266
Views
53
CrossRef citations to date
0
Altmetric
Review

Histone deacetylase inhibitors (HDACIs): multitargeted anticancer agents

, , &
Pages 47-60 | Published online: 25 Feb 2013

Abstract

Histone deacetylase (HDAC) inhibitors are an emerging class of therapeutics with potential as anticancer drugs. The rationale for developing HDAC inhibitors (and other chromatin-modifying agents) as anticancer therapies arose from the understanding that in addition to genetic mutations, epigenetic changes such as dysregulation of HDAC enzymes can alter phenotype and gene expression, disturb homeostasis, and contribute to neoplastic growth. The family of HDAC inhibitors is large and diverse. It includes a range of naturally occurring and synthetic compounds that differ in terms of structure, function, and specificity. HDAC inhibitors have multiple cell type-specific effects in vitro and in vivo, such as growth arrest, cell differentiation, and apoptosis in malignant cells. HDAC inhibitors have the potential to be used as monotherapies or in combination with other anticancer therapies. Currently, there are two HDAC inhibitors that have received approval from the US FDA for the treatment of cutaneous T-cell lymphoma: vorinostat (suberoylanilide hydroxamic acid, Zolinza) and depsipeptide (romidepsin, Istodax). More recently, depsipeptide has also gained FDA approval for the treatment of peripheral T-cell lymphoma. Many more clinical trials assessing the effects of various HDAC inhibitors on hematological and solid malignancies are currently being conducted. Despite the proven anticancer effects of particular HDAC inhibitors against certain cancers, many aspects of HDAC enzymes and HDAC inhibitors are still not fully understood. Increasing our understanding of the effects of HDAC inhibitors, their targets and mechanisms of action will be critical for the advancement of these drugs, especially to facilitate the rational design of HDAC inhibitors that are effective as antineoplastic agents. This review will discuss the use of HDAC inhibitors as multitargeted therapies for malignancy. Further, we outline the pharmacology and mechanisms of action of HDAC inhibitors while discussing the safety and efficacy of these compounds in clinical studies to date.

Video abstract

Point your SmartPhone at the code above. If you have a QR code reader the video abstract will appear. Or use:

http://dvpr.es/V0uUng

Introduction

Within eukaryotic cells, chromatin architecture consists of tightly packed DNA, histones, and nonhistone proteins.Citation1 The basic organizing unit of chromatin is the nucleosome and comprises a histone octamer core containing two units each of H2A, H2B, H3, and H4, with 147 base pairs of DNA wrapped tightly around the protein core 1.65 times.Citation1,Citation2 This allows for highly dynamic chromatin architecture. Chromatin undergoes a continual process of condensation and decondensation, which regulates the access of the cellular machinery to specific DNA sequences to facilitate metabolic processes, including transcription, replication, and repair.Citation3,Citation4 The amino acid N-terminal tails of each of the core histones are substrates for a variety of enzyme-catalyzed, reversible, posttranslational modifications, including acetylation, phosphorylation, methylation, and ubiquitination. The combined effect of these modifications creates an epigenetic marking system, or a “histone code,” which governs gene expression.Citation5

Of the various posttranslational modifications, histone acetylation is relatively well characterized. The acetylation status of histones is controlled by the opposing actions of two classes of enzymes: histone acetyltransferases (HATs), which transfer acetyl groups to lysine residues within the N-terminal tails of core histones, and histone deacetylases (HDACs), which remove the acetyl groups.Citation6,Citation7 The acetylation status of histones influences chromatin conformation and affects the accessibility of transcription factors and effector proteins to the DNA, thereby modifying gene expression.Citation7

There are two well-characterized mechanisms by which histone acetylation increases transcriptional activity.Citation4 Firstly, the transfer of the acetyl moiety of acetyl coenzyme A by HATs results in the acetylation of the ɛ-amino tails of lysine residue in histones.Citation8 This neutralizes the positive charge of the histone tails and reduces the affinity of histones for the negatively charged DNA backbone, thereby loosening the structure of the chromatin.Citation6,Citation8 This enables the transcriptional machinery to access the DNA and enhances gene transcription.Citation4 Conversely, HDACs remove the acetyl group from the histone tails, reversing the effects of HATs and altering transcription.Citation9 Secondly, histone acetylation mediates the recruitment of nonhistone proteins to the DNA. Modification of histone tails creates sites that are recognized by effector proteins, which have bromodomains that specifically interact with the modified residues.Citation10,Citation11 Subsequently, the recruited effector proteins modulate DNA transcription.Citation11 It has been widely accepted that enhanced histone acetylation is associated with transcriptionally active DNA, whereas hypoacetylation of histones is associated with transcriptional repression.Citation4,Citation7 However, recent findings have indicated that histone hyperacetylation may not necessarily translate to increased gene expression alone, but also has effects on gene repression or inactivation.Citation12,Citation13

There are currently 18 mammalian HDAC enzymes that have been identified (). These enzymes are classified into four main classes, based on their homology to yeast.Citation14,Citation15 The “classical,” metal-dependent HDAC enzymes involve class I, II, and IV HDACs and the sirtuins; the nonmetal-dependent enzymes represent class III.Citation16 The sirtuins (1–7) share homology to the yeast silent information regulator 2 and differ from the classical HDAC enzymes, as they require the consumption of nicotinamide adenine dinucleotide to deacetylate the lysine residues. The sirtuins have been associated with cell proliferation and cell-cycle control.Citation17 The classical HDAC enzymes are metal-dependent as they contain zinc catalytic binding domains.Citation15 Class I enzymes contain HDAC1, −2, −3, and −8 and are expressed ubiquitously and share homology with the yeast transcriptional regulator RDP3.Citation16 These isotypes are usually expressed within the nucleus and act as transcriptional corepressors. The class II enzymes share homology with the yeast HDAC1 and are subdivided into class IIa, consisting of HDAC4, −5, −7, and −9, and class IIb, containing HDAC6 and −10.Citation18 These isotypes show tissue-specific distribution and are known to shuttle between the nucleus and cytoplasm, although histone proteins broadly represent their main target. The class IIb enzymes differ in that they primarily localize to the cytoplasm and differ structurally by containing two catalytic sites.Citation19 HDAC11 shares homology with the class I isotypes, but shows more tissue-specific distribution with cytoplasmic localization. As it shares relationships with both class I and class II HDACs and structural homology to yeast, it has been designated a distinct class IV.Citation15

Table 1 Classification of the eleven metal-dependent histone deacetylase (HDAC) enzymes

Although HDACs cause the deacetylation of histones, phylogenetic studies indicate that histones are not the primary substrates for HDACs.Citation15 In fact, HATs and HDACs can also regulate gene expression indirectly by mediating the posttranslational acetylation and deacetylation of various nonhistone protein substrates.Citation20 HDACs have more than 50 nonhistone protein substrates, such as DNA-binding proteins, transcription factors, signal-transduction molecules, DNA-repair proteins, and chaperone proteins.Citation20,Citation21 (). The posttranslational modification of these nonhistone proteins can affect many vital regulatory processes, including gene expression, mRNA stability, protein activity, and protein stability.Citation22 For example, HDAC-mediated deacetylation of DNA-binding transcription factors affects their DNA-binding activity, which in turn alters expression of the gene.

Table 2 Partial list of nonhistone protein substrates of HDACs

Abnormalities in the activity or expression of HDACs and HATs can lead to an imbalance between the acetylation and deacetylation of their substrates. Given the importance of histone acetylation and deacetylation in altering chromatin architecture and regulating gene transcription, it follows that abnormalities in histone acetylation status can play a significant role in human disease.Citation23,Citation24 Furthermore, HDACs and HATs have many nonhistone protein substrates, and consequently, the biological implications of HDAC and HAT dysregulation can extend beyond altered gene expression.Citation20

Irregularities in histone acetylation status have been implicated in the development and progression of many diseases, particularly cancer. In particular, loss of acetylation at Lys16 and trimethylation at Lys20 of histone H4 has been found to be a common hallmark of human cancer.Citation25 Furthermore, several lines of evidence have demonstrated the involvement of various HDACs in many malignancies. For example, overexpression of specific HDACs has been identified in a range of human cancers, including HDAC1 in gastricCitation26 and prostateCitation27 cancer, HDAC1 and −6 in breast cancer,Citation28,Citation29 and HDAC2 and −3 in colorectal cancer.Citation30,Citation31 Furthermore, murine knockout models have given rise to possible side effects from the absence of HDACs. Experiments using HDAC1, HDAC2, HDAC3, HDAC4, or HDAC7 knockout mice showed either embryonic lethality or death soon after birth.Citation32Citation36 HDAC3, HDAC5, and HDAC9 knockouts have shown severe cardiac defects involving hypertrophy and fibrosis, and HDAC8 has displayed craniofacial defects.Citation34,Citation37,Citation38 In other cancers, HDAC enzymes are aberrantly recruited to gene promoters. One well-characterized example of aberrant HDAC recruitment in human cancer is that induced by the oncogenic PML–RARα fusion protein. PML–RARα causes acute promyelocytic leukemia by recruiting HDAC-containing complexes to specific target genes, which constitutively repress gene expression.Citation39 This has led to the development of HDAC inhibitors (HDACIs) as anticancer therapies.

Pharmacology, classification, and functions of histone deacetylase inhibitors

HDACIs are a family of naturally derived and synthetically produced compounds that target the classical HDAC enzymes. They are a diverse group of compounds, which vary in structure, biological activity, and specificity. Although histone hyperacetylation is generally associated with transcriptional activation, inhibition of HDACs (which in turn favors histone hyperacetylation) does not necessarily result in a global increase in gene transcription. It has been estimated that up to 20% of all known genes are affected by HDACIs, and of these genes, about half are downregulated and half are upregulated.Citation9

At present, two HDACIs – vorinostat (suberoylanilide hydroxamic acid, Zolinza) and depsipeptide (romidepsin, Istodax) – have received approval from the US Food and Drug Administration (FDA) for treatment of refractory cutaneous T-cell lymphoma (CTCL), and more recently, depsipeptide has gained FDA approval for peripheral T-cell lymphoma (PTCL).Citation40,Citation41 Developments have been made to create chemically distinct HDACIs, with several undergoing intensive clinical trials in various malignancies, many of them focusing on hematological entities, such as the leukemias, lymphomas, and myelodysplastic syndrome.Citation42,Citation43

Broadly, HDACIs can be classified into different structural groups (): the hydroxamic acids, cyclic peptides, bibenzimides, and short-chain fatty acids. The hydroxamates include vorinostat, givinostat, abexinostat, panobinostat, belinostat, and the prototypical HDACI trichostatin A. The cyclic peptides include compounds such as depsipeptide and trapoxin. Benzamides include entinostat and mocetinostat, and together with the hydroxamates and cyclic peptides, have relatively potent inhibition activity within the nanomolar range. Generally, the hydroxamates exert nonspecific HDAC-inhibition activity affecting all classes of HDACs.Citation44,Citation45 Other compounds can exert their properties specifically on class I HDACs, eg, the benzamide entinostat (MS-275), or class I and IIa HDACs, as in the case for the short-chain fatty acids valproic acid (VPA) and butyrate.Citation46 Isotype-selective compounds are also increasingly becoming available, eg, tubacin, mocetinostat, and PC-34501 selectively inhibit HDAC6, −1, and −8, respectively.Citation47Citation50 However, there has been much debate over whether isotype and class-specific HDACIs are preferred over broad-spectrum HDACIs.

Table 3 Characteristics of histone deacetylase (HDAC) inhibitors currently undergoing clinical trials

Mechanisms of HDAC inhibitors

The cellular response to HDACIs is complex and is likely to involve transcriptional and nontranscriptional phenomena. By blocking the activity of HDAC enzymes, HDACIs promote the acetylation of histones and nonhistone proteins. HDACI-mediated modification of histones can result in increased or decreased gene expression (). In addition, targeting histones can influence other DNA-based processes, including DNA replication and repair. Alternatively, through their actions on nonhistone proteins, such as transcription factors and heat shock proteins, HDACIs can alter transcription indirectly, or they may modulate a wide range of cellular processes other than gene expression, through nontranscriptional mechanisms. As a result of these processes, HDACIs are able to elicit a multitude of biological effects on cells, such as apoptosis, cell-cycle arrest, necrosis, autophagy, differentiation, and migration.Citation19,Citation51

Figure 1 HDAC inhibitors promote the acetylation of histones and nonhistone proteins by inhibiting the activity of HDAC enzymes.

Notes: HDAC inhibitor-mediated modification of histones and nonhistone proteins (examples shown) can result in increased or decreased gene expression, influencing other DNA-based processes, including DNA replication and repair. As a result of these processes, HDAC inhibitors are able to elicit a multitude of biological effects on cells, such as apoptosis, cell-cycle arrest, and angiogenesis.
Abbreviations: HDACIs, histone deacetylase inhibitors; HIF, hypoxia-inducible factor; VEGF, vascular endothelial growth factor.
Figure 1 HDAC inhibitors promote the acetylation of histones and nonhistone proteins by inhibiting the activity of HDAC enzymes.

HDACIs have been found to upregulate the cell cyclin-dependent kinase inhibitor p21 and subsequently block the cyclin/CDK complexes, leading to cell G1 cycle arrest in malignant cell lines.Citation52,Citation53 Furthermore, HDACIs cause reduced cyclin-dependent kinase activity via the downregulation of cyclins, causing Rb dephosphorylation and indirectly effecting E2F transcription activity.Citation54

Many in vitro studies have shown the combination of HDACIs with DNA-damaging agents and ionizing radiation cause DNA double-strand breaks, measured by the induction of phosphorylated histone H2AX.Citation55 Although HDACIs may not independently induce DNA double-strand breaks, their involvement in DNA damage may be via several mechanisms. One hypothesis suggests that following alteration of chromatin structure by hyperacetylation, exposure to and severity of DNA-damaging agents is increased.Citation56 Secondly, genes involved in both the homologous recombination and nonhomologous double-strand break-repair pathways are downregulated by HDACIs, such as Ku86, BRCA1, and RAD51.Citation57,Citation58

Apoptosis, the process of programmed cell death, is mediated by intrinsic and extrinsic pathways and is important for tissue homeostasis and development. Apoptosis has been characterized by plasma membrane blebbing and DNA degradation and fractionation of the cell into small vesicles, which are engulfed by phagocytes.Citation59 HDACIs have been shown to induce apoptosis in both solid and hematological malignancies using both transcription-dependent and transcription-independent mechanisms.Citation9,Citation60

HDAC inhibition meddles with the balance between pro- and antiapoptotic proteins involved in cell death. Death receptors and ligands that characterize the extrinsic apoptosis pathways are upregulated by HDACIs and TRAIL (tumor necrosis factor-related apoptosis-inducing ligand) sensitivity may be restored in TRAIL-resistant malignant cells.Citation61 The intrinsic apoptotic pathway is characterized by mitochondrial disruption in response to stress. HDACIs downregulate prosurvival proteins such as Bcl-2 and Bcl-1, which maintain mitochondrial integrity,Citation62 and upregulate proapoptotic proteins such as Bim, Bak, and Bax, which function as sensors of cellular stress and initiate the intrinsic pathway.Citation54,Citation63 Furthermore, hyperacetylation in malignant cells has shown to stabilize p53, promoting cell-cycle arrest and expression of proapoptotic genes.Citation64

The ability of HDACIs to produce these effects suggests that they may be utilized as effective anticancer drugs, eg, by causing apoptosis, DNA damage, or growth arrest in malignant cells. Notably, it has been widely reported that the actions of HDACIs demonstrate relative selectivity for transformed cells over normal cells.Citation65 Furthermore, as well as their effects on tumor cells, HDACIs may also have indirect effects on tumor growth by regulating the host immune response and the tumor vasculature.Citation60

Angiogenesis – the formation of new blood vessels from preexisting vasculature – is driven by the release of vascular endothelial growth factor (VEGF) from surrounding endothelial progenitor cells, macrophages, and fibroblasts.Citation66 Hypoxia-inducible factor (HIF)-1α mediates the expression of several genes involved in angiogenesis and other signaling pathways via the increased expression of VEGF, which induces tumor blood-vessel formation.Citation67 HDAC inhibition has been found to regulate HIF-1α activity indirectly in hypoxic conditions by suppressing HIF-1α and VEGF in malignant in vitro and in vivo models, thus blocking angiogenesis.Citation68 Studies performed under hypoxic conditions in malignant cell lines have shown HDAC1 to be upregulated, subsequently leading to the reduced expression of p53 and von Hippel–Lindau tumor-suppressor genes, with downstream effects of increased HIF-1α and VEGF expression and stimulating angiogenesis of endothelial cells. Treatments with the classical HDACI trichostatin A reversed these effects by upregulating p53 and von Hippel–Lindau tumor-suppressor genes and downregulating HIF-1α and VEGF.Citation68

Furthermore, HIF-1α can also be suppressed indirectly and independently by p300 acetylation.Citation69 Hyperacetylation of chaperone protein Hsp90 via inhibition of HDAC6 by HDACIs leads to increased affinity to HIF-1α. As a result, HIF-1α disrupts Hsp90 chaperone function and exposes HIF-1α to proteasomal degradation by Hsp70.Citation70

HDAC inhibitors in clinical trials

Currently, there are over 80 clinical trials investigating more than eleven different HDACIs for both solid and hematological malignancies as either monotherapies or in combination with various other antitumor agents.

Vorinostat

The hydroxamic acid vorinostat was FDA-approved in 2006 for CTCL, which previously could not be treated with multiple or systemic drugs.Citation40 FDA approval was based on two phase II clinical trials with a 30% response rate in patients with CTCL. Although response rates were similar to previously used therapies, vorinostat showed relatively higher relief from pruritus in comparison to other agents used in the advanced form of the disease. Vorinostat was generally well tolerated, with adverse side effects including diarrhea, fatigue, and nausea. Some patients experienced pulmonary embolism and thrombocytopenia, and there is evidence of long-term safety.Citation71Citation73 Similar response rates have been observed in patients with relapsed non-Hodgkin’s lymphoma and mantle-cell lymphoma; however, the response rates in solid cancers has been ineffective or modest at best.Citation74 Studies in either relapsed or refractory breast, colorectal, or non-small-cell lung cancer had no response.Citation75Citation77 Vorinostat used as a single agent in patients with squamous cell carcinoma of head and neck or ovarian cancer was well tolerated and safe but ineffective.Citation78,Citation79 Studies with breast cancer patients showed no response, with side effects following treatment.Citation75 Although clinical results with Vorinostat used as a single agent have been unsuccessful in treatment of solid malignancies, preclinical data strongly suggest combination with conventional cancer therapies would be beneficial. outlines a list of combinatorial therapies with vorinostat currently under clinical investigations.

Table 4 Partial list of current clinical trials involving histone deacetylase inhibitors as single and combination therapies

Depsipeptide

Depsipeptide represents a bicyclic peptide that has demonstrated potent cytotoxic activity against malignant cells in both in vitro investigations and in vivo tumor xenograft models. A plethora of clinical trials have been undertaken with depsipeptide, representing phase I/II and III trials in patients with colorectal, renal, and breast neoplasms and sarcomas, as well as a wide range of hematological malignancies. Nonhematological toxicities have been mild to moderate, with no record of life-threatening or cardiac toxicities. In summary, depsipeptide can be administered with acceptable short-term toxicity; however, monotherapy appears to have limited clinical activity in acute myeloid leukemia and myelodysplastic syndrome patients.Citation80Citation85

Entinostat

Entinostat (formerly known as MS-2750) has been shown to exhibit many antitumor activities in a range of preclinical investigations. Phase I clinical studies were performed in patients with relapsed or refractory acute myeloid leukemia or refractory solid tumors. Results demonstrated safety and were well tolerated up to 8 mg/mCitation2. No grade 4 toxicities were observed, and dose-limiting toxicities were reversible with no long-term adverse outcomes. Common low-grade toxicities included nausea/vomiting, constipation, fatigue, and cytopenias. HDAC inhibition was observed in PBMCs, and pharmacokinetic analysis suggested a 39 to 80-hour half-life.Citation86Citation88

Valproic acid

VPA is a short-chain fatty acid that has been used in the clinic for the treatment of epilepsy for more than 30 years. Given its HDAC-inhibition activities, VPA has been extensively tested as a monotherapy, but also in combination with other anticancer modalities. In phase I clinical trials, patients with acute myeloid leukemia or myelodysplastic syndrome were treated with VPA, with improvement in 24% of patients. Patients who were either not responsive or who relapsed were also administered all-trans retinoic acid, and the response duration was halved with no additional side effects. Overall, the combination of epigenetic therapy appeared to be more successful in leukemias and was associated with a reverse of aberrant epigenetic marks.Citation89 In separate studies, patients who had acute myeloid leukemia or high-risk myelodysplastic syndrome were administered the combination therapy of the DNA hypomethylating agent azacitidine, all-trans retinoic acid, and VPA. The study reported significant clinical activity and a safe combination.Citation90 Phase I clinical studies have also been performed on solid malignancies, with reports of well-tolerated toxicities.Citation91Citation93 In a clinical trial to assess whether VPA can modulate the effectiveness of temozolomide radiochemotherapy in patients with glioblastoma, it was suggested the combined therapy with VPA was more effective over patients treated with an enzyme-inducing antiepileptic drug. Furthermore, patients treated with VPA had greater success over patients who were not administered any antiepileptics. This study suggests that the observed outcome of combining VPA with temozolomide-based chemoradiotherapy is due to the inhibition of HDAC by VPA. However further investigations are required to determine whether VPA increases temozolomide bioavailability or sensitizes for radiochemotherapy due to its HDAC-inhibition properties.Citation94

Novel HDAC inhibitors

Other than those mentioned earlier, some of the more recent HDACIs that have been tested include abexinostat, givinostat, and mocetinostat. Abexinostat (PCI-24781; formerly CRA-024781) is a broad-spectrum phenyl hydroxamate. Preclinical studies involving combination with radiotherapy have suggested it may act in DNA-repair mechanisms, leading to apoptosis.Citation57,Citation95 In a phase I clinical study involving refractory advanced solid tumors, patients were relatively successful, with adverse side effects including anemia, thrombocytopenia, diarrhea, nausea, vomiting, and fatigue.Citation96 Givinostat (ITF2357) is a synthetic HDACI containing a hydroxamic acid moiety linked to an aromatic ring. Both in vitro and in vivo studies involving human tumor cell lines have shown ITF2357 – used either alone or in combination with other agents – has cytotoxic effects and inhibitory effects on proinflammatory cytokines.Citation97,Citation98 In a phase II open-label nonrandomized clinical study involving heavily pretreated, relapsed, or refractory Hodgkin’s lymphoma patients, preliminary data showed that the oral application of ITF2375 had antitumor activity with an acceptable safety profile. The toxicity profile included grade 1 leukopenia in 30%, grade 2 thrombocytopenia in 33%, fatigue in 50%, grade 1 diarrhea in 40%, and cardiac QT persistence leading to drug discontinuation in 20% of treated patients.Citation99 Mocetinostat (MGCD0103) is a novel HDACI that has strong isotype selectivity to HDAC1 and some weak inhibition to HDAC2, −3, and −11. Studies have found MGCD0103 regulates aberrant gene expression and controls tumorigenic growth in malignancies.Citation100 Phase I and II clinical trials included treatment of advanced solid tumors, relapsed or refractory acute or chronic myeloid leukemia, myelodysplastic syndrome, acute lymphocytic leukemia, diffuse large B-cell lymphoma, follicular lymphoma, and Hodgkin’s lymphoma. MGCD0103 was well tolerated and had antileukemia activity, with side effects consisting mainly of fatigue, nausea, vomiting, and dehydration.Citation101Citation104 A phase I/II trial with MGCD0103 alone or in combination with gemcitabine was performed in patients with solid tumors recently. Preclinical studies found the combination therapy to be more effective than using MGCD0103 alone.Citation105

In summary, extensive cell-based assays and clinical studies with HDACIs have been shown to reduce proliferation, induce cell death and apoptosis, cause cell-cycle arrest, and prevent differentiation and migration selectively in malignant and transformed cells with little effect in normal cells.Citation19,Citation51,Citation106 This provides them with an advantageous stand-alone therapeutic window in oncology. In addition to their intrinsic cytotoxic properties when tested as a single treatment, HDACIs have been shown to induce additive cytotoxic effects when used in combination with conventional anticancer therapies, such as chemotherapy (anthracyclines and retinoic acid) and radiotherapy.Citation9,Citation19,Citation51,Citation107Citation112 Furthermore, studies with HDACIs in combination with ultraviolet radiation and potent iodinated DNA minor groove-binding ligands have been shown to augment photosensitization and cytotoxicity in tumor cells.Citation113,Citation114

Efficacy and safety issues with the use of HDAC inhibitors

Currently, within clinical trials, the overall response rate of patients to HDACIs has been promising, with generally approximately 30% patient success. However, the outcomes from long-term case studies have yet to be reported. There has also been little indication whether class-specific HDACIs such as MS-275 or panspecific HDACIs such as vorinostat have been more successful. The toxicity profiles of HDACIs can be compared between the different types, with side effects mainly consisting of diarrhea, myelosuppression and cardiac QT persistence. Most HDACIs have a half-life of 2–8 hours in plasma and will undergo hepatic metabolization and subsequent intestinal excretions.Citation80,Citation102,Citation115Citation118

Furthermore, the use of HDACIs in nononcological models, such as heart disease including cardiac hypertrophy and myocardial ischemia/reperfusion, has been investigated with the therapeutic potential remaining controversial.Citation119Citation121 Investigations from our laboratory aiming to explore the combinatorial effects of the broad-spectrum HDACI trichostatin A with chemotherapy using the anthracycline doxorubicin to induce hypertrophy in rat cardiac myocytes also suggested detrimental effects caused by the HDACI.Citation122,Citation123 We reported that trichostatin A augmented doxorubicin-induced hypertrophy by altering the expression of hypertrophy-associated genes.Citation122 In addition, further investigations indicated that pretreatment but not posttreatment of cardiac myocytes exposed to trichostatin A and the short-chain fatty acids, VPA, and sodium butyrate augmented DNA damage induced by doxorubicin.Citation122Citation124 It has been proposed that the uncertainty around the therapeutic potential of HDACIs in heart disease stems from the disparate actions of class I and II HDACs.Citation125Citation128 Given the differential findings and the disparity of the actions of the HDACs, particularly in the heart, it is suggested that using isotype- or class-specific HDACIs over broad-spectrum inhibitors may be more successful in this regard.

Conclusion

HDACIs are a promising new group of anticancer agents that have shown positive responses in preclinical and clinical trials. HDACIs have been shown to induce malignant cell death over a large range of solid and hematological malignancies, with generally normal cell resistance. The mechanism of action of HDACIs requires further precise investigations, and normal cell resistance is not understood. However, this will provide an advantageous therapeutic potential over current conventional oncological modalities, which display adverse side effects in normal cells. Furthermore, HDACIs display their biological effects across multiple pathways within the malignant cell, including extrinsic and intrinsic apoptosis, autophagy, inhibiting proliferation, migration, and tumor angiogenesis and effects in the immune response. The fact the normal cells are relatively resistant combined with the multiple defects induced in cancer cells has allowed for relatively high tolerance within clinical investigations. Although clinical trials have been promising, a proportion of patients appear to be unresponsive to HDACI therapy. Accumulating evidence suggests that HDACI therapy may be more successful in combination with other targeted anticancer agents. To date, a range of structurally different inhibitors has been developed that broadly inhibits several HDACs. By developing HDACIs and increasing our understanding of their target HDAC enzymes as well as the effects of targeted inhibitors, we will foster anticancer modalities that are safer and more effective over the current nontargeted agents and current conventional oncological modalities. This will also provide justification for the use of HDACIs as potential therapies for nononcological applications, where we can gain fewer off-target effects by targeting effective biological pathways and processes to reverse or inhibit disease states.

Acknowledgments

The support of the Australian Institute of Nuclear Science and Engineering is acknowledged. TCK was the recipient of AINSE awards. The Epigenomic Medicine Laboratory is supported by the Australian Research Council Future Fellowship and McCord Research. The Allergy and Immune Disorders Laboratory is supported by MCRI, and PVL is the recipient of an Australian National Health and Medical Research Council training fellowship. Both laboratories are supported in part by the Victorian government’s Operational Infrastructure Support Program.

Disclosure

The authors report no conflicts of interest in this work.

References

  • KornbergRDChromatin structure: a repeating unit of histones and DNAScience197418441398688714825889
  • KornbergRDStructure of chromatinAnnu Rev Biochem197746931954332067
  • MorrisonAJShenXChromatin remodelling beyond transcription: the INO80 and SWR1 complexesNat Rev Mol Cell Biol200910637338419424290
  • KouzaridesTChromatin modifications and their functionCell2007128469370517320507
  • JenuweinTAllisCDTranslating the histone codeScience200129355321074108011498575
  • SmithBCDenuJMChemical mechanisms of histone lysine and arginine modificationsBiochim Biophys Acta200917891455718603028
  • BernsteinBEMeissnerALanderESThe mammalian epigenomeCell2007128466968117320505
  • RothSYDenuJMAllisCDHistone acetyltransferasesAnnu Rev Biochem2001708112011395403
  • MinucciSPelicciPGHistone deacetylase inhibitors and the promise of epigenetic (and more) treatments for cancerNat Rev Cancer200661385116397526
  • SanchezRZhouMMThe role of human bromodomains in chromatin biology and gene transcriptionCurr Opin Drug Discov Devel2009125659665
  • IzzoASchneiderRChatting histone modifications in mammalsBrief Funct Genomics201095–642944321266346
  • DuanHHeckmanCABoxerLMHistone deacetylase inhibitors down-regulate bcl-2 expression and induce apoptosis in t(14;18) lymphomasMol Cell Biol20052551608161915713621
  • Rada-IglesiasAEnrothSAmeurAButyrate mediates decrease of histone acetylation centered on transcription start sites and down-regulation of associated genesGenome Res200717670871917567991
  • de RuijterAJvan GennipAHCaronHNKempSvan KuilenburgABHistone deacetylases (HDACs): characterization of the classical HDAC familyBiochem J2003370Pt 373774912429021
  • GregorettiIVLeeYMGoodsonHVMolecular evolution of the histone deacetylase family: functional implications of phylogenetic analysisJ Mol Biol20043381173115050820
  • YangXJSetoEThe Rpd3/Hda1 family of lysine deacetylases: from bacteria and yeast to mice and menNat Rev Mol Cell Biol20089320621818292778
  • TannerKGLandryJSternglanzRDenuJMSilent information regulator 2 family of NAD-dependent histone/protein deacetylases generates a unique product, 1-O-acetyl-ADP-riboseProc Natl Acad Sci USA20009726141781418211106374
  • MartinMKettmannRDequiedtFClass IIa histone deacetylases: regulating the regulatorsOncogene200726375450546717694086
  • MarksPAHistone deacetylase inhibitors: a chemical genetics approach to understanding cellular functionsBiochim Biophys Acta2010179910–1271772520594930
  • OckerMDeacetylase inhibitors – focus on non-histone targets and effectsWorld J Biol Chem201015556121540990
  • GlozakMASenguptaNZhangXSetoEAcetylation and deacetylation of non-histone proteinsGene2005363152316289629
  • SpangeSWagnerTHeinzelTKrämerOHAcetylation of non-histone proteins modulates cellular signalling at multiple levelsInt J Biochem Cell Biol200941118519818804549
  • CressWDSetoEHistone deacetylases, transcriptional control, and cancerJ Cell Physiol2000184111610825229
  • OckerMSchneider-StockRHistone deacetylase inhibitors: signalling towards p21cip1/waf1Int J Biochem Cell Biol2007397–81367137417412634
  • FragaMFBallestarEVillar-GareaALoss of acetylation at Lys16 and trimethylation at Lys20 of histone H4 is a common hallmark of human cancerNat Genet200537439140015765097
  • ChoiJHKwonHJYoonBIExpression profile of histone deacetylase 1 in gastric cancer tissuesJpn J Cancer Res200192121300130411749695
  • HalkidouKGaughanLCookSLeungHYNealDERobsonCNUpregulation and nuclear recruitment of HDAC1 in hormone refractory prostate cancerProstate200459217718915042618
  • ZhangZYamashitaHToyamaTHDAC6 expression is correlated with better survival in breast cancerClin Cancer Res200410206962696815501975
  • ZhangZYamashitaHToyamaTQuantitation of HDAC1 mRNA expression in invasive carcinoma of the breastBreast Cancer Res Treat2005941111616172792
  • ZhuPMartinEMengwasserJSchlagPJanssenKPGöttlicherMInduction of HDAC2 expression upon loss of APC in colorectal tumorigenesisCancer Cell20045545546315144953
  • WilsonAJByunDSPopovaNHistone deacetylase 3 (HDAC3) and other class I HDACs regulate colon cell maturation and p21 expression and are deregulated in human colon cancerJ Biol Chem200628119135481355816533812
  • LaggerGO’CarrollDRemboldMEssential function of histone deacetylase 1 in proliferation control and CDK inhibitor repressionEMBO J200221112672268112032080
  • MontgomeryRLDavisCAPotthoffMJHistone deacetylases 1 and 2 redundantly regulate cardiac morphogenesis, growth, and contractilityGenes Dev200721141790180217639084
  • MontgomeryRLPotthoffMJHaberlandMMaintenance of cardiac energy metabolism by histone deacetylase 3 in miceJ Clin Invest2008118113588359718830415
  • VegaRBMatsudaKOhJHistone deacetylase 4 controls chondrocyte hypertrophy during skeletogenesisCell2004119455556615537544
  • ChangSYoungBDLiSQiXRichardsonJAOlsonENHistone deacetylase 7 maintains vascular integrity by repressing matrix metalloproteinase 10Cell2006126232133416873063
  • ChangSMcKinseyTAZhangCLRichardsonJAHillJAOlsonENHistone deacetylases 5 and 9 govern responsiveness of the heart to a subset of stress signals and play redundant roles in heart developmentMol Cell Biol200424198467847615367668
  • HaberlandMMokalledMHMontgomeryRLOlsonENEpigenetic control of skull morphogenesis by histone deacetylase 8Genes Dev200923141625163019605684
  • LinRJSternsdorfTTiniMEvansRMTranscriptional regulation in acute promyelocytic leukemiaOncogene200120497204721511704848
  • MarksPABreslowRDimethyl sulfoxide to vorinostat: development of this histone deacetylase inhibitor as an anticancer drugNat Biotechnol2007251849017211407
  • Campas-MoyaCRomidepsin for the treatment of cutaneous T-cell lymphomaDrugs Today (Barc)2009451178779520126671
  • PrinceHMBishtonMJHarrisonSJClinical studies of histone deacetylase inhibitorsClin Cancer Res200915123958396919509172
  • LaneAAChabnerBAHistone deacetylase inhibitors in cancer therapyJ Clin Oncol200927325459546819826124
  • MarksPAThe clinical development of histone deacetylase inhibitors as targeted anticancer drugsExpert Opin Investig Drugs201019910491066
  • Schneider-StockROckerMEpigenetic therapy in cancer: molecular background and clinical development of histone deacetylase and DNA methyltransferase inhibitorsIDrugs200710855756117665331
  • HuEDulESungCMIdentification of novel isoform-selective inhibitors within class I histone deacetylasesJ Pharmacol Exp Ther20033072720812975486
  • ParmigianiRBXuWSVenta-PerezGHDAC6 is a specific deacetylase of peroxiredoxins and is involved in redox regulationProc Natl Acad Sci USA2008105289633963818606987
  • HaggartySJKoellerKMWongJCGrozingerCMSchreiberSLDomain-selective small-molecule inhibitor of histone deacetylase 6 (HDAC6)-mediated tubulin deacetylationProc Natl Acad Sci USA200310084389439412677000
  • NamdarMPerezGNgoLMarksPASelective inhibition of histone deacetylase 6 (HDAC6) induces DNA damage and sensitizes transformed cells to anticancer agentsProc Natl Acad Sci USA201010746200032000821037108
  • TangWLuoTGreenbergEFBradnerJESchreiberSLDiscovery of histone deacetylase 8 selective inhibitorsBioorg Med Chem Lett20112192601260521334896
  • MarksPAXuWSHistone deacetylase inhibitors: potential in cancer therapyJ Cell Biochem2009107460060819459166
  • RichonVMSandhoffTWRifkindRAMarksPAHistone deacetylase inhibitor selectively induces p21WAF1 expression and gene-associated histone acetylationProc Natl Acad Sci USA20009718100141001910954755
  • SandorVSenderowiczAMertinsSP21-dependent g(1)arrest with downregulation of cyclin D1 and upregulation of cyclin E by the histone deacetylase inhibitor FR901228Br J Cancer200083681782510952788
  • ZhaoYTanJZhuangLJiangXLiuETYuQInhibitors of histone deacetylases target the Rb-E2F1 pathway for apoptosis induction through activation of proapoptotic protein BimProc Natl Acad Sci USA200510244160901609516243973
  • HarikrishnanKNKaragiannisTCChowMZEl-OstaAEffect of valproic acid on radiation-induced DNA damage in euchromatic and heterochromatic compartmentsCell Cycle20087446847618239454
  • KaragiannisTCEl-OstaAChromatin modifications and DNA double-strand breaks: the current state of playLeukemia200721219520017151702
  • AdimoolamSSirisawadMChenJThiemannPFordJMBuggyJJHDAC inhibitor PCI-24781 decreases RAD51 expression and inhibits homologous recombinationProc Natl Acad Sci USA200710449194821948718042714
  • ZhangYCarrTDimtchevAZaerNDritschiloAJungMAttenuated DNA damage repair by trichostatin A through BRCA1 suppressionRadiat Res2007168111512417722998
  • HotchkissRSStrasserAMcDunnJESwansonPECell deathN Engl J Med2009361161570158319828534
  • BoldenJEPeartMJJohnstoneRWAnticancer activities of histone deacetylase inhibitorsNat Rev Drug Discov20065976978416955068
  • SrivastavaRKKurzrockRShankarSMS-275 sensitizes TRAIL-resistant breast cancer cells, inhibits angiogenesis and metastasis, and reverses epithelial-mesenchymal transition in vivoMol Cancer Ther20109123254326621041383
  • RikiishiHAutophagic and apoptotic effects of HDAC inhibitors on cancer cellsJ Biomed Biotechnol2011201183026021629704
  • ZhangYAdachiMKawamuraRImaiKBmf is a possible mediator in histone deacetylase inhibitors FK228 and CBHA-induced apoptosisCell Death Differ200613112914015947789
  • XuYRegulation of p53 responses by post-translational modificationsCell Death Differ200310440040312719715
  • MaXEzzeldinHHDiasioRBHistone deacetylase inhibitors: current status and overview of recent clinical trialsDrugs200969141911193419747008
  • AdamsRHAlitaloKMolecular regulation of angiogenesis and lymphangiogenesisNat Rev Mol Cell Biol20078646447817522591
  • LinEYPollardJWTumor-associated macrophages press the angiogenic switch in breast cancerCancer Res200767115064506617545580
  • KimMSKwonHJLeeYMHistone deacetylases induce angiogenesis by negative regulation of tumor suppressor genesNat Med20017443744311283670
  • FathDMKongXLiangDHistone deacetylase inhibitors repress the transactivation potential of hypoxia-inducible factors independently of direct acetylation of HIF-alphaJ Biol Chem200628119136121361916543236
  • KongXLinZLiangDFathDSangNCaroJHistone deacetylase inhibitors induce VHL and ubiquitin-independent proteasomal degradation of hypoxia-inducible factor 1alphaMol Cell Biol20062662019202816507982
  • DuvicMTalpurRNiXPhase 2 trial of oral vorinostat (suberoylanilide hydroxamic acid SAHA) for refractory cutaneous T-cell lymphoma (CTCL)Blood20071091313916960145
  • OlsenEAKimYHKuzelTMPhase IIb multicenter trial of vorinostat in patients with persistent progressive or treatment refractory cutaneous T-cell lymphomaJ Clin Oncol200725213109311517577020
  • KhanOLa ThangueNBDrug insight: histone deacetylase inhibitor-based therapies for cutaneous T-cell lymphomasNat Clin Pract Oncol200851271472618839006
  • KirschbaumMFrankelPPopplewellLPhase II study of vorinostat for treatment of relapsed or refractory indolent non-Hodgkin’s lymphoma and mantle cell lymphomaJ Clin Oncol20112991198120321300924
  • LuuTHMorganRJLeongLA phase II trial of vorinostat (suberoylanilide hydroxamic acid) in metastatic breast cancer: a California Cancer Consortium studyClin Cancer Res200814217138714218981013
  • TraynorAMDubeySEickhoffJCVorinostat (NSC# 701852) in patients with relapsed non-small cell lung cancer: a Wisconsin Oncology Network phase II studyJ Thorac Oncol20094452252619347984
  • VansteenkisteJVan CutsemEDumezHEarly phase II trial of oral vorinostat in relapsed or refractory breast colorectal or non-small cell lung cancerInvest New Drugs200826548348818425418
  • BlumenscheinGRJrKiesMSPapadimitrakopoulouVAPhase II trial of the histone deacetylase inhibitor vorinostat (Zolinza, suberoylanilide hydroxamic acid SAHA) in patients with recurrent and/or metastatic head and neck cancerInvest New Drugs2008261818717960324
  • ModesittSCSillMHoffmanJSBenderDPA phase II study of vorinostat in the treatment of persistent or recurrent epithelial ovarian or primary peritoneal carcinoma: a Gynecologic Oncology Group studyGynecol Oncol2008109218218618295319
  • SandorVBakkeSRobeyRWPhase I trial of the histone deacetylase inhibitor depsipeptide (FR901228 NSC 630176) in patients with refractory neoplasmsClin Cancer Res20028371872811895901
  • MarshallJLRizviNKauhJA phase I trial of depsipeptide (FR901228) in patients with advanced cancerJ Exp Ther Oncol20022632533212440223
  • StadlerWMMargolinKFerberSMcCullochWThompsonJAA phase II study of depsipeptide in refractory metastatic renal cell cancerClin Genitourin Cancer200651576016859580
  • ByrdJCMarcucciGParthunMRA phase 1 and pharmacodynamic study of depsipeptide (FK228) in chronic lymphocytic leukemia and acute myeloid leukemiaBlood2005105395996715466934
  • KlimekVMFircanisSMaslakPTolerability pharmacodynamics and pharmacokinetics studies of depsipeptide (romidepsin) in patients with acute myelogenous leukemia or advanced myelodysplastic syndromesClin Cancer Res200814382683218245545
  • SchrumpDSFischetteMRNguyenDMClinical and molecular responses in lung cancer patients receiving romidepsinClin Cancer Res200814118819818172270
  • KummarSGutierrezMGardnerERPhase I trial of MS-275, a histone deacetylase inhibitor, administered weekly in refractory solid tumors and lymphoid malignanciesClin Cancer Res20071318 Pt 15411541717875771
  • GoreLRothenbergMLO’BryantCLA phase I and pharmacokinetic study of the oral histone deacetylase inhibitor, MS-275, in patients with refractory solid tumors and lymphomasClin Cancer Res200814144517452518579665
  • GojoIJiemjitATrepelJBPhase 1 and pharmacologic study of MS-275, a histone deacetylase inhibitor, in adults with refractory and relapsed acute leukemiasBlood200710972781279017179232
  • BlumWKlisovicRBHackansonBPhase I study of decitabine alone or in combination with valproic acid in acute myeloid leukemiaJ Clin Oncol200725253884389117679729
  • SorianoAOYangHFaderlSSafety and clinical activity of the combination of 5-azacytidine, valproic acid, and all-trans retinoic acid in acute myeloid leukemia and myelodysplastic syndromeBlood200711072302230817596541
  • AtmacaAAl-BatranSEMaurerAValproic acid (VPA) in patients with refractory advanced cancer: a dose escalating phase I clinical trialBr J Cancer200797217718217579623
  • MünsterPMarchionDBicakuEPhase I trial of histone deacetylase inhibition by valproic acid followed by the topoisomerase II inhibitor epirubicin in advanced solid tumors: a clinical and translational studyJ Clin Oncol200725151979198517513804
  • CandelariaMGallardo-RincónDArceCA phase II study of epigenetic therapy with hydralazine and magnesium valproate to overcome chemotherapy resistance in refractory solid tumorsAnn Oncol20071891529153817761710
  • WellerMGorliaTCairncrossJGProlonged survival with valproic acid use in the EORTC/NCIC temozolomide trial for glioblastomaNeurology201177121156116421880994
  • BanuelosCABanáthJPMacPhailSHZhaoJReitsemaTOlivePLRadiosensitization by the histone deacetylase inhibitor PCI-24781Clin Cancer Res20071322 Pt 16816682618006784
  • UndeviaSJanischLSchilskyRLPhase I study of the safety, pharmacokinetics (PK) and pharmacodynamics (PD) of the histone deacetylase inhibitor (HDACi) PCI-24781J Clin Oncol200826 Suppl14514
  • BarbettiVGozziniARovidaESelective anti-leukaemic activity of low-dose histone deacetylase inhibitor ITF2357 on AML1/ETO-positive cellsOncogene200827121767177817891169
  • GueriniVBarbuiVSpinelliOThe histone deacetylase inhibitor ITF2357 selectively targets cells bearing mutated JAK2(V617F)Leukemia200822474074718079739
  • VivianiSBonfanteVFasolaCValagussaPGianniAMPhase II study of the histone-deacetylase inhibitor ITF2357 in relasped/refractory Hodgkin’s lymphoma patientsJ Clin Oncol200826 Suppl8532
  • Le TourneauCSiuLLPromising antitumor activity with MGCD0103, a novel isotype-selective histone deacetylase inhibitorExpert Opin Investig Drugs200817812471254
  • SiuLLPiliRDuranIPhase I study of MGCD0103 given as a three-times-per-week oral dose in patients with advanced solid tumorsJ Clin Oncol200826121940194718421048
  • Garcia-ManeroGAssoulineSCortesJPhase 1 study of the oral isotype specific histone deacetylase inhibitor MGCD0103 in leukemiaBlood2008112498198918495956
  • CrumpMAndreadisCAssoulineSTreatment of relapsed or refractory non-hodgkin lymphoma with the oral isotype-selective histone deacetylase inhibitor MGCD0103: Interim results from a phase II studyJ Clin Oncol200826 Suppl8528
  • BociekRGKuruvillaJProBIsotype-selective histone deacetylase (HDAC) inhibitor MGCD0103 demonstrates clinical activity and safety in patients with relapsed/refractory classical Hodgkin lymphoma (HL)J Clin Oncol200826 Suppl8507
  • HurwitzHNelsonBO’DwyerPJPhase I/II: The oral isotype-selective HDAC inhibitor MGCD0103 in combination with gemcitabine (Gem) in patients (pts) with refractory solid tumorsJ Clin Oncol200826 Suppl4625
  • KwaFABalcerczykALicciardiPEl-OstaAKaragiannisTCChromatin modifying agents – the cutting edge of anticancer therapyDrug Discov Today20111613–1454354721664485
  • De los SantosMZambranoASánchez-PachecoAArandaAHistone deacetylase inhibitors regulate retinoic acid receptor beta expression in neuroblastoma cells by both transcriptional and posttranscriptional mechanismsMol Endocrinol200721102416242617622583
  • De los SantosMZambranoAArandaACombined effects of retinoic acid and histone deacetylase inhibitors on human neuroblastoma SH-SY5Y cellsMol Cancer Ther2007641425143217431121
  • KimHJBaeSCHistone deacetylase inhibitors: molecular mechanisms of action and clinical trials as anti-cancer drugsAm J Transl Res20113216617921416059
  • KaragiannisTCHarikrishnanKNEl-OstaADisparity of histone deacetylase inhibition on repair of radiation-induced DNA damage on euchromatin and constitutive heterochromatin compartmentsOncogene200726273963397117213813
  • KaragiannisTCHarikrishnanKNEl-OstaAThe histone deacetylase inhibitor, trichostatin A, enhances radiation sensitivity and accumulation of gammaH2A.XCancer Biol Ther20054778779316082178
  • Sanchez-GonzalezBYangHBueso-RamosCAntileukemia activity of the combination of an anthracycline with a histone deacetylase inhibitorBlood200610841174118216675713
  • BriggsBVerverisKRoddALFoongLJSilvaFMKaragiannisTCPhotosensitization by iodinated DNA minor groove binding ligands: evaluation of DNA double-strand break induction and repairJ Photochem Photobiol B2011103214515221440453
  • SinghTRShankarSSrivastavaRKHDAC inhibitors enhance the apoptosis-inducing potential of TRAIL in breast carcinomaOncogene200524294609462315897906
  • KellyWKO’ConnorOAKrugLMPhase I study of an oral histone deacetylase inhibitor suberoylanilide hydroxamic acid in patients with advanced cancerJ Clin Oncol200523173923393115897550
  • RubinEHAgrawalNGFriedmanEJA study to determine the effects of food and multiple dosing on the pharmacokinetics of vorinostat given orally to patients with advanced cancerClin Cancer Res200612237039704517145826
  • Garcia-ManeroGYangHBueso-RamosCPhase 1 study of the histone deacetylase inhibitor vorinostat (suberoylanilide hydroxamic acid [SAHA]) in patients with advanced leukemias and myelodysplastic syndromesBlood200811131060106617962510
  • Lech-MarandaERobakEKoryckaARobakTDepsipeptide (FK228) as a novel histone deacetylase inhibitor: mechanism of action and anticancer activityMini Rev Med Chem20077101062106917979809
  • CaoDJWangZVBattiproluPKHistone deacetylase (HDAC) inhibitors attenuate cardiac hypertrophy by suppressing autophagyProc Natl Acad Sci USA2011108104123412821367693
  • ChoYKEomGHKeeHJSodium valproate, a histone deacetylase inhibitor, but not captopril, prevents right ventricular hypertrophy in ratsCirc J201074476077020208383
  • BogaardHJMizunoSHussainiAASuppression of histone deacetylases worsens right ventricular dysfunction after pulmonary artery banding in ratsAm J Respir Crit Care Med2011183101402141021297075
  • KaragiannisTCLinAJVerverisKTrichostatin A accentuates doxorubicin-induced hypertrophy in cardiac myocytesAging (Albany NY)201021065966820930262
  • VerverisKRoddALTangMMEl-OstaAKaragiannisTCHistone deacetylase inhibitors augment doxorubicin-induced DNA damage in cardiomyocytesCell Mol Life Sci201168244101411421584806
  • VerverisKKaragiannisTCPotential non-oncological applications of histone deacetylase inhibitorsAm J Transl Res20113545446722046487
  • McKinseyTAOlsonENDual roles of histone deacetylases in the control of cardiac growthNovartis Found Symp2004259132141 discussion 141–145, 163–16915171251
  • BacksJOlsonENControl of cardiac growth by histone acetylation/deacetylationCirc Res2006981152416397154
  • HaberlandMMontgomeryRLOlsonENThe many roles of histone deacetylases in development and physiology: implications for disease and therapyNat Rev Genet2009101324219065135
  • ZhangCLMcKinseyTAChangSAntosCLHillJAOlsonENClass II histone deacetylases act as signal-responsive repressors of cardiac hypertrophyCell2002110447948812202037
  • GuWRoederRGActivation of p53 sequence-specific DNA binding by acetylation of the p53 C-terminal domainCell19979045956069288740
  • ThevenetLMéjeanCMoniotBRegulation of human SRY subcellular distribution by its acetylation/deacetylationEMBO J200423163336334515297880
  • YuanZLGuanYJChatterjeeDChinYEStat3 dimerization regulated by reversible acetylation of a single lysine residueScience2005307570726927315653507
  • BoyesJByfieldPNakataniYOgryzkoVRegulation of activity of the transcription factor GATA-1 by acetylationNature199839667115945989859997
  • HayakawaFTowatariMOzawaYTomitaAPrivalskyMLSaitoHFunctional regulation of GATA-2 by acetylationJ Leukoc Biol200475352954015001660
  • Martínez-BalbásMABauerUMNielsenSJBrehmAKouzaridesTRegulation of E2F1 activity by acetylationEMBO J200019466267110675335
  • SartorelliVPuriPLHamamoriYAcetylation of MyoD directed by PCAF is necessary for the execution of the muscle programMol Cell19994572573410619020
  • YaoYLYangWMSetoERegulation of transcription factor YY1 by acetylation and deacetylationMol Cell Biol200121175979599111486036
  • MunshiNMerikaMYieJSengerKChenGThanosDAcetylation of HMG I(Y) by CBP turns off IFN beta expression by disrupting the enhanceosomeMol Cell1998244574679809067
  • LührsHHockRSchauberJModulation of HMG-N2 binding to chromatin by butyrate-induced acetylation in human colon adenocarcinoma cellsInt J Cancer200297556757311807779
  • KiernanRBrèsVNgRWPost-activation turn-off of NF-kappa B-dependent transcription is regulated by acetylation of p65J Biol Chem200327842758276612419806
  • LuoJLiMTangYLaszkowskaMRoederRGGuWAcetylation of p53 augments its site-specific DNA binding both in vitro and in vivoProc Natl Acad Sci USA200410182259226414982997
  • MunshiNAgaliotiTLomvardasSMerikaMChenGThanosDCoordination of a transcriptional switch by HMGI(Y) acetylationScience200129355321133113611498590
  • WangRCherukuriPLuoJActivation of Stat3 sequence-specific DNA binding and transcription by p300/CREB-binding protein-mediated acetylationJ Biol Chem200528012115281153415649887
  • FuMWangCReutensATP300 and p300/cAMP-response element-binding protein-associated factor acetylate the androgen receptor at sites governing hormone-dependent transactivationJ Biol Chem200027527208532086010779504
  • GaughanLLoganIRCookSNealDERobsonCNTip60 and histone deacetylase 1 regulate androgen receptor activity through changes to the acetylation status of the receptorJ Biol Chem200227729259042951311994312
  • WangCFuMAngelettiRHDirect acetylation of the estrogen receptor alpha hinge region by p300 regulates transactivation and hormone sensitivityJ Biol Chem200127621183751838311279135
  • YamagataTMitaniKOdaHAcetylation of GATA-3 affects T-cell survival and homing to secondary lymphoid organsEMBO J200019174676468710970860
  • ZhangWBiekerJJAcetylation and modulation of erythroid Krüppel-like factor (EKLF) activity by interaction with histone acetyltransferasesProc Natl Acad Sci USA19989517985598609707565
  • MarzioGWagenerCGutierrezMICartwrightPHelinKGiaccaME2F family members are differentially regulated by reversible acetylationJ Biol Chem200027515108871089210753885
  • JinYHJeonEJLiQLTransforming growth factor-beta stimulates p300-dependent RUNX3 acetylation which inhibits ubiquitination-mediated degradationJ Biol Chem200427928294092941715138260
  • JeongJWBaeMKAhnMYRegulation and destabilization of HIF-1alpha by ARD1-mediated acetylationCell2002111570972012464182
  • ItoAKawaguchiYLaiCHMDM2-HDAC1-mediated deacetylation of p53 is required for its degradationEMBO J200221226236624512426395
  • PatelJHDuYArdPGThe c-MYC oncoprotein is a substrate of the acetyltransferases hGCN5/PCAF and TIP60Mol Cell Biol20042424108261083415572685
  • GaughanLLoganIRNealDERobsonCNRegulation of androgen receptor and histone deacetylase 1 by Mdm2-mediated ubiquitylationNucleic Acids Res2005331132615640443
  • KawaiHLiHAvrahamSJiangSAvrahamHKOverexpression of histone deacetylase HDAC1 modulates breast cancer progression by negative regulation of estrogen receptor alphaInt J Cancer2003107335335814506733
  • GrönroosEHellmanUHeldinCHEricssonJControl of Smad7 stability by competition between acetylation and ubiquitinationMol Cell200210348349312408818
  • FuMWangCWangJAndrogen receptor acetylation governs trans activation and MEKK1-induced apoptosis without affecting in vitro sumoylation and trans-repression functionMol Cell Biol200222103373338811971970
  • ZhangWKadamSEmersonBMBiekerJJSite-specific acetylation by p300 or CREB binding protein regulates erythroid Krüppel-like factor transcriptional activity via its interaction with the SWI-SNF complexMol Cell Biol20012172413242211259590
  • BannisterAJMiskaEAGörlichDKouzaridesTAcetylation of importin-alpha nuclear import factors by CBP/p300Curr Biol200010846747010801418
  • ChenLFFischleWVerdinEGreeneWCDuration of nuclear NF-kappaB action regulated by reversible acetylationScience200129355351653165711533489
  • CohenHYLavuSBittermanKJAcetylation of the C terminus of Ku70 by CBP and PCAF controls Bax-mediated apoptosisMol Cell200413562763815023334
  • KovacsJJMurphyPJGaillardSHDAC6 regulates Hsp90 acetylation and chaperone-dependent activation of glucocorticoid receptorMol Cell200518560160715916966