesis Using Bevacizumab Combined with Chemotherapy and a Histone Deacetylase Inhibitor (Valproic Acid) in Advanced Sarcomas. Cancers (Basel). 2018;10(2). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Bernstein ID. Monoclonal antibodies to the myeloid stem cells: therapeutic implications of CMA-676, a humanized anti-CD33 antibody calicheamicin conjugate. Leukemia. 2000;14(3):474–475. [DOI] [PubMed] [Google Scholar]
  • 190.Vasu S, He S, Cheney C, et al. Decitabine enhances anti-CD33 monoclonal antibody BI 836858-mediated natural killer ADCC against AML blasts. Blood. 2016;127(23):2879–2889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Balaian L, Ball ED. Cytotoxic activity of gemtuzumab ozogamicin (Mylotarg) in acute myeloid leukemia correlates with the expression of protein kinase Syk. Leukemia. 2006;20(12):2093–2101. [DOI] [PubMed] [Google Scholar]
  • 192.Nand S, Othus M, Godwin JE, et al. A phase 2 trial of azacitidine and gemtuzumab ozogamicin therapy in older patients with acute myeloid leukemia. Blood. 2013;122(20):3432–3439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Linenberger ML, Hong T, Flowers D, et al. Multidrug-resistance phenotype and clinical responses to gemtuzumab ozogamicin. Blood. 2001;98(4):988–994. [DOI] [PubMed] [Google Scholar]
  • 194.Ball ED MB, Balaian L, et al. Update of a Phase I/II Trial of 5-Azacytidine Prior to Gemtuzumab Ozogamicin (GO) for Patients with Relapsed Acute Myeloid Leukemia with Correlative Biomarker Studies. Blood. 2010. [Google Scholar]
  • 195.Medeiros BC, Tanaka TN, Balaian L, et al. A Phase I/II Trial of the Combination of Azacitidine and Gemtuzumab Ozogamicin for Treatment of Relapsed Acute Myeloid Leukemia. Clin Lymphoma Myeloma Leuk. 2018;18(5):346–352 e345. [DOI] [PubMed] [Google Scholar]
  • 196.Larson RA, Sievers EL, Stadtmauer EA, et al. Final report of the efficacy and safety of gemtuzumab ozogamicin (Mylotarg) in patients with CD33-positive acute myeloid leukemia in first recurrence. Cancer. 2005;104(7):1442–1452. [DOI] [PubMed] [Google Scholar]
  • 197.Cai C, Yuan X, Balk SP. Androgen receptor epigenetics. Transl Androl Urol. 2013;2(3):148–157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Cucchiara V, Yang JC, Mirone V, Gao AC, Rosenfeld MG, Evans CP. Epigenomic Regulation of Androgen Receptor Signaling: Potential Role in Prostate Cancer Therapy. Cancers (Basel). 2017;9(1). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Lavery DN, Bevan CL. Androgen receptor signalling in prostate cancer: the functional consequences of acetylation. J Biomed Biotechnol. 2011;2011:862125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Urbanucci A, Marttila S, Janne OA, Visakorpi T. Androgen receptor overexpression alters binding dynamics of the receptor to chromatin and chromatin structure. Prostate. 2012;72(11):1223–1232. [DOI] [PubMed] [Google Scholar]
  • 201.Xia Q, Sung J, Chowdhury W, et al. Chronic administration of valproic acid inhibits prostate cancer cell growth in vitro and in vivo. Cancer Res. 2006;66(14):7237–7244. [DOI] [PubMed] [Google Scholar]
  • 202.Marks P, Rifkind RA, Richon VM, Breslow R, Miller T, Kelly WK. Histone deacetylases and cancer: causes and therapies. Nat Rev Cancer. 2001;1(3):194–202. [DOI] [PubMed] [Google Scholar]
  • 203.Azad N, Zahnow CA, Rudin CM, Baylin SB. The future of epigenetic therapy in solid tumours--lessons from the past. Nat Rev Clin Oncol. 2013;10(5):256–266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Taylor BS, Schultz N, Hieronymus H, et al. Integrative genomic profiling of human prostate cancer. Cancer Cell. 2010;18(1):11–22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Robinson D, Van Allen EM, Wu YM, et al. Integrative clinical genomics of advanced prostate cancer. Cell. 2015;161(5):1215–1228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Kumar A, Coleman I, Morrissey C, et al. Substantial interindividual and limited intraindividual genomic diversity among tumors from men with metastatic prostate cancer. Nat Med. 2016;22(4):369–378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Sharma SV, Lee DY, Li B, et al. A chromatin-mediated reversible drug-tolerant state in cancer cell subpopulations. Cell. 2010;141(1):69–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Wang LG, Johnson EM, Kinoshita Y, et al. Androgen receptor overexpression in prostate cancer linked to Pur alpha loss from a novel repressor complex. Cancer Res. 2008;68(8):2678–2688. [DOI] [PubMed] [Google Scholar]
  • 209.Liu X, Gomez-Pinillos A, Liu X, Johnson EM, Ferrari AC. Induction of bicalutamide sensitivity in prostate cancer cells by an epigenetic Puralpha-mediated decrease in androgen receptor levels. Prostate. 2010;70(2):179–189. [DOI] [PubMed] [Google Scholar]
  • 210.Ferrari AC, Alumkal JJ, Stein MN, et al. Epigenetic Therapy with Panobinostat Combined with Bicalutamide Rechallenge in Castration-Resistant Prostate Cancer. Clin Cancer Res. 2018. [DOI] [PubMed] [Google Scholar]
  • 211.Krug U, Buchner T, Berdel WE, Muller-Tidow C. The treatment of elderly patients with acute myeloid leukemia. Dtsch Arztebl Int. 2011;108(51–52):863–870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Pettit K, Odenike O. Defining and Treating Older Adults with Acute Myeloid Leukemia Who Are Ineligible for Intensive Therapies. Front Oncol. 2015;5:280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Kantarjian H, O’Brien S, Cortes J, et al. Results of intensive chemotherapy in 998 patients age 65 years or older with acute myeloid leukemia or high-risk myelodysplastic syndrome: predictive prognostic models for outcome. Cancer. 2006;106(5):1090–1098. [DOI] [PubMed] [Google Scholar]
  • 214.Kantarjian H, Ravandi F, O’Brien S, et al. Intensive chemotherapy does not benefit most older patients (age 70 years or older) with acute myeloid leukemia. Blood. 2010;116(22):4422–4429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Lagadinou ED, Sach A, Callahan K, et al. BCL-2 inhibition targets oxidative phosphorylation and selectively eradicates quiescent human leukemia stem cells. Cell Stem Cell. 2013;12(3):329–341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Pan R, Ruvolo VR, Wei J, et al. Inhibition of Mcl-1 with the pan-Bcl-2 family inhibitor (−)BI97D6 overcomes ABT-737 resistance in acute myeloid leukemia. Blood. 2015;126(3):363–372. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Konopleva M, Pollyea DA, Potluri J, et al. Efficacy and Biological Correlates of Response in a Phase II Study of Venetoclax Monotherapy in Patients with Acute Myelogenous Leukemia. Cancer Discov. 2016;6(10):1106–1117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.DiNardo CD, Pratz KW, Letai A, et al. Safety and preliminary efficacy of venetoclax with decitabine or azacitidine in elderly patients with previously untreated acute myeloid leukaemia: a non-randomised, open-label, phase 1b study. Lancet Oncol. 2018;19(2):216–228. [DOI] [PubMed] [Google Scholar]
  • 219.Cornely OA, Maertens J, Winston DJ, et al. Posaconazole vs. fluconazole or itraconazole prophylaxis in patients with neutropenia. N Engl J Med. 2007;356(4):348–359. [DOI] [PubMed] [Google Scholar]
  • 220.Wang DF, Helquist P, Wiech NL, Wiest O. Toward selective histone deacetylase inhibitor design: homology modeling, docking studies, and molecular dynamics simulations of human class I histone deacetylases. J Med Chem. 2005;48(22):6936–6947. [DOI] [PubMed] [Google Scholar]
  • 221.Butler KV, Kozikowski AP. Chemical origins of isoform selectivity in histone deacetylase inhibitors. Curr Pharm Des. 2008;14(6):505–528. [DOI] [PubMed] [Google Scholar]
  • 222.Boran AD, Iyengar R. Systems approaches to polypharmacology and drug discovery. Curr Opin Drug Discov Devel. 2010;13(3):297–309. [PMC free article] [PubMed] [Google Scholar]
  • 223.Knight ZA, Lin H, Shokat KM. Targeting the cancer kinome through polypharmacology. Nat Rev Cancer. 2010;10(2):130–137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Rask-Andersen M, Zhang J, Fabbro D, Schioth HB. Advances in kinase targeting: current clinical use and clinical trials. Trends Pharmacol Sci. 2014;35(11):604–620. [DOI] [PubMed] [Google Scholar]
  • 225.Daub H Quantitative proteomics of kinase inhibitor targets and mechanisms. ACS Chem Biol. 2015;10(1):201–212. [DOI] [PubMed] [Google Scholar]
  • 226.Cai X, Zhai HX, Wang J, et al. Discovery of 7-(4-(3-ethynylphenylamino)-7-methoxyquinazolin-6-yloxy)-N-hydroxyheptanamide (CUDc-101) as a potent multi-acting HDAC, EGFR, and HER2 inhibitor for the treatment of cancer. J Med Chem. 2010;53(5):2000–2009. [DOI] [PubMed] [Google Scholar]
  • 227.Lai CJ, Bao R, Tao X, et al. CUDC-101, a multitargeted inhibitor of histone deacetylase, epidermal growth factor receptor, and human epidermal growth factor receptor 2, exerts potent anticancer activity. Cancer Res. 2010;70(9):3647–3656. [DOI] [PubMed] [Google Scholar]
  • 228.Zhang L, Zhang Y, Mehta A, et al. Dual inhibition of HDAC and EGFR signaling with CUDC-101 induces potent suppression of tumor growth and metastasis in anaplastic thyroid cancer. Oncotarget. 2015;6(11):9073–9085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Galloway TJ, Wirth LJ, Colevas AD, et al. A Phase I Study of CUDC-101, a Multitarget Inhibitor of HDACs, EGFR, and HER2, in Combination with Chemoradiation in Patients with Head and Neck Squamous Cell Carcinoma. Clin Cancer Res. 2015;21(7):1566–1573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Shimizu T, LoRusso PM, Papadopoulos KP, et al. Phase I first-in-human study of CUDC-101, a multitargeted inhibitor of HDACs, EGFR, and HER2 in patients with advanced solid tumors. Clin Cancer Res. 2014;20(19):5032–5040. [DOI] [PubMed] [Google Scholar]
  • 231.Courtney KD, Corcoran RB, Engelman JA. The PI3K pathway as drug target in human cancer. J Clin Oncol. 2010;28(6):1075–1083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Chen Y, Peubez C, Smith V, et al. CUDC-907 blocks multiple pro-survival signals and abrogates microenvironment protection in CLL. J Cell Mol Med. 2019;23(1):340–348. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Sun K, Atoyan R, Borek MA, et al. Dual HDAC and PI3K Inhibitor CUDC-907 Downregulates MYC and Suppresses Growth of MYC-dependent Cancers. Mol Cancer Ther. 2017;16(2):285–299. [DOI] [PubMed] [Google Scholar]
  • 234.Oki Y, Kelly KR, Flinn I, et al. CUDC-907 in relapsed/refractory diffuse large B-cell lymphoma, including patients with MYC-alterations: results from an expanded phase I trial. Haematologica. 2017;102(11):1923–1930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Knight T, Qiao X, Edwards H, Lin H, Taub JW, Ge Y. Novel Therapy for FLT3-ITD Acute Myeloid Leukemia Utilizing the Combination of CUDC-907 and Gilteritinib. Blood. 2018;132(Suppl 1):1427. [Google Scholar]
  • 236.Qian C, Lai CJ, Bao R, et al. Cancer network disruption by a single molecule inhibitor targeting both histone deacetylase activity and phosphatidylinositol 3-kinase signaling. Clin Cancer Res. 2012;18(15):4104–4113. [DOI] [PubMed] [Google Scholar]
  • 237.Development History and FDA Approval Process for Nexavar Drugs.com, 2007 https://www.drugs.com/history/nexavar.html
  • 238.Lencioni R, Llovet JM. Modified RECIST (mRECIST) assessment for hepatocellular carcinoma. Semin Liver Dis. 2010;30(1):52–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.Development History and FDA Approval Process for Stivarga Drugs.com https://www.drugs.com/history/stivarga.html 2017.
  • 240.Llovet JM. Regorafenib in Second-Line Setting for Hepatocellular Carcinoma: Balancing Benefit With Toxicity. February 25, 2017. https://www.ascopost.com/issues/february-25-2017/regorafenib-in-second-line-setting-for-hepatocellular-carcinoma-balancing-benefit-with-toxicity/
  • 241.Zhu AX, Kudo M, Assenat E, et al. Effect of everolimus on survival in advanced hepatocellular carcinoma after failure of sorafenib: the EVOLVE-1 randomized clinical trial. JAMA. 2014;312(1):57–67. [DOI] [PubMed] [Google Scholar]
  • 242.Chen D, Soh CK, Goh WH, Wang H. Design, Synthesis, and Preclinical Evaluation of Fused Pyrimidine-Based Hydroxamates for the Treatment of Hepatocellular Carcinoma. J Med Chem. 2018;61(4):1552–1575. [DOI] [PubMed] [Google Scholar]
  • 243.Poulsen A, Nagaraj H, Lee A, et al. Structure and ligand-based design of mTOR and PI3-kinase inhibitors leading to the clinical candidates VS-5584 (SB2343) and SB2602. J Chem Inf Model. 2014;54(11):3238–3250. [DOI] [PubMed] [Google Scholar]
  • 244.Ota S, Tonou-Fujimori N, Yamasu K. The roles of the FGF signal in zebrafish embryos analyzed using constitutive activation and dominant-negative suppression of different FGF receptors. Mech Dev. 2009;126(1–2):1–17. [DOI] [PubMed] [Google Scholar]
  • 245.Trueb B Biology of FGFRL1, the fifth fibroblast growth factor receptor. Cell Mol Life Sci. 2011;68(6):951–964. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.Brooks AN, Kilgour E, Smith PD. Molecular pathways: fibroblast growth factor signaling: a new therapeutic opportunity in cancer. Clin Cancer Res. 2012;18(7):1855–1862. [DOI] [PubMed] [Google Scholar]
  • 247.Neuzillet Y, van Rhijn BW, Prigoda NL, et al. FGFR3 mutations, but not FGFR3 expression and FGFR3 copy-number variations, are associated with favourable non-muscle invasive bladder cancer. Virchows Arch. 2014;465(2):207–213. [DOI] [PubMed] [Google Scholar]
  • 248.Taylor JGt, Cheuk AT, Tsang PS, et al. Identification of FGFR4-activating mutations in human rhabdomyosarcomas that promote metastasis in xenotransplanted models. J Clin Invest. 2009;119(11):3395–3407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Laird AD, Vajkoczy P, Shawver LK, et al. SU6668 is a potent antiangiogenic and antitumor agent that induces regression of established tumors. Cancer Res. 2000;60(15):4152–4160. [PubMed] [Google Scholar]
  • 250.Hilberg F, Roth GJ, Krssak M, et al. BIBF 1120: triple angiokinase inhibitor with sustained receptor blockade and good antitumor efficacy. Cancer Res. 2008;68(12):4774–4782. [DOI] [PubMed] [Google Scholar]
  • 251.De Luca A, Frezzetti D, Gallo M, Normanno N. FGFR-targeted therapeutics for the treatment of breast cancer. Expert Opin Investig Drugs. 2017;26(3):303–311. [DOI] [PubMed] [Google Scholar]
  • 252.Babina IS, Turner NC. Advances and challenges in targeting FGFR signalling in cancer. Nat Rev Cancer. 2017;17(5):318–332. [DOI] [PubMed] [Google Scholar]
  • 253.Putcha P, Yu J, Rodriguez-Barrueco R, et al. HDAC6 activity is a non-oncogene addiction hub for inflammatory breast cancers. Breast Cancer Res. 2015;17(1):149. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Yu S, Cai X, Wu C, et al. Targeting HSP90-HDAC6 Regulating Network Implicates Precision Treatment of Breast Cancer. Int J Biol Sci. 2017;13(4):505–517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Liu J, Peng X, Dai Y, et al. Design, synthesis and biological evaluation of novel FGFR inhibitors bearing an indazole scaffold. Org Biomol Chem. 2015;13(28):7643–7654. [DOI] [PubMed] [Google Scholar]
  • 256.Liu J, Qian C, Zhu Y, et al. Design, synthesis and evaluate of novel dual FGFR1 and HDAC inhibitors bearing an indazole scaffold. Bioorg Med Chem. 2018;26(3):747–757. [DOI] [PubMed] [Google Scholar]
  • 257.Bergman JA, Woan K, Perez-Villarroel P, Villagra A, Sotomayor EM, Kozikowski AP. Selective histone deacetylase 6 inhibitors bearing substituted urea linkers inhibit melanoma cell growth. J Med Chem. 2012;55(22):9891–9899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Baker SJ, Reddy EP. CDK4: A Key Player in the Cell Cycle, Development, and Cancer. Genes Cancer. 2012;3(11–12):658–669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Dickson MA. Molecular pathways: CDK4 inhibitors for cancer therapy. Clin Cancer Res. 2014;20(13):3379–3383. [DOI] [PubMed] [Google Scholar]
  • 260.Lim JS, Turner NC, Yap TA. CDK4/6 Inhibitors: Promising Opportunities beyond Breast Cancer. Cancer Discov. 2016;6(7):697–699. [DOI] [PubMed] [Google Scholar]
  • 261.O’Leary B, Finn RS, Turner NC. Treating cancer with selective CDK4/6 inhibitors. Nat Rev Clin Oncol. 2016;13(7):417–430. [DOI] [PubMed] [Google Scholar]
  • 262.Sherr CJ, Beach D, Shapiro GI. Targeting CDK4 and CDK6: From Discovery to Therapy. Cancer Discov. 2016;6(4):353–367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Di Giovanni C, Novellino E, Chilin A, Lavecchia A, Marzaro G. Investigational drugs targeting cyclin-dependent kinases for the treatment of cancer: an update on recent findings (2013–2016). Expert Opin Investig Drugs. 2016;25(10):1215–1230. [DOI] [PubMed] [Google Scholar]
  • 264.Hamilton E, Infante JR. Targeting CDK4/6 in patients with cancer. Cancer Treat Rev. 2016;45:129–138. [DOI] [PubMed] [Google Scholar]
  • 265.Lu H, Chang DJ, Baratte B, Meijer L, Schulze-Gahmen U. Crystal structure of a human cyclin-dependent kinase 6 complex with a flavonol inhibitor, fisetin. J Med Chem. 2005;48(3):737–743. [DOI] [PubMed] [Google Scholar]
  • 266.Sanchez-Martinez C, Gelbert LM, Lallena MJ, de Dios A. Cyclin dependent kinase (CDK) inhibitors as anticancer drugs. Bioorg Med Chem Lett. 2015;25(17):3420–3435. [DOI] [PubMed] [Google Scholar]
  • 267.Rader J, Russell MR, Hart LS, et al. Dual CDK4/CDK6 inhibition induces cell-cycle arrest and senescence in neuroblastoma. Clin Cancer Res. 2013;19(22):6173–6182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Logan JE, Mostofizadeh N, Desai AJ, et al. PD-0332991, a potent and selective inhibitor of cyclin-dependent kinase 4/6, demonstrates inhibition of proliferation in renal cell carcinoma at nanomolar concentrations and molecular markers predict for sensitivity. Anticancer Res. 2013;33(8):2997–3004. [PubMed] [Google Scholar]
  • 269.Finn RS, Dering J, Conklin D, et al. PD 0332991, a selective cyclin D kinase 4/6 inhibitor, preferentially inhibits proliferation of luminal estrogen receptor-positive human breast cancer cell lines in vitro. Breast Cancer Res. 2009;11(5):R77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Toogood PL, Harvey PJ, Repine JT, et al. Discovery of a potent and selective inhibitor of cyclin-dependent kinase 4/6. J Med Chem. 2005;48(7):2388–2406. [DOI] [PubMed] [Google Scholar]
  • 271.Baughn LB, Di Liberto M, Wu K, et al. A novel orally active small molecule potently induces G1 arrest in primary myeloma cells and prevents tumor growth by specific inhibition of cyclin-dependent kinase 4/6. Cancer Res. 2006;66(15):7661–7667. [DOI] [PubMed] [Google Scholar]
  • 272.Reddy MV, Akula B, Cosenza SC, et al. Discovery of 8-cyclopentyl-2-[4-(4-methyl-piperazin-1-yl)-phenylamino]-7-oxo-7,8-dihydro-pyrid o[2,3-d]pyrimidine-6-carbonitrile (7x) as a potent inhibitor of cyclin-dependent kinase 4 (CDK4) and AMPK-related kinase 5 (ARK5). J Med Chem. 2014;57(3):578–599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Heilmann AM, Perera RM, Ecker V, et al. CDK4/6 and IGF1 receptor inhibitors synergize to suppress the growth of p16INK4A-deficient pancreatic cancers. Cancer Res. 2014;74(14):3947–3958. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.Huang JM, Sheard MA, Ji L, Sposto R, Keshelava N. Combination of vorinostat and flavopiridol is selectively cytotoxic to multidrug-resistant neuroblastoma cell lines with mutant TP53. Mol Cancer Ther. 2010;9(12):3289–3301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 275.Li Y, Luo X, Guo Q, et al. Discovery of N1-(4-((7-Cyclopentyl-6-(dimethylcarbamoyl)-7 H-pyrrolo[2,3- d]pyrimidin-2-yl)amino)phenyl)- N8-hydroxyoctanediamide as a Novel Inhibitor Targeting Cyclin-dependent Kinase 4/9 (CDK4/9) and Histone Deacetlyase1 (HDAC1) against Malignant Cancer. J Med Chem. 2018;61(7):3166–3192. [DOI] [PubMed] [Google Scholar]
  • 276.Patnaik A, Rosen LS, Tolaney SM, et al. Efficacy and Safety of Abemaciclib, an Inhibitor of CDK4 and CDK6, for Patients with Breast Cancer, Non-Small Cell Lung Cancer, and Other Solid Tumors. Cancer Discov. 2016;6(7):740–753. [DOI] [PubMed] [Google Scholar]
  • 277.Huwe A, Mazitschek R, Giannis A. Small molecules as inhibitors of cyclin-dependent kinases. Angew Chem Int Ed Engl. 2003;42(19):2122–2138. [DOI] [PubMed] [Google Scholar]
  • 278.Huang Z, Zhou W, Li Y, et al. Novel hybrid molecule overcomes the limited response of solid tumours to HDAC inhibitors via suppressing JAK1-STAT3-BCL2 signalling. Theranostics. 2018;8(18):4995–5011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 279.Peng FW, Xuan J, Wu TT, et al. Design, synthesis and biological evaluation of N-phenylquinazolin-4-amine hybrids as dual inhibitors of VEGFR-2 and HDAC. Eur J Med Chem. 2016;109:1–12. [DOI] [PubMed] [Google Scholar]
  • 280.Harris PA, Boloor A, Cheung M, et al. Discovery of 5-[[4-[(2,3-dimethyl-2H-indazol-6-yl)methylamino]-2-pyrimidinyl]amino]-2-methylbenzenesulfonamide (pazopanib), a novel and potent vascular endothelial growth factor receptor inhibitor. J Med Chem. 2008;51(15):4632–4640. [DOI] [PubMed] [Google Scholar]
  • 281.Brotelle T, Bay JO. Pazopanib for treatment of renal cell carcinoma and soft tissue sarcomas. Bulletin Du Cancer. 2014;101(6):641–646. [DOI] [PubMed] [Google Scholar]
  • 282.Zang J, Liang XW, Huang YX, et al. Discovery of Novel Pazopanib-Based HDAC and VEGFR Dual Inhibitors Targeting Cancer Epigenetics and Angiogenesis Simultaneously. J Med Chem. 2018;61(12):5304–5322. [DOI] [PubMed] [Google Scholar]
  • 283.Hurley LH. DNA and its associated processes as targets for cancer therapy. Nat Rev Cancer. 2002;2(3):188–200. [DOI] [PubMed] [Google Scholar]
  • 284.Griffith D, Morgan MP, Marmion CJ. A novel anti-cancer bifunctional platinum drug candidate with dual DNA binding and histone deacetylase inhibitory activity. Chem Commun (Camb) 2009(44):6735–6737. [DOI] [PubMed] [Google Scholar]
  • 285.Liu C, Ding HY, Li XX, et al. A DNA/HDAC dual-targeting drug CY190602 with significantly enhanced anticancer potency. Embo Mol Med. 2015;7(4):438–449. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Guerrant W, Patil V, Canzoneri JC, Oyelere AK. Dual targeting of histone deacetylase and topoisomerase II with novel bifunctional inhibitors. J Med Chem. 2012;55(4):1465–1477. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Yao L, Ohlson S, Dymock BW. Design and synthesis of triple inhibitors of janus kinase (JAK), histone deacetylase (HDAC) and Heat Shock Protein 90 (HSP90). Bioorg Med Chem Lett. 2018;28(8):1357–1362. [DOI] [PubMed] [Google Scholar]
  • 288.Rawlings JS, Rosler KM, Harrison DA. The JAK/STAT signaling pathway. J Cell Sci. 2004;117(Pt 8):1281–1283. [DOI] [PubMed] [Google Scholar]
  • 289.Bohonowych JE, Hance MW, Nolan KD, Defee M, Parsons CH, Isaacs JS. Extracellular Hsp90 mediates an NF-kappaB dependent inflammatory stromal program: implications for the prostate tumor microenvironment. Prostate. 2014;74(4):395–407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 290.Kekatpure VD, Dannenberg AJ, Subbaramaiah K. HDAC6 modulates Hsp90 chaperone activity and regulates activation of aryl hydrocarbon receptor signaling. J Biol Chem. 2009;284(12):7436–7445. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 291.Mesa RA, Yasothan U, Kirkpatrick P. Ruxolitinib. Nat Rev Drug Discov. 2012;11(2):103–104. [DOI] [PubMed] [Google Scholar]
  • 292.Alicea-Velazquez NL, Boggon TJ. The use of structural biology in Janus kinase targeted drug discovery. Curr Drug Targets. 2011;12(4):546–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Massey AJ, Schoepfer J, Brough PA, et al. Preclinical antitumor activity of the orally available heat shock protein 90 inhibitor NVP-BEP800. Mol Cancer Ther. 2010;9(4):906–919. [DOI] [PubMed] [Google Scholar]
  • 294.Brough PA, Barril X, Borgognoni J, et al. Combining hit identification strategies: fragment-based and in silico approaches to orally active 2-aminothieno[2,3-d]pyrimidine inhibitors of the Hsp90 molecular chaperone. J Med Chem. 2009;52(15):4794–4809. [DOI] [PubMed] [Google Scholar]
  • 295.Li ZR, Suo FZ, Hu B, et al. Identification of osimertinib (AZD9291) as a lysine specific demethylase 1 inhibitor. Bioorg Chem. 2018;84:164–169. [DOI] [PubMed] [Google Scholar]
  • 296.Smith CC, Wang Q, Chin CS, et al. Validation of ITD mutations in FLT3 as a therapeutic target in human acute myeloid leukaemia. Nature. 2012;485(7397):260–263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 297.Li J, Favata M, Kelley JA, et al. INCB16562, a JAK1/2 Selective Inhibitor, Is Efficacious against Multiple Myeloma Cells and Reverses the Protective Effects of Cytokine and Stromal Cell Support. Neoplasia. 2010;12(1):28–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Chao Q, Sprankle KG, Grotzfeld RM, et al. Identification of N-(5-tert-butyl-isoxazol-3-yl)-N’-{4-[7-(2-morpholin-4-yl-ethoxy)imidazo[2,1-b][1,3]benzothiazol-2-yl]phenyl}urea dihydrochloride (AC220), a uniquely potent, selective, and efficacious FMS-like tyrosine kinase-3 (FLT3) inhibitor. J Med Chem. 2009;52(23):7808–7816. [DOI] [PubMed] [Google Scholar]
  • 299.Fiskus W, Sharma S, Qi J, et al. BET protein antagonist JQ1 is synergistically lethal with FLT3 tyrosine kinase inhibitor (TKI) and overcomes resistance to FLT3-TKI in AML cells expressing FLT-ITD. Mol Cancer Ther. 2014;13(10):2315–2327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 300.Elkins JM, Fedele V, Szklarz M, et al. Comprehensive characterization of the Published Kinase Inhibitor Set. Nat Biotechnol. 2016;34(1):95–103. [DOI] [PubMed] [Google Scholar]
  • 301.Ember SW, Zhu JY, Olesen SH, et al. Acetyl-lysine binding site of bromodomain-containing protein 4 (BRD4) interacts with diverse kinase inhibitors. ACS Chem Biol. 2014;9(5):1160–1171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 302.Urick AK, Hawk LM, Cassel MK, et al. Dual Screening of BPTF and Brd4 Using Protein-Observed Fluorine NMR Uncovers New Bromodomain Probe Molecules. ACS Chem Biol. 2015;10(10):2246–2256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 303.Divakaran A, Talluri SK, Ayoub AM, et al. Molecular Basis for the N-Terminal Bromodomain-and-Extra-Terminal-Family Selectivity of a Dual Kinase-Bromodomain Inhibitor. J Med Chem. 2018;61(20):9316–9334. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 304.Mai A, Rotili D, Tarantino D, et al. Small-molecule inhibitors of histone acetyltransferase activity: identification and biological properties. J Med Chem. 2006;49(23):6897–6907. [DOI] [PubMed] [Google Scholar]
  • 305.De Bellis F, Carafa V, Conte M, et al. Context-selective death of acute myeloid leukemia cells triggered by the novel hybrid retinoid-HDAC inhibitor MC2392. Cancer Res. 2014;74(8):2328–2339. [DOI] [PubMed] [Google Scholar]
  • 306.Sharma P, Allison JP. The future of immune checkpoint therapy. Science. 2015;348(6230):56–61. [DOI] [PubMed] [Google Scholar]
  • 307.Dounay AB, Tuttle JB, Verhoest PR. Challenges and Opportunities in the Discovery of New Therapeutics Targeting the Kynurenine Pathway. J Med Chem. 2015;58(22):8762–8782. [DOI] [PubMed] [Google Scholar]
  • 308.Macchiarulo A, Camaioni E, Nuti R, Pellicciari R. Highlights at the gate of tryptophan catabolism: a review on the mechanisms of activation and regulation of indoleamine 2,3-dioxygenase (IDO), a novel target in cancer disease. Amino Acids. 2009;37(2):219–229. [DOI] [PubMed] [Google Scholar]
  • 309.Curti A, Trabanelli S, Salvestrini V, Baccarani M, Lemoli RM. The role of indoleamine 2,3-dioxygenase in the induction of immune tolerance: focus on hematology. Blood. 2009;113(11):2394–2401. [DOI] [PubMed] [Google Scholar]
  • 310.Munn DH, Sharma MD, Baban B, et al. GCN2 kinase in T cells mediates proliferative arrest and anergy induction in response to indoleamine 2,3-dioxygenase. Immunity. 2005;22(5):633–642. [DOI] [PubMed] [Google Scholar]
  • 311.Uyttenhove C, Pilotte L, Theate I, et al. Evidence for a tumoral immune resistance mechanism based on tryptophan degradation by indoleamine 2,3-dioxygenase. Nat Med. 2003;9(10):1269–1274. [DOI] [PubMed] [Google Scholar]
  • 312.Godin-Ethier J, Hanafi LA, Piccirillo CA, Lapointe R. Indoleamine 2,3-dioxygenase expression in human cancers: clinical and immunologic perspectives. Clin Cancer Res. 2011;17(22):6985–6991. [DOI] [PubMed] [Google Scholar]
  • 313.Muller AJ, DuHadaway JB, Donover PS, Sutanto-Ward E, Prendergast GC. Inhibition of indoleamine 2,3-dioxygenase, an immunoregulatory target of the cancer suppression gene Bin1, potentiates cancer chemotherapy. Nat Med. 2005;11(3):312–319. [DOI] [PubMed] [Google Scholar]
  • 314.Li M, Bolduc AR, Hoda MN, et al. The indoleamine 2,3-dioxygenase pathway controls complement-dependent enhancement of chemo-radiation therapy against murine glioblastoma. J Immunother Cancer. 2014;2:21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 315.Dunn J, Rao S. Epigenetics and immunotherapy: The current state of play. Mol Immunol. 2017;87:227–239. [DOI] [PubMed] [Google Scholar]
  • 316.Booth L, Roberts JL, Poklepovic A, Kirkwood J, Dent P. HDAC inhibitors enhance the immunotherapy response of melanoma cells. Oncotarget. 2017;8(47):83155–83170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 317.Yue EW, Sparks R, Polam P, et al. INCB24360 (Epacadostat), a Highly Potent and Selective Indoleamine-2,3-dioxygenase 1 (IDO1) Inhibitor for Immuno-oncology. ACS Med Chem Lett. 2017;8(5):486–491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 318.Fang K, Dong G, Li Y, et al. Discovery of Novel Indoleamine 2,3-Dioxygenase 1 (IDO1) and Histone Deacetylase (HDAC) Dual Inhibitors. ACS Med Chem Lett. 2018;9(4):312–317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 319.Sampath D, Zabka TS, Misner DL, O’Brien T, Dragovich PS. Inhibition of nicotinamide phosphoribosyltransferase (NAMPT) as a therapeutic strategy in cancer. Pharmacol Ther. 2015;151:16–31. [DOI] [PubMed] [Google Scholar]
  • 320.Galli U, Travelli C, Massarotti A, et al. Medicinal chemistry of nicotinamide phosphoribosyltransferase (NAMPT) inhibitors. J Med Chem. 2013;56(16):6279–6296. [DOI] [PubMed] [Google Scholar]
  • 321.Watson M, Roulston A, Belec L, et al. The small molecule GMX1778 is a potent inhibitor of NAD+ biosynthesis: strategy for enhanced therapy in nicotinic acid phosphoribosyltransferase 1-deficient tumors. Mol Cell Biol. 2009;29(21):5872–5888. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 322.Hasmann M, Schemainda I. FK866, a highly specific noncompetitive inhibitor of nicotinamide phosphoribosyltransferase, represents a novel mechanism for induction of tumor cell apoptosis. Cancer Res. 2003;63(21):7436–7442. [PubMed] [Google Scholar]
  • 323.Holen K, Saltz LB, Hollywood E, Burk K, Hanauske AR. The pharmacokinetics, toxicities, and biologic effects of FK866, a nicotinamide adenine dinucleotide biosynthesis inhibitor. Invest New Drugs. 2008;26(1):45–51. [DOI] [PubMed] [Google Scholar]
  • 324.von Heideman A, Berglund A, Larsson R, Nygren P. Safety and efficacy of NAD depleting cancer drugs: results of a phase I clinical trial of CHS 828 and overview of published data. Cancer Chemother Pharmacol. 2010;65(6):1165–1172. [DOI] [PubMed] [Google Scholar]
  • 325.Chen W, Dong G, Wu Y, Zhang W, Miao C, Sheng C. Dual NAMPT/HDAC Inhibitors as a New Strategy for Multitargeting Antitumor Drug Discovery. ACS Med Chem Lett. 2018;9(1):34–38. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 326.Brazeau JF, Rosse G. Thiazolocarboxamide Analogues as NAMPT Inhibitors. ACS Med Chem Lett. 2014;5(4):277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 327.Dong G, Chen W, Wang X, et al. Small Molecule Inhibitors Simultaneously Targeting Cancer Metabolism and Epigenetics: Discovery of Novel Nicotinamide Phosphoribosyltransferase (NAMPT) and Histone Deacetylase (HDAC) Dual Inhibitors. J Med Chem. 2017;60(19):7965–7983. [DOI] [PubMed] [Google Scholar]
  • 328.Aranha O, Wood DP Jr., Sarkar FH. Ciprofloxacin mediated cell growth inhibition, S/G2-M cell cycle arrest, and apoptosis in a human transitional cell carcinoma of the bladder cell line. Clin Cancer Res. 2000;6(3):891–900. [PubMed] [Google Scholar]
  • 329.Herold C, Ocker M, Ganslmayer M, Gerauer H, Hahn EG, Schuppan D. Ciprofloxacin induces apoptosis and inhibits proliferation of human colorectal carcinoma cells. Br J Cancer. 2002;86(3):443–448. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 330.Wang G, Peng Z, Peng S, Qiu J, Li Y, Lan Y. (E)-N-Aryl-2-oxo-2-(3,4,5-trimethoxyphenyl)acetohydrazonoyl cyanides as tubulin polymerization inhibitors: Structure-based bioisosterism design, synthesis, biological evaluation, molecular docking and in silico ADME prediction. Bioorg Med Chem Lett. 2018;28(20):3350–3355. [DOI] [PubMed] [Google Scholar]
  • 331.Xia Y, Yang ZY, Xia P, et al. Antitumor Agents. 211. Fluorinated 2-phenyl-4-quinolone derivatives as antimitotic antitumor agents. J Med Chem. 2001;44(23):3932–3936. [DOI] [PubMed] [Google Scholar]
  • 332.Janin YL. Antituberculosis drugs: ten years of research. Bioorg Med Chem. 2007;15(7):2479–2513. [DOI] [PubMed] [Google Scholar]
  • 333.Xuan Wang XJ, Shiyou Sun and Yongqiong Liu. Synthesis and biological evaluation of novel quinolone derivatives dual targeting histone deacetylase and tubulin polymerization as antiproliferative agents. RSC Adv. 2018;8:16494–16502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 334.Chao MW, Lai MJ, Liou JP, et al. The synergic effect of vincristine and vorinostat in leukemia in vitro and in vivo. J Hematol Oncol. 2015;8:82. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 335.Lamaa D, Lin HP, Zig L, et al. Design and Synthesis of Tubulin and Histone Deacetylase Inhibitor Based on iso-Combretastatin A-4. J Med Chem. 2018;61(15):6574–6591. [DOI] [PubMed] [Google Scholar]
  • 336.Graham B, Curry J, Smyth T, et al. The heat shock protein 90 inhibitor, AT13387, displays a long duration of action in vitro and in vivo in non-small cell lung cancer. Cancer Sci. 2012;103(3):522–527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 337.Proia DA, Bates RC. Ganetespib and HSP90: translating preclinical hypotheses into clinical promise. Cancer Res. 2014;74(5):1294–1300. [DOI] [PubMed] [Google Scholar]
  • 338.Johnson ML, Yu HA, Hart EM, et al. Phase I/II Study of HSP90 Inhibitor AUY922 and Erlotinib for EGFR-Mutant Lung Cancer With Acquired Resistance to Epidermal Growth Factor Receptor Tyrosine Kinase Inhibitors. J Clin Oncol. 2015;33(15):1666–1673. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 339.Ojha R, Huang HL, HuangFu WC, et al. 1-Aroylindoline-hydroxamic acids as anticancer agents, inhibitors of HSP90 and HDAC. Eur J Med Chem. 2018;150:667–677. [DOI] [PubMed] [Google Scholar]
  • 340.Ugarte A, Gil-Bea F, Garcia-Barroso C, et al. Decreased levels of guanosine 3’, 5’-monophosphate (cGMP) in cerebrospinal fluid (CSF) are associated with cognitive decline and amyloid pathology in Alzheimer’s disease. Neuropathol Appl Neurobiol. 2015;41(4):471–482. [DOI] [PubMed] [Google Scholar]
  • 341.Cuadrado-Tejedor M, Hervias I, Ricobaraza A, et al. Sildenafil restores cognitive function without affecting beta-amyloid burden in a mouse model of Alzheimer’s disease. Br J Pharmacol. 2011;164(8):2029–2041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 342.Garcia-Barroso C, Ricobaraza A, Pascual-Lucas M, et al. Tadalafil crosses the blood-brain barrier and reverses cognitive dysfunction in a mouse model of AD. Neuropharmacology. 2013;64:114–123. [DOI] [PubMed] [Google Scholar]
  • 343.Reneerkens OA, Rutten K, Akkerman S, et al. Phosphodiesterase type 5 (PDE5) inhibition improves object recognition memory: indications for central and peripheral mechanisms. Neurobiol Learn Mem. 2012;97(4):370–379. [DOI] [PubMed] [Google Scholar]
  • 344.Rabal O, Sanchez-Arias JA, Cuadrado-Tejedor M, et al. Design, Synthesis, and Biological Evaluation of First-in-Class Dual Acting Histone Deacetylases (HDACs) and Phosphodiesterase 5 (PDE5) Inhibitors for the Treatment of Alzheimer’s Disease. J Med Chem. 2016;59(19):8967–9004. [DOI] [PubMed] [Google Scholar]
  • 345.Sanchez-Arias JA, Rabal O, Cuadrado-Tejedor M, et al. Impact of Scaffold Exploration on Novel Dual-Acting Histone Deacetylases and Phosphodiesterase 5 Inhibitors for the Treatment of Alzheimer’s Disease. ACS Chem Neurosci. 2017;8(3):638–661. [DOI] [PubMed] [Google Scholar]
  • 346.Cuadrado-Tejedor M, Garcia-Barroso C, Sanzhez-Arias J, et al. Concomitant histone deacetylase and phosphodiesterase 5 inhibition synergistically prevents the disruption in synaptic plasticity and it reverses cognitive impairment in a mouse model of Alzheimer’s disease. Clin Epigenetics. 2015;7:108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 347.Cuadrado-Tejedor M, Garcia-Barroso C, Sanchez-Arias JA, et al. A First-in-Class Small-Molecule that Acts as a Dual Inhibitor of HDAC and PDE5 and that Rescues Hippocampal Synaptic Impairment in Alzheimer’s Disease Mice. Neuropsychopharmacology. 2017;42(2):524–539. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 348.Rabal O, Sanchez-Arias JA, Cuadrado-Tejedor M, et al. Design, synthesis, biological evaluation and in vivo testing of dual phosphodiesterase 5 (PDE5) and histone deacetylase 6 (HDAC6)-selective inhibitors for the treatment of Alzheimer’s disease. Eur J Med Chem. 2018;150:506–524. [DOI] [PubMed] [Google Scholar]
  • 349.Borrell-Pages M, Canals JM, Cordelieres FP, et al. Cystamine and cysteamine increase brain levels of BDNF in Huntington disease via HSJ1b and transglutaminase. J Clin Invest. 2006;116(5):1410–1424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 350.Basso M, Chen HH, Tripathy D, et al. Designing Dual Transglutaminase 2/Histone Deacetylase Inhibitors Effective at Halting Neuronal Death. ChemMedChem. 2018;13(3):227–230. [DOI] [PubMed] [Google Scholar]
  • 351.Atlante S, Chegaev K, Cencioni C, et al. Structural and biological characterization of new hybrid drugs joining an HDAC inhibitor to different NO-donors. Eur J Med Chem. 2018;144:612–625. [DOI] [PubMed] [Google Scholar]
  • 352.Colussi C, Mozzetta C, Gurtner A, et al. HDAC2 blockade by nitric oxide and histone deacetylase inhibitors reveals a common target in Duchenne muscular dystrophy treatment. Proc Natl Acad Sci U S A. 2008;105(49):19183–19187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 353.Fischer ES, Bohm K, Lydeard JR, et al. Structure of the DDB1-CRBN E3 ubiquitin ligase in complex with thalidomide. Nature. 2014;512(7512):49–53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 354.Ito T, Ando H, Suzuki T, et al. Identification of a primary target of thalidomide teratogenicity. Science. 2010;327(5971):1345–1350. [DOI] [PubMed] [Google Scholar]
  • 355.Lopez-Girona A, Mendy D, Ito T, et al. Cereblon is a direct protein target for immunomodulatory and antiproliferative activities of lenalidomide and pomalidomide. Leukemia. 2012;26(11):2326–2335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 356.Winter GE, Buckley DL, Paulk J, et al. DRUG DEVELOPMENT. Phthalimide conjugation as a strategy for in vivo target protein degradation. Science. 2015;348(6241):1376–1381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 357.Lu J, Qian Y, Altieri M, et al. Hijacking the E3 Ubiquitin Ligase Cereblon to Efficiently Target BRD4. Chem Biol. 2015;22(6):755–763. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 358.Itoh Y Chemical Protein Degradation Approach and its Application to Epigenetic Targets. Chem Rec. 2018. [DOI] [PubMed] [Google Scholar]
  • 359.Wissmann M, Yin N, Muller JM, et al. Cooperative demethylation by JMJD2C and LSD1 promotes androgen receptor-dependent gene expression. Nat Cell Biol. 2007;9(3):347–353. [DOI] [PubMed] [Google Scholar]
  • 360.Rotili D, Tomassi S, Conte M, et al. Pan-histone demethylase inhibitors simultaneously targeting Jumonji C and lysine-specific demethylases display high anticancer activities. J Med Chem. 2014;57(1):42–55. [DOI] [PubMed] [Google Scholar]
  • 361.Rose NR, Ng SS, Mecinovic J, et al. Inhibitor scaffolds for 2-oxoglutarate-dependent histone lysine demethylases. J Med Chem. 2008;51(22):7053–7056. [DOI] [PubMed] [Google Scholar]
  • 362.Hopkinson RJ, Tumber A, Yapp C, et al. 5-Carboxy-8-hydroxyquinoline is a Broad Spectrum 2-Oxoglutarate Oxygenase Inhibitor which Causes Iron Translocation. Chem Sci. 2013;4(8):3110–3117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 363.Chen MW, Hua KT, Kao HJ, et al. H3K9 histone methyltransferase G9a promotes lung cancer invasion and metastasis by silencing the cell adhesion molecule Ep-CAM. Cancer Res. 2010;70(20):7830–7840. [DOI] [PubMed] [Google Scholar]
  • 364.Lehnertz B, Pabst C, Su L, et al. The methyltransferase G9a regulates HoxA9-dependent transcription in AML. Genes Dev. 2014;28(4):317–327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 365.Upadhyay AK, Rotili D, Han JW, et al. An analog of BIX-01294 selectively inhibits a family of histone H3 lysine 9 Jumonji demethylases. J Mol Biol. 2012;416(3):319–327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 366.Rotili D, Tarantino D, Marrocco B, et al. Properly substituted analogues of BIX-01294 lose inhibition of G9a histone methyltransferase and gain selective anti-DNA methyltransferase 3A activity. PLoS One. 2014;9(5):e96941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 367.Kubicek S, O’Sullivan RJ, August EM, et al. Reversal of H3K9me2 by a small-molecule inhibitor for the G9a histone methyltransferase. Mol Cell. 2007;25(3):473–481. [DOI] [PubMed] [Google Scholar]
  • 368.Vedadi M, Barsyte-Lovejoy D, Liu F, et al. A chemical probe selectively inhibits G9a and GLP methyltransferase activity in cells. Nat Chem Biol. 2011;7(8):566–574. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 369.Speranzini V, Rotili D, Ciossani G, et al. Polymyxins and quinazolines are LSD1/KDM1A inhibitors with unusual structural features. Sci Adv. 2016;2(9):e1601017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 370.Esteve PO, Chin HG, Smallwood A, et al. Direct interaction between DNMT1 and G9a coordinates DNA and histone methylation during replication. Genes Dev. 2006;20(22):3089–3103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 371.Wozniak RJ, Klimecki WT, Lau SS, Feinstein Y, Futscher BW. 5-Aza-2’-deoxycytidine -mediated reductions in G9A histone methyltransferase and histone H3 K9 di-methylation levels are linked to tumor suppressor gene reactivation. Oncogene. 2007;26(1):77–90. [DOI] [PubMed] [Google Scholar]
  • 372.Sharma S, Gerke DS, Han HF, et al. Lysine methyltransferase G9a is not required for DNMT3A/3B anchoring to methylated nucleosomes and maintenance of DNA methylation in somatic cells. Epigenetics Chromatin. 2012;5(1):3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 373.Rabal O, San Jose-Eneriz E, Agirre X, et al. Discovery of Reversible DNA Methyltransferase and Lysine Methyltransferase G9a Inhibitors with Antitumoral in Vivo Efficacy. J Med Chem. 2018;61(15):6518–6545. [DOI] [PubMed] [Google Scholar]
  • 374.Rabal O, Sanchez-Arias JA, San Jose-Eneriz E, et al. Detailed Exploration around 4-Aminoquinolines Chemical Space to Navigate the Lysine Methyltransferase G9a and DNA Methyltransferase Biological Spaces. J Med Chem. 2018;61(15):6546–6573. [DOI] [PubMed] [Google Scholar]
  • 375.San Jose-Eneriz E, Agirre X, Rabal O, et al. Discovery of first-in-class reversible dual small molecule inhibitors against G9a and DNMTs in hematological malignancies. Nat Commun. 2017;8:15424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 376.Dokmanovic M, Clarke C, Marks PA. Histone deacetylase inhibitors: overview and perspectives. Mol Cancer Res. 2007;5(10):981–989. [DOI] [PubMed] [Google Scholar]
  • 377.Inoue S, Riley J, Gant TW, Dyer MJS, Cohen GM. Apoptosis induced by histone deacetylase inhibitors in leukemic cells is mediated by Bim and Noxa. Leukemia. 2007;21(8):1773–1782. [DOI] [PubMed] [Google Scholar]
  • 378.Xu WS, Parmigiani RB, Marks PA. Histone deacetylase inhibitors: molecular mechanisms of action. Oncogene. 2007;26(37):5541–5552. [DOI] [PubMed] [Google Scholar]
  • 379.Zang L, Kondengaden SM, Zhang Q, et al. Structure based design, synthesis and activity studies of small hybrid molecules as HDAC and G9a dual inhibitors. Oncotarget. 2017;8(38):63187–63207. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 380.Kargbo RB. Histone Deacetylase Inhibitors as Treatment for Targeting Multiple Components in Cancer Therapy. ACS Med Chem Lett. 2018;9(3):167–168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 381.Ding J, Li T, Wang X, et al. The histone H3 methyltransferase G9A epigenetically activates the serine-glycine synthesis pathway to sustain cancer cell survival and proliferation. Cell Metab. 2013;18(6):896–907. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 382.Gibaja V, Shen F, Harari J, et al. Development of secondary mutations in wild-type and mutant EZH2 alleles cooperates to confer resistance to EZH2 inhibitors. Oncogene. 2016;35(5):558–566. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 383.Curry E, Green I, Chapman-Rothe N, et al. Dual EZH2 and EHMT2 histone methyltransferase inhibition increases biological efficacy in breast cancer cells. Clin Epigenetics. 2015;7:84. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 384.Reuter S, Gupta SC, Park B, Goel A, Aggarwal BB. Epigenetic changes induced by curcumin and other natural compounds. Genes Nutr. 2011;6(2):93–108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 385.Mai A, Cheng D, Bedford MT, et al. epigenetic multiple ligands: mixed histone/protein methyltransferase, acetyltransferase, and class III deacetylase (sirtuin) inhibitors. J Med Chem. 2008;51(7):2279–2290. [DOI] [PubMed] [Google Scholar]
  • 386.Pina IC, Gautschi JT, Wang GY, et al. Psammaplins from the sponge Pseudoceratina purpurea: inhibition of both histone deacetylase and DNA methyltransferase. J Org Chem. 2003;68(10):3866–3873. [DOI] [PubMed] [Google Scholar]
  • 387.Pereira R, Benedetti R, Perez-Rodriguez S, et al. Indole-Derived Psammaplin A Analogues as Epigenetic Modulators with Multiple Inhibitory Activities. J Med Chem. 2012;55(22):9467–9491. [DOI] [PubMed] [Google Scholar]
  • 388.Jones PA, Baylin SB. The epigenomics of cancer. Cell. 2007;128(4):683–692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 389.Robertson KD, Ait-Si-Ali S, Yokochi T, Wade PA, Jones PL, Wolffe AP. DNMT1 forms a complex with Rb, E2F1 and HDAC1 and represses transcription from E2F-responsive promoters. Nat Genet. 2000;25(3):338–342. [DOI] [PubMed] [Google Scholar]
  • 390.Fuks F, Burgers WA, Brehm A, Hughes-Davies L, Kouzarides T. DNA methyltransferase Dnmt1 associates with histone deacetylase activity. Nat Genet. 2000;24(1):88–91. [DOI] [PubMed] [Google Scholar]
  • 391.Rountree MR, Bachman KE, Baylin SB. DNMT1 binds HDAC2 and a new co-repressor, DMAP1, to form a complex at replication foci. Nat Genet. 2000;25(3):269–277. [DOI] [PubMed] [Google Scholar]
  • 392.Griffiths EA, Gore SD. DNA methyltransferase and histone deacetylase inhibitors in the treatment of myelodysplastic syndromes. Semin Hematol. 2008;45(1):23–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 393.Pathania R, Ramachandran S, Mariappan G, et al. Combined Inhibition of DNMT and HDAC Blocks the Tumorigenicity of Cancer Stem-like Cells and Attenuates Mammary Tumor Growth. Cancer Res. 2016;76(11):3224–3235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 394.Kuck D, Singh N, Lyko F, Medina-Franco JL. Novel and selective DNA methyltransferase inhibitors: Docking-based virtual screening and experimental evaluation. Bioorg Med Chem. 2010;18(2):822–829. [DOI] [PubMed] [Google Scholar]
  • 395.Yuan Z, Sun Q, Li D, et al. Design, synthesis and anticancer potential of NSC-319745 hydroxamic acid derivatives as DNMT and HDAC inhibitors. Eur J Med Chem. 2017;134:281–292. [DOI] [PubMed] [Google Scholar]
  • 396.Duan YC, Ma YC, Qin WP, et al. Design and synthesis of tranylcypromine derivatives as novel LSD1/HDACs dual inhibitors for cancer treatment. Eur J Med Chem. 2017;140:392–402. [DOI] [PubMed] [Google Scholar]
  • 397.Valente S, Rodriguez V, Mercurio C, et al. Pure Diastereomers of a Tranylcypromine-Based LSD1 Inhibitor: Enzyme Selectivity and In-Cell Studies. ACS Med Chem Lett. 2015;6(2):173–177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 398.Rodriguez V, Valente S, Rovida S, et al. Pyrrole- and indole-containing tranylcypromine derivatives as novel lysine-specific demethylase 1 inhibitors active on cancer cells. MedChemComm. 2015;6(4):665–670. [Google Scholar]
  • 399.Valente S, Rodriguez V, Mercurio C, et al. Pure enantiomers of benzoylamino-tranylcypromine: LSD1 inhibition, gene modulation in human leukemia cells and effects on clonogenic potential of murine promyelocytic blasts. Eur J Med Chem. 2015;94:163–174. [DOI] [PubMed] [Google Scholar]
  • 400.Vianello P, Botrugno OA, Cappa A, et al. Discovery of a Novel Inhibitor of Histone Lysine-Specific Demethylase 1A (KDM1A/LSD1) as Orally Active Antitumor Agent. J Med Chem. 2016;59(4):1501–1517. [DOI] [PubMed] [Google Scholar]
  • 401.Ourailidou ME, Lenoci A, Zwergel C, Rotili D, Mai A, Dekker FJ. Towards the development of activity-based probes for detection of lysine-specific demethylase-1 activity. Bioorg Med Chem. 2017;25(3):847–856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 402.Kalin JH, Wu M, Gomez AV, et al. Targeting the CoREST complex with dual histone deacetylase and demethylase inhibitors. Nat Commun. 2018;9(1):53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 403.Milelli A, Marchetti C, Turrini E, et al. Novel polyamine-based Histone deacetylases-Lysine demethylase 1 dual binding inhibitors. Bioorg Med Chem Lett. 2018;28(6):1001–1004. [DOI] [PubMed] [Google Scholar]
  • 404.Varghese S, Senanayake T, Murray-Stewart T, et al. Polyaminohydroxamic acids and polyaminobenzamides as isoform selective histone deacetylase inhibitors. J Med Chem. 2008;51(8):2447–2456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 405.Zhu Q, Huang Y, Marton LJ, Woster PM, Davidson NE, Casero RA Jr. Polyamine analogs modulate gene expression by inhibiting lysine-specific demethylase 1 (LSD1) and altering chromatin structure in human breast cancer cells. Amino Acids. 2012;42(2–3):887–898. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 406.Pasini A, Marchetti C, Sissi C, et al. Novel Polyamine-Naphthalene Diimide Conjugates Targeting Histone Deacetylases and DNA for Cancer Phenotype Reprogramming. ACS Med Chem Lett. 2017;8(12):1218–1223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 407.Palermo G, Minniti E, Greco ML, et al. An optimized polyamine moiety boosts the potency of human type II topoisomerase poisons as quantified by comparative analysis centered on the clinical candidate F14512. Chem Commun (Camb). 2015;51(76):14310–14313. [DOI] [PubMed] [Google Scholar]
  • 408.Minarini A, Milelli A, Tumiatti V, Rosini M, Bolognesi ML, Melchiorre C. Synthetic polyamines: an overview of their multiple biological activities. Amino Acids. 2010;38(2):383–392. [DOI] [PubMed] [Google Scholar]
  • 409.Bhadury J, Nilsson LM, Muralidharan SV, et al. BET and HDAC inhibitors induce similar genes and biological effects and synergize to kill in Myc-induced murine lymphoma. Proc Natl Acad Sci U S A. 2014;111(26):E2721–2730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 410.Chesi M, Matthews GM, Garbitt VM, et al. Drug response in a genetically engineered mouse model of multiple myeloma is predictive of clinical efficacy. Blood. 2012;120(2):376–385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 411.Picaud S, Strocchia M, Terracciano S, et al. 9H-purine scaffold reveals induced-fit pocket plasticity of the BRD9 bromodomain. J Med Chem. 2015;58(6):2718–2736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 412.Noguchi-Yachide T, Sakai T, Hashimoto Y, Yamaguchi T. Discovery and structure-activity relationship studies of N6-benzoyladenine derivatives as novel BRD4 inhibitors. Bioorg Med Chem. 2015;23(5):953–959. [DOI] [PubMed] [Google Scholar]
  • 413.Amemiya S, Yamaguchi T, Hashimoto Y, Noguchi-Yachide T. Synthesis and evaluation of novel dual BRD4/HDAC inhibitors. Bioorg Med Chem. 2017;25(14):3677–3684. [DOI] [PubMed] [Google Scholar]
  • 414.Hewings DS, Wang MH, Philpott M, et al. 3,5-Dimethylisoxazoles Act As Acetyl-lysine-mimetic Bromodomain Ligands. J Med Chem. 2011;54(19):6761–6770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 415.Zhang Z, Hou S, Chen H, et al. Targeting epigenetic reader and eraser: Rational design, synthesis and in vitro evaluation of dimethylisoxazoles derivatives as BRD4/HDAC dual inhibitors. Bioorg Med Chem Lett. 2016;26(12):2931–2935. [DOI] [PubMed] [Google Scholar]

RESOURCES