2,357
Views
31
CrossRef citations to date
0
Altmetric
Perspective

Histone deacetylases 1 and 2 regulate DNA replication and DNA repair: potential targets for genome stability-mechanism-based therapeutics for a subset of cancers

Pages 1779-1785 | Received 24 Feb 2015, Accepted 13 Apr 2015, Published online: 17 Jun 2015

Abstract

Histone deacetylases 1 and 2 (HDAC1,2) belong to the class I HDAC family, which are targeted by the FDA-approved small molecule HDAC inhibitors currently used in cancer therapy. HDAC1,2 are recruited to DNA break sites during DNA repair and to chromatin around forks during DNA replication. Cancer cells use DNA repair and DNA replication as survival mechanisms and to evade chemotherapy-induced cytotoxicity. Hence, it is vital to understand how HDAC1,2 function during the genome maintenance processes (DNA replication and DNA repair) in order to gain insights into the mode-of-action of HDAC inhibitors in cancer therapeutics. The first-in-class HDAC1,2-selective inhibitors and Hdac1,2 conditional knockout systems greatly facilitated dissecting the precise mechanisms by which HDAC1,2 control genome stability in normal and cancer cells. In this perspective, I summarize the findings on the mechanistic functions of class I HDACs, specifically, HDAC1,2 in genome maintenance, unanswered questions for future investigations and views on how this knowledge could be harnessed for better-targeted cancer therapeutics for a subset of cancers.

Histone Deacetylases (HDACs) and Histone Deacetylase Inhibitors (HDIs)

Histone deacetylases remove acetyl groups from histone and non-histone proteins;Citation1,2 they are properly termed as lysine deacetylases or KDACs, but for historical reasons remain better known as HDACs. Eighteen HDACs have been identified so far in mammalian cells, which are divided into four classes: Class I HDACs include HDACs 1, 2, 3 and 8, and are homologous to yeast Rpd3. Class II consists of HDACs 4, 5, 6, 7, 9 and 10, and they have high similarity to yeast Hda1. Additionally, Class IIa enzymes (HDACs 4, 5, 7, 9) depend on Class I HDACs for their activity. Sirtuins, similar to yeast Sir2, form Class III HDACs and their activity is dependent on NAD. HDAC11 is the only Class IV HDAC with features similar to Class I and Class II HDACs.Citation2,3

Histone deacetylase inhibitors (HDIs), small molecules that inhibit HDAC activities, are potent anti-proliferative agents that selectively kill cancer cells. Currently available HDIs belong to four classes based on their chemical structure: hydroxamate, cyclic peptide, aliphatic acids and benzamide.Citation4 When used as a monotherapy agent, HDIs have shown a very positive response in patients with hematologic malignancies, but they have been less effective on solid tumors.Citation5 Two broad-spectrum or pan−HDIs, SAHA or Vorinostat (a hydroxamate class inhibitor) and Depsipeptide or Romidepsin (a cyclic peptide class inhibitor), are currently approved by the FDA for the treatment of refractory cutaneous T-cell lymphoma (CTCL).Citation6,7 This approval has spurred the creation and testing of several novel HDIs and at least 20 of them are currently in preclinical or clinical trials.Citation8 An effective HDI to treat B-cell malignancies is still not available in the clinic and several phase I studies using SAHA/vorinostat are currently underway for lymphomas.

While SAHA and Depsipeptide are effective anti-neoplastic agents, the broad-spectrum mechanism of action of these FDA-approved HDIs was not fully understood. HDACs and HDIs have been studied mainly in the context of gene transcription and not genome stability in cancers. For instance, acute myeloid leukemia (AML) is associated with misregulated gene expression due to recruitment of HDACs to ectopic loci by mutant transcription factors.Citation9,10 HDI treatment is therefore thought to kill these AML cells by reversing the repressive effects of aberrantly targeted HDACs.Citation11 To develop and use specific selective inhibitors to individual HDACs, which might lead to a safer and more effective HDAC inhibitors with less side-effects, it is imperative to understand the transcription-independent functions of individual HDACs in normal and cancer cells. This knowledge is required for not only better design, but also for the better understanding of the mode-of-action of selective HDAC inhibitors.

Genetic analyses of Class I HDACs using conditional knockout mouse models and knockdown systems reveal their important functions in the genome maintenance

Several HDAC inhibitors are currently being used in clinical trials for B-cell lymphomas and two of these inhibitors are FDA-approved drugs for the treatment of refractory cutaneous T-cell lymphoma. However, these FDA-approved broad-spectrum HDAC inhibitors inhibit ten different HDACs, and therefore have several adverse side effects in cancer patients, which include cardiac toxicity, thrombocytopenia and gastrointestinal toxicity. Hence, using selective inhibitors against specific class I will cause maximum cytotoxicity in cancer cells, but at the same time minimize the side effects caused by the use of pan-HDAC inhibitors. Before, we use selective HDAC inhibitors in the clinic; it is imperative to understand biological functions of individual HDACs. Class I HDACs consist of HDAC1, 2, 3 and 8. Our previous studies uncovered novel functions for HDAC3 in genome maintenance processes (DNA repair and DNA replication).Citation12-14

Targeted deletion of Hdac3 in the germ line led to embryonic lethality.Citation12 Using conditional Hdac3 knockout mice, we showed that the deletion of Hdac3 leads to S-phase-dependent DNA double-strand breaks in cycling cells and not in quiescent cells, which could provide a therapeutic window.Citation12 Cell cycle analysis of Hdac3-null cells revealed a delay in the progression of cells through the S-phase, accumulation of S-phase dependent DNA damage and activation of the S-phase cell cycle checkpoint response.Citation12 Moreover, HDAC3-null cells displayed a reduction in both homologous recombination and non-homologous recombination pathways,Citation14 suggesting that the absence of HDAC3 impairs double-strand DNA break repair. Even though short-term loss of Hdac3 causes cell death in primary mouse embryo fibroblasts, long-term deletion of Hdac3 in livers leads to hepatocellular carcinoma in mice.Citation14 Thus, genetic knockout studies provide not only insights into the biological functions of HDACs, but also provide valuable information regarding the efficacy of HDIs in cancer therapy when used for a short duration and the demerits of their sustained long-term use.Citation12-14 HDAC3 activity is also required for the removal of H3K4 acetylation (H3K4ac) at centromeres and loss of HDAC3 leads to dissociation of sister chromatids in HeLa cells.Citation15 The increase in H3K4ac in HDAC3-depleted cells occurs before the entry of cells into mitosis,Citation15 which is consistent with our findings that HDAC3 dissociates from chromatin before the metaphase stage.Citation14 Hence, in addition to DNA repair and DNA replication, HDAC3 maintains genome stability by controlling centromeric functions.Citation15 Recent intriguing discoveries demonstrate that HDAC3 could regulate transcription independent of its catalytic activity in mouse livers.Citation16 In this study, the interaction of HDAC3 with its stable partners NCoR and SMRT was found to be important for the deacetylase-independent functions of HDAC3 during transcription.Citation16 Interestingly, we previously found that knockdown of both NCoR and SMRT decreased the total cellular levels of HDAC3 and activated the DNA damage response, demonstrating the importance of NCoR-SMRT-HDAC3 nexus in regulating genome stability.Citation14 Hence, targeting NCOR and SMRT, in addition to targeting Hdac3, would be another strategy to trigger DNA damage response in cancer cells.

A great deal of knowledge gained from the elegant studies using conditional single gene or double gene knockout mouse models in various cell lineages or tissues have together shed light on the vital, redundant and non-redundant biological functions of HDAC1 and HDAC2, which are two highly similar enzymes. Targeted deletion of Hdac1 led to embryonic lethality.Citation17 Distinct phenotypes in Hdac2-null (Hdac2-/-) mice were observed. In one study, Hdac2-/- pups died within a month, due to cardiac defects and abnormalities in myocyte proliferation.Citation18 In another study, 50 percent of Hdac2-/- pups died perinatally, whereas the remaining littermates survived.Citation19 Other studies indicate that Hdac2-/- mice are viable.Citation20,21 However, mice that survived had smaller hearts when compared to the control littermates. The reason behind different phenotypes in Hdac2-/- mice could be attributed to the different strategies used to make these knockout mice and (or) due to different mouse strains. In the vast majority of cell types (including those made during hematopoiesis), targeted deletion of either Hdac1 or Hdac2 has minimal effects on proliferation and the cell cycle, likely due to compensation for one by the other.Citation22,23 However, combined deletion of both Hdac1 and Hdac2 (Hdac1,2) dramatically impairs proliferation in multiple cell types by blocking cells at the G1 to S phase stage.Citation23 Simultaneous deletion of Hdac1 and Hdac2 in early B-cell progenitors leads to a dramatic block in B-cell development and apoptosis. However, mature non-dividing terminally differentiated B-cells can tolerate loss of both Hdac1 and Hdac2, but their deletion affects proliferation upon exogenous mitogenic stimulation.Citation23 Loss of HDAC1,2 function in embryonic stem cells led to chromatin bridges and mitotic instability.Citation24 Similarly, loss of HDAC1,2 function in fibrosarcoma cells led to mitotic catastrophy.Citation25 HDAC1,2 dissociate from mitotic chromosomes during mitosis and reappear in the daughter nuclei.Citation26 Interestingly, loss of HDAC1 function in tumor cells led to defective entry into mitosis, which in turn leads to apoptosis.Citation27 These findings together reveal that HDAC1,2 play a crucial role when cells progress through the cell cycle.

In mice with Hdac2 alleles but without the Hdac1 alleles (i.e., Hdac1-null mice), immature thymocytes accumulation and lymphoblastic lymphomas were observed.Citation28 Similarly, mice with Hdac1 and no Hdac2 alleles (i.e., Hdac2-null mice) also developed lymphomas.Citation28 However, lymphamogenesis was not observed when both HDAC1 and HDAC2 activities are lost in T-cells.Citation28 Instead, deletion of Hdac1 and Hdac2 led to a block in thymocyte development. A similar dosage dependent effect upon loss of either Hdac1 or Hdac2 was observed on epidermal proliferation and differentiation.Citation29 In this elegant study, deletion of a single Hdac2 allele in Hdac1-knockout mice caused severe defects in the development of epidermal lineages and spontaneous tumor formation.Citation29 Therefore, the use of genetic knockout mouse models in parallel with the isotype selective inhibitors of HDACs is required to compare and assess the mode-of-action, specificity, benefits, and pitfalls of complete and partial inhibition of individual HDACs.

The deletion of Hdac8, another class I HDAC, in mice led to death within 4–6 h following birth.Citation30 Conditional deletion of Hdac8 specifically in the brain caused loss of cranial neural crest cells and instability of the skull in mice.Citation30 Hence, HDAC8 function is vital for skull development. HDAC8 is a deacetylase for SMC3 (Sister Maintenance of Chromosome protein 3, a core cohesion component), and mutations in HDAC8 that disrupt SMC3 deacetylation result in an improper renewal of cohesin components and inadequate recycling of the cohesin components in the next cell cycle. This results in decreased cohesin occupancy in the genome leading to clinical features of Cornelia de Lange syndrome (CdLS), which is characterized by the congenital malformation disorder.Citation31

Hence, all four class I HDACs are required for genome maintenance in mammalian cells. However, HDAC1,2 are recruited to sites of DNA replication and DNA damage break sites, suggesting a direct role for these two enzymes during DNA replication and double-strand break (DSB) repair.Citation32-34 HDAC3 associates with nascent chromatin during DNA replication.Citation33 Even though loss of HDAC3 impairs DSB repair,Citation14 it is not localized to DSB DNA damage sites.Citation14,32 Hence, the repair defects observed in Hdac3-null cells is likely due to the defective chromatin structure. Direct functions for HDAC8 at the replication fork and at DNA damage sites remain to be investigated.

HDAC1,2 maintain genome stability during DNA replication - the engine that drives rapid proliferation of cycling cells

Defective DNA repair and/or DNA replication are major causes of genome instability, which can trigger cell death. A single unrepaired double strand break is sufficient to cause cell death. Hence, it is crucial to understand functions for HDACs in genome maintenance pathways. HDAC1,2 interact with PCNACitation35 and localize to sites of replication.Citation33,34 Deletion of both HDAC1 and HDAC2 using conditional knockout system results in the arrest of cells in the G1 phase,Citation23 and hence it was difficult to study functions of HDAC1,2 in the S-phase progression. Selective inhibitors make an excellent tool to study biological functions of HDAC1,2 as we could transiently inhibit their functions for a short duration of time when there is an impact on DNA replication and repair and before cells arrest in G1 phase.

Using selective inhibitors, we examined whether HDAC1,2 activities are required for efficient DNA replication in mammalian cells. A decrease in replication fork velocity was observed upon HDAC1,2 inhibition, demonstrating that HDAC1,2 activities are required for efficient replication fork elongation.Citation34 If stalled forks are not restarted in a timely fashion, it can result in fork collapse, formation of DSBs and activation of the DNA damage response.Citation36 Increased γH2AX (phosphorylated form of H2AX and a marker of double-strand breaks) was observed in cells treated with HDAC1,2 inhibitor or following knockdown of Hdac1,2. RNA-seq analysis of S-phase cells treated with HDAC1,2 selective inhibitor showed no change in the expression of genes involved in DNA replication or DNA repair.Citation34 Hence, our findings showed a direct role for HDAC1,2 in ensuring the proper progression of the replication fork. Activation of DNA damage response in S-phase cells due to obstruction of DNA fork progression provided the mechanistic basis for how HDAC1,2 selective inhibitors might be able to directly target regulators of genome stability in order to kill the rapidly cycling cancer cells.

Upon dissection of the mechanism behind these replication defects, we found that these enzymes control chromatin structure at and around replicating regions by targeting histone acetylation. We developed a technique termed modified BrdU-ChIP assay to examine histone modifications and replication proteins on nascent DNA.Citation34 This technique is a complement to the elegant iPOND technique that permits one to follow the dynamics of replication proteins associated with nascent DNA.Citation33 Using the modified BrdU-ChIP approach, we showed that inhibition of HDAC1,2 activity increases H4K16ac present on newly synthesized DNA and at replicating origins.Citation34 Hence, HDAC1,2 regulate H4K16ac associated with nascent chromatin during DNA replication. H4K16ac is known to block inter-nucleosomal interactions and disrupt chromatin packaging.Citation37 Therefore, we asked whether HDAC1,2 regulate nascent chromatin packaging around replication forks by targeting H4K16ac? An increase in the release of BrdU-labeled nascent DNA associated with di- and tri-nucleosomes was observed following micrococcal nuclease digestion in HDAC1,2 inhibitor-treated cells, confirming the role for HDAC1,2 activities in chromatin compaction during DNA replication.Citation34 SMARCA5 is a ISWI-family ATP-dependent chromatin remodeler, whose activity is inhibited by H4K16ac and H4K12ac marks,Citation38 and both of these marks are targets of HDAC1,2.Citation34 Hence, we asked is there a connection between SMARCA5 and HDAC1,2 functions at the replication fork? We found that SMARCA5 is present on nascent DNA and importantly, loss of SMARCA5 also reduced fork velocity similar to the loss of HDAC1,2 activities.Citation34 We found that the level of SMARCA5 associated with replication origins increases whereas the level of H4K16ac at replication origins decreases when cells enter the S-phase.Citation34 Thus, HDAC1,2 activity appears to be required for the deacetylation of H4K16ac to facilitate SMARCA5-mediated remodeling of chromatin around the replication fork during S-phase. Overall, our studies demonstrated the functional interplay between a chromatin remodeler (SMARCA5), regulatory histone modifications (H4K12ac and H4K16ac) and histone deacetylases (HDAC1,2 that target H4K12ac and H4K16ac) at the replication fork. Moreover, these studies highlight the mechanism by which selective inhibition of HDAC1,2 is able to directly target chromatin structure and the chromatin remodeling around replication forks in order to obstruct the progression of DNA replication, trigger DNA damage response and selectively kill the rapidly cycling cancer cells.

HDAC1,2 is involved in DSB repair pathways that contribute to chemoresistance

Diffuse large B-cell lymphoma (DLBCL), a type of non-Hodgkin's lymphoma, is the most common lymphoid malignancy in the United States accounting for 40% of adult lymphomas.Citation39 A major advancement has been made in treating DLBCL with the addition of rituximab (an anti-CD20 monoclonal antibody) to the standard chemotherapy regimen CHOP (vincistrine, doxorubicin, cyclophosphamide, prednisone).Citation40,41 Despite overall improvement in treating DLBCL using this cocktail regimen, one-third of the patients fail standard therapy and have a poor outcome.Citation40,41 A vast majority of B-cell lymphomas are derived from the germinal centers of lymphomas.Citation42 These lymphoma cells constitutively express BCL6 oncogene, due to translocations or mutations that result in deregulated BCL6 expression.Citation42 BCL6 oncoprotein acts as a key transcriptional repressor of the ATM/ATR/p53 DNA damage-signaling pathway and facilitates hyperproliferation to provide a survival advantage to lymphoma cells.Citation43,44 The transcriptional repression is mediated through recruitment of HDACs 1, 2 and 3, via recruitment of SMRT, NCoR and BCoR to the BCL6-regulated genes.Citation45,46 Hence, small molecule inhibitors targeting the BCL6 oncoprotein or HDACs (specifically, HDAC1, 2 or HDAC3) would have therapeutic benefits in this subset of lymphoma.

Cancer cells in general, and refractory cancer cells in particular, utilize DNA repair activities as a survival mechanism to overcome the DNA damage caused by many chemotherapy drugs. Hence, inhibiting specific DNA repair pathways could induce cytotoxicity selectively in cancer cells. HDAC1,2 localize to double-strand damage sites and facilitate DNA repair.Citation32 Can we use this property of HDAC1,2 to target specific cancers resulting from increased DNA repair activities due to mutations in genes that code for proteins involved in DNA repair?

Sequencing of DLBCL lymph node biopsy samples has identified somatic mutations in EZH2 (enhancer of zeste homolog 2).Citation47 EZH2 is the enzymatic component of the polycomb repression complex 2 (PRC2) and catalyzes trimethylation of H3 at the K27 residue (H3K27me3).Citation48 In about 22% of germinal center derived lymphomas, gain-of-function mutations in the tyrosine residue (Y641) within the EZH2 catalytic SET domain is observed.Citation47 This gain-of-function mutation in EZH2 (EZH2GOF) results in high levels of H3K27me3 levels in DLBCL cells and has been implicated to promote lymphomagenesis.Citation49 In fact, a positive relationship has been observed between increased H3K27me3 and chemoresistance in ovarian cancer.Citation50 GSK126, a potent inhibitor of EZH2 activity, decreases H3K27me3 and promotes death in EZH2GOF DLBCL cells.Citation51 Thus, inhibiting EZH2 activity using small molecules to decrease the aberrant H3K27me3 is one potential strategy to overcome lymphomagenesis and/or chemoresistance in these refractory EZH2-activating mutant DLBCL cells.

H3K27me3 is also associated with DSB repair,Citation52 in addition to transcription. Hence, one could postulate that increased levels of H3K27me3 at break sites would protect these EZH2GOF DLBCL cells from DNA damage due to increased DNA repair. Hence, inhibiting EZH2-specific DNA repair pathway could cause cytotoxicity and DNA damage selectively in DLBCL cells that harbor the EZH2GOF mutation.Citation51 Our results showed that selective inhibition of HDAC1,2 increases global H3K27ac without decreasing pre-existing H3K27me3, but it decreases H3K27me3 specifically at DNA break sites, and causes cytotoxicity in the EZH2GOFDLBCL cells.Citation53 Hence, our results further indicate a new mechanism whereby HDAC1,2 inhibition induces cytotoxicity in the EZH2GOF DLBCL cells by just altering the H3K27ac/H3K27me3 ratio and the level of H3K27me3 specifically at DSB sites.Citation53 In the future, we will further decipher the cross talk between HDAC1,2 and EZH2-mediated repair signaling in DLBCL cells to nail down the precise mechanism by which HDAC1,2 inhibition overcomes the H3K27me3-mediated chemoresistance in DLBCL cells.

We have also found that EZH2GOF DLBCL cells overexpress BBAP or DTX3L (Deltex (DTX)-3-like E3 histone ubiquitin ligase),Citation53 a chromatin-modifying enzyme that has a demonstrated role in chemoresistance.Citation54 BBAP enzyme catalyzes monoubiquitination of H4K91Citation55 and hence the activity of BBAP is not counteracted by the EZH2 inhibitor, GSK126. Therefore, in addition to overcoming EZH2 and H3K27me3-mediated chemoresistance, it is also important to overcome the chemoresistance mediated by BBAP. Importantly, BBAP catalyzes H4K91 monoubiquitylation (H4K91ub1),Citation55,56 that facilitates DNA repair signaling mediated by 53BP1.Citation55 We have found that selective inhibition of HDAC1,2 results in an increase in H4K91ac, decreases H4K91ub1 levels during DNA repair following treatment with doxorubicin (a chemotherapy agent) and thereby sensitizes the refractory EZH2GOFDLBCL cells to doxorubicin.Citation53 Hence, inhibition of HDAC1,2 activities is able to impair the DNA repair processes by altering the H3K27ac-H3K27me3 switch and the H4K91ac-H4K91ubiquityl switch during DNA repair mediated by EZH2 and BBAP enzymes, respectively, in order to overcome chemoresistance in the refractory EZH2GOFDLBCL cells. Previously, knockdown of HDAC1,2 was reported to impair 53BP1 recruitment to DNA damage sites and this phenotype was attributed to the increase in H4K16ac.Citation57 Hence, increased H4K91ac and H4K16ac is likely to contribute to the observed 53BP1-mediated DSB repair defects upon inhibition of HDAC1,2. H4K91 is also linked to chromatin assembly during DNA repair in yeast,Citation58 in addition to 53BP1 signaling during DNA repair in mammalian cells. It is possible that increased H4K91ac following HDAC1,2 inhibition could prevent H2A-H2B deposition onto the H3-H4 tetramer and disrupt nucleosome assembly, which in turn might result in an unstable nucleosome/chromatin that is prone to DNA damage in cancer cells, thus providing another mechanism for HDAC1,2 inhibitor action. Overall, selective inhibition of HDAC1,2 activity using small molecules could therefore provide a novel DNA repair mechanism-based therapeutic approach in the EZH2GOFDLBCL cells by simultaneously negating two parallel and important pathways of DSB repair mediated by EZH2- and BBAP. These findings along with our DNA replication studies reveal that HDIs, more specifically selective HDAC1,2 inhibitors, act as genome-stability mechanism based cancer therapeutics.

Summary and Perspectives

In our recent publications (Bhaskara et al. (2013) and Johnson et al. (2015)),Citation34,53 we showed novel functions for HDAC1,2 during DNA replication and DNA repair that contributes to chemoresistance in cancer cells. These studies provide the basis to design a novel genome stability-targeted HDAC inhibitor therapy for a subset of cancers. HDAC1,2 inhibition causes DNA damage activation as a result of replication stress response and defective nascent chromatin structure in rapidly cycling cancer cells.Citation34 At the same time, defective DSB repair results in the failure of cancer cells to repair DNA damage resulting from collapsed replication forks and from treatment of cancer cells with chemotherapy agents, such as, doxorubicin. We are currently testing whether this mechanism based strategy would be applicable for other cancers with perturbed specific repair pathways that involve HDAC1,2. Can HDAC1,2 inhibition then provide therapeutic benefits for all cancers with increased DSB DNA repair? The answer is probably not. There are 18 HDACs in mammalian cellsCitation59 and class III Sirtuin family HDACs have been associated with DSB repair.Citation60-63 Hence, one could speculate that repair defects due to loss of HDAC1,2 activity might be compensated by other HDACs or the Sirtuin family members that are also involved in DSB repair. However, HDAC1,2 inhibition might still be a efficient way to target specific repair pathways that are addicted to HDAC1,2 and do not require Sirtuin functions. Hence, it is critical to dissect out the repair mechanisms specifically targeted by HDAC1,2 and not compensated by other HDAC family members. Moreover, chromosomal translocations that recruit HDAC1,2 as well as HDAC3 in certain cancers, cannot be targeted by inhibiting HDAC1,2 or HDAC3 alone. In these cases, combined action of HDAC1,2 and HDAC3 selective inhibitors may be necessary to achieve maximum clinical potency. Hence, mechanistic studies are going to be critical in order to avoid using selective HDAC inhibitors as a pan-cancer therapy, which might lead to unwanted toxicity with little clinical effectiveness. Moreover, Mi-2/NuRD complex that contains HDAC1,2 is required for proper heterochromatin formation and for S-phase progression.Citation64 NuRD complex also localizes to double strand break sites.Citation52 Do HDAC1,2 regulate the chromatin during DNA replication and DNA repair via the NuRD complex needs to be investigated in future. While transient inhibition of HDACs causes cytotoxicity in cancer cells, conditional deletion of Hdac3 specifically in the liver leads to hepatomegaly65 and hepatocellular carcinoma.14 Therefore, while long-term loss of HDAC function may result in secondary tumors, short-term inhibition of HDAC activity is a viable option for treating cancers. Overall, the comprehensive knowledge of how class I HDACs regulate genome stability will help us in the innovative design of monotherapy or combination therapies with inhibitors against other chromatin-modifying enzymes for certain cancers. The use of new mechanism-based strategies will augment our ongoing efforts to understand cancer epigenetics as well as provide a potential novel effective cancer therapeutic strategy.

Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

Acknowledgments

I sincerely thank my colleague in the HDAC field, Christian Seiser (Max F. Perutz Laboratories, Vienna Biocenter, Medical University of Vienna) for his critical reading of this review and valuable suggestions. I also thank Mahesh Chandrasekharan (Huntsman Cancer Institute, University of Utah) for his valuable comments on the review. I thank all the members of my lab for their valuable contributions to the work described on HDAC1,2.

Funding

Funding for the work described here was provided by development funds to S.B. from the Huntsman Cancer Institute and the Department of Radiation Oncology.

References

  • Glozak MA, Seto E. Histone deacetylases and cancer. Oncogene 2007; 26:5420-32; PMID:17694083; http://dx.doi.org/10.1038/sj.onc.1210610
  • Haberland M, Montgomery RL, Olson EN. The many roles of histone deacetylases in development and physiology: implications for disease and therapy. Nat Rev Genet 2009; 10:32-42; PMID:19065135; http://dx.doi.org/10.1038/nrg2485
  • Moser MA, Hagelkruys A, Seiser C. Transcription and beyond: the role of mammalian class I lysine deacetylases. Chromosoma 2014; 123:67-78; PMID:24170248; http://dx.doi.org/10.1007/s00412-013-0441-x
  • New M, Olzscha H, La Thangue NB. HDAC inhibitor-based therapies: can we interpret the code? Mol Oncol 2012; 6:637-56; PMID:23141799; http://dx.doi.org/10.1016/j.molonc.2012.09.003
  • Robey RW, Chakraborty AR, Basseville A, Luchenko V, Bahr J, Zhan Z, Bates SE. Histone deacetylase inhibitors: emerging mechanisms of resistance. Mol Pharm 2011; 8:2021-31; PMID:21899343; http://dx.doi.org/10.1021/mp200329f
  • Zain J, O'Connor OA. Targeting histone deacetyalses in the treatment of B- and T-cell malignancies. Invest New Drugs 2010; 28(Suppl 1):S58-78; PMID:21132350; http://dx.doi.org/10.1007/s10637-010-9591-3
  • Petrella A, Fontanella B, Carratu A, Bizzarro V, Rodriquez M, Parente L. Histone deacetylase inhibitors in the treatment of hematological malignancies. Mini Rev Med Chem 2011; 11:519-27; PMID:21561404; http://dx.doi.org/10.2174/138955711795843347
  • Giannini G, Cabri W, Fattorusso C, Rodriquez M. Histone deacetylase inhibitors in the treatment of cancer: overview and perspectives. Future Med Chem 2012; 4:1439-60; PMID:22857533; http://dx.doi.org/10.4155/fmc.12.80
  • Amann JM, Nip J, Strom DK, Lutterbach B, Harada H, Lenny N, Downing JR, Meyers S, Hiebert SW. ETO, a target of t(8;21) in acute leukemia, makes distinct contacts with multiple histone deacetylases and binds mSin3A through its oligomerization domain. Mol Cell Biol 2001; 21:6470-83; PMID:11533236; http://dx.doi.org/10.1128/MCB.21.19.6470-6483.2001
  • Lutterbach B, Westendorf JJ, Linggi B, Patten A, Moniwa M, Davie JR, Huynh KD, Bardwell VJ, Lavinsky RM, Rosenfeld MG, et al. ETO, a target of t(8;21) in acute leukemia, interacts with the N-CoR and mSin3 corepressors. Mol Cell Biol 1998; 18:7176-84; PMID:9819404
  • Wang J, Saunthararajah Y, Redner RL, Liu JM. Inhibitors of histone deacetylase relieve ETO-mediated repression and induce differentiation of AML1-ETO leukemia cells. Cancer Res 1999; 59:2766-9; PMID:10383127
  • Bhaskara S, Chyla BJ, Amann JM, Knutson SK, Cortez D, Sun ZW, Hiebert SW. Deletion of histone deacetylase 3 reveals critical roles in S phase progression and DNA damage control. Mol Cell 2008; 30:61-72; PMID:18406327; http://dx.doi.org/10.1016/j.molcel.2008.02.030
  • Bhaskara S, Hiebert SW. Role for histone deacetylase 3 in maintenance of genome stability. Cell Cycle 2011; 10:727-8; PMID:21311228; http://dx.doi.org/10.4161/cc.10.5.14866
  • Bhaskara S, Knutson SK, Jiang G, Chandrasekharan MB, Wilson AJ, Zheng S, Yenamandra A, Locke K, Yuan JL, Bonine-Summers AR, et al. Hdac3 is essential for the maintenance of chromatin structure and genome stability. Cancer Cell 2010; 18:436-47; PMID:21075309; http://dx.doi.org/10.1016/j.ccr.2010.10.022
  • Eot-Houllier G, Fulcrand G, Watanabe Y, Magnaghi-Jaulin L, Jaulin C. Histone deacetylase 3 is required for centromeric H3K4 deacetylation and sister chromatid cohesion. Genes Dev 2008; 22:2639-44; PMID:18832068; http://dx.doi.org/10.1101/gad.484108
  • Sun Z, Feng D, Fang B, Mullican SE, You SH, Lim HW, Everett LJ, Nabel CS, Li Y, Selvakumaran V, et al. Deacetylase-independent function of HDAC3 in transcription and metabolism requires nuclear receptor corepressor. Mol Cell 2013; 52:769-82; PMID:24268577; http://dx.doi.org/10.1016/j.molcel.2013.10.022
  • Lagger G, O'Carroll D, Rembold M, Khier H, Tischler J, Weitzer G, Schuettengruber B, Hauser C, Brunmeir R, Jenuwein T, et al. Essential function of histone deacetylase 1 in proliferation control and CDK inhibitor repression. Embo J 2002; 21:2672-81; PMID:12032080; http://dx.doi.org/10.1093/emboj/21.11.2672
  • Montgomery RL, Davis CA, Potthoff MJ, Haberland M, Fielitz J, Qi X, Hill JA, Richardson JA, Olson EN. Histone deacetylases 1 and 2 redundantly regulate cardiac morphogenesis, growth, and contractility. Genes Dev 2007; 21:1790-802; PMID:17639084; http://dx.doi.org/10.1101/gad.1563807
  • Trivedi CM, Luo Y, Yin Z, Zhang M, Zhu W, Wang T, Floss T, Goettlicher M, Noppinger PR, Wurst W, et al. Hdac2 regulates the cardiac hypertrophic response by modulating Gsk3 beta activity. Nat Med 2007; 13:324-31; PMID:17322895; http://dx.doi.org/10.1038/nm1552
  • Guan JS, Haggarty SJ, Giacometti E, Dannenberg JH, Joseph N, Gao J, Nieland TJ, Zhou Y, Wang X, Mazitschek R, et al. HDAC2 negatively regulates memory formation and synaptic plasticity. Nature 2009; 459:55-60; PMID:19424149; http://dx.doi.org/10.1038/nature07925
  • Zimmermann S, Kiefer F, Prudenziati M, Spiller C, Hansen J, Floss T, Wurst W, Minucci S, Gottlicher M. Reduced body size and decreased intestinal tumor rates in HDAC2-mutant mice. Cancer Res 2007; 67:9047-54; PMID:17909008; http://dx.doi.org/10.1158/0008-5472.CAN-07-0312
  • Wilting RH, Yanover E, Heideman MR, Jacobs H, Horner J, van der Torre J, DePinho RA, Dannenberg JH. Overlapping functions of Hdac1 and Hdac2 in cell cycle regulation and haematopoiesis. EMBO J 2010; 29:2586-97; PMID:20571512; http://dx.doi.org/10.1038/emboj.2010.136
  • Yamaguchi T, Cubizolles F, Zhang Y, Reichert N, Kohler H, Seiser C, Matthias P. Histone deacetylases 1 and 2 act in concert to promote the G1-to-S progression. Genes Dev 2010; 24:455-69; PMID:20194438; http://dx.doi.org/10.1101/gad.552310
  • Jamaladdin S, Kelly RD, O'Regan L, Dovey OM, Hodson GE, Millard CJ, Portolano N, Fry AM, Schwabe JW, Cowley SM. Histone deacetylase (HDAC) 1 and 2 are essential for accurate cell division and the pluripotency of embryonic stem cells. Proc Natl Acad Sci U S A 2014; 111:9840-5; PMID:24958871; http://dx.doi.org/10.1073/pnas.1321330111
  • Haberland M, Johnson A, Mokalled MH, Montgomery RL, Olson EN. Genetic dissection of histone deacetylase requirement in tumor cells. Proc Natl Acad Sci U S A 2009; 106:7751-5; PMID:19416910; http://dx.doi.org/10.1073/pnas.0903139106
  • He S, Khan DH, Winter S, Seiser C, Davie JR. Dynamic distribution of HDAC1 and HDAC2 during mitosis: association with F-actin. J Cell Physiol 2013; 228:1525-35; PMID:23280436; http://dx.doi.org/10.1002/jcp.24311
  • Senese S, Zaragoza K, Minardi S, Muradore I, Ronzoni S, Passafaro A, Bernard L, Draetta GF, Alcalay M, Seiser C, et al. Role for histone deacetylase 1 in human tumor cell proliferation. Mol Cell Biol 2007; 27:4784-95; PMID:17470557; http://dx.doi.org/10.1128/MCB.00494-07
  • Heideman MR, Wilting RH, Yanover E, Velds A, de Jong J, Kerkhoven RM, Jacobs H, Wessels LF, Dannenberg JH. Dosage-dependent tumor suppression by histone deacetylases 1 and 2 through regulation of c-Myc collaborating genes and p53 function. Blood 2013; 121:2038-50; PMID:23327920; http://dx.doi.org/10.1182/blood-2012-08-450916
  • Winter M, Moser MA, Meunier D, Fischer C, Machat G, Mattes K, Lichtenberger BM, Brunmeir R, Weissmann S, Murko C, et al. Divergent roles of HDAC1 and HDAC2 in the regulation of epidermal development and tumorigenesis. EMBO J 2013; 32:3176-91; PMID:24240174; http://dx.doi.org/10.1038/emboj.2013.243
  • Haberland M, Mokalled MH, Montgomery RL, Olson EN. Epigenetic control of skull morphogenesis by histone deacetylase 8. Genes Dev 2009; 23:1625-30; PMID:19605684; http://dx.doi.org/10.1101/gad.1809209
  • Deardorff MA, Bando M, Nakato R, Watrin E, Itoh T, Minamino M, Saitoh K, Komata M, Katou Y, Clark D, et al. HDAC8 mutations in Cornelia de Lange syndrome affect the cohesin acetylation cycle. Nature 2012; 489:313-7; PMID:22885700; http://dx.doi.org/10.1038/nature11316
  • Miller KM, Tjeertes JV, Coates J, Legube G, Polo SE, Britton S, Jackson SP. Human HDAC1 and HDAC2 function in the DNA-damage response to promote DNA nonhomologous end-joining. Nat Struct Mol Biol 2010; 17:1144-51; PMID:20802485; http://dx.doi.org/10.1038/nsmb.1899
  • Sirbu BM, Couch FB, Feigerle JT, Bhaskara S, Hiebert SW, Cortez D. Analysis of protein dynamics at active, stalled, and collapsed replication forks. Genes Dev 2011; 25:1320-7; PMID:21685366; http://dx.doi.org/10.1101/gad.2053211
  • Bhaskara S, Jacques V, Rusche JR, Olson EN, Cairns BR, Chandrasekharan MB. Histone deacetylases 1 and 2 maintain S-phase chromatin and DNA replication fork progression. Epigenetics Chromatin 2013; 6:27; PMID:23947532; http://dx.doi.org/10.1186/1756-8935-6-27
  • Milutinovic S, Zhuang Q, Szyf M. Proliferating cell nuclear antigen associates with histone deacetylase activity, integrating DNA replication and chromatin modification. J Biol Chem 2002; 277:20974-8; PMID:11929879; http://dx.doi.org/10.1074/jbc.M202504200
  • Kinner A, Wu W, Staudt C, Iliakis G. Gamma-H2AX in recognition and signaling of DNA double-strand breaks in the context of chromatin. Nucleic Acids Res 2008; 36:5678-94; PMID:18772227; http://dx.doi.org/10.1093/nar/gkn550
  • Shogren-Knaak M, Ishii H, Sun JM, Pazin MJ, Davie JR, Peterson CL. Histone H4-K16 acetylation controls chromatin structure and protein interactions. Science 2006; 311:844-7; PMID:16469925; http://dx.doi.org/10.1126/science.1124000
  • Clapier CR, Cairns BR. Regulation of ISWI involves inhibitory modules antagonized by nucleosomal epitopes. Nature 2012; 492:280-4; PMID:23143334; http://dx.doi.org/10.1038/nature11625
  • Sabattini E, Bacci F, Sagramoso C, Pileri SA. WHO classification of tumours of haematopoietic and lymphoid tissues in 2008: an overview. Pathologica 2010; 102:83-7; PMID:21171509
  • Foon KA, Takeshita K, Zinzani PL. Novel therapies for aggressive B-cell lymphoma. Adv Hematol 2012; 2012:302570; PMID:22536253; http://dx.doi.org/10.1155/2012/302570
  • Friedberg JW. Relapsed/refractory diffuse large B-cell lymphoma. Hematology Am Soc Hematol Educ Program 2011; 2011:498-505; PMID:22160081; http://dx.doi.org/10.1182/asheducation-2011.1.498
  • Ci W, Polo JM, Cerchietti L, Shaknovich R, Wang L, Yang SN, Ye K, Farinha P, Horsman DE, Gascoyne RD, et al. The BCL6 transcriptional program features repression of multiple oncogenes in primary B cells and is deregulated in DLBCL. Blood 2009; 113:5536-48; PMID:19307668; http://dx.doi.org/10.1182/blood-2008-12-193037
  • Ranuncolo SM, Polo JM, Melnick A. BCL6 represses CHEK1 and suppresses DNA damage pathways in normal and malignant B-cells. Blood Cells Mol Dis 2008; 41:95-9; PMID:18346918; http://dx.doi.org/10.1016/j.bcmd.2008.02.003
  • Ranuncolo SM, Wang L, Polo JM, Dell'Oso T, Dierov J, Gaymes TJ, Rassool F, Carroll M, Melnick A. BCL6-mediated attenuation of DNA damage sensing triggers growth arrest and senescence through a p53-dependent pathway in a cell context-dependent manner. J Biol Chem 2008; 283:22565-72; PMID:18524763; http://dx.doi.org/10.1074/jbc.M803490200
  • Huynh KD, Fischle W, Verdin E, Bardwell VJ. BCoR, a novel corepressor involved in BCL-6 repression. Genes Dev 2000; 14:1810-23; PMID:10898795
  • Parekh S, Polo JM, Shaknovich R, Juszczynski P, Lev P, Ranuncolo SM, Yin Y, Klein U, Cattoretti G, Dalla Favera R, et al. BCL6 programs lymphoma cells for survival and differentiation through distinct biochemical mechanisms. Blood 2007; 110:2067-74; PMID:17545502; http://dx.doi.org/10.1182/blood-2007-01-069575
  • Morin RD, Johnson NA, Severson TM, Mungall AJ, An J, Goya R, Paul JE, Boyle M, Woolcock BW, Kuchenbauer F, et al. Somatic mutations altering EZH2 (Tyr641) in follicular and diffuse large B-cell lymphomas of germinal-center origin. Nat Genet 2010; 42:181-5; PMID:20081860; http://dx.doi.org/10.1038/ng.518
  • Cao R, Zhang Y. The functions of E(Z)/EZH2-mediated methylation of lysine 27 in histone H3. Curr Opin Genet Dev 2004; 14:155-64; PMID:15196462; http://dx.doi.org/10.1016/j.gde.2004.02.001
  • Beguelin W, Popovic R, Teater M, Jiang Y, Bunting KL, Rosen M, Shen H, Yang SN, Wang L, Ezponda T, et al. EZH2 is required for germinal center formation and somatic EZH2 mutations promote lymphoid transformation. Cancer Cell 2013; 23:677-92; PMID:23680150; http://dx.doi.org/10.1016/j.ccr.2013.04.011
  • Chapman-Rothe N, Curry E, Zeller C, Liber D, Stronach E, Gabra H, Ghaem-Maghami S, Brown R. Chromatin H3K27me3/H3K4me3 histone marks define gene sets in high-grade serous ovarian cancer that distinguish malignant, tumour-sustaining and chemo-resistant ovarian tumour cells. Oncogene 2013; 32:4586-92; PMID:23128397; http://dx.doi.org/10.1038/onc.2012.477
  • McCabe MT, Ott HM, Ganji G, Korenchuk S, Thompson C, Van Aller GS, Liu Y, Graves AP, Della Pietra A, 3rd, Diaz E, et al. EZH2 inhibition as a therapeutic strategy for lymphoma with EZH2-activating mutations. Nature 2012; 492:108-12; PMID:23051747; http://dx.doi.org/10.1038/nature11606
  • Chou DM, Adamson B, Dephoure NE, Tan X, Nottke AC, Hurov KE, Gygi SP, Colaiacovo MP, Elledge SJ. A chromatin localization screen reveals poly (ADP ribose)-regulated recruitment of the repressive polycomb and NuRD complexes to sites of DNA damage. Proc Natl Acad Sci U S A 2010; 107:18475-80; PMID:20937877; http://dx.doi.org/10.1073/pnas.1012946107
  • Johnson DP, Spitz GS, Tharkar S, Quayle SN, Shearstone JR, Jones S, McDowell ME, Wellman H, Tyler JK, Cairns BR, et al. HDAC1,2 inhibition impairs EZH2- and BBAP- mediated DNA repair to overcome chemoresistance in EZH2 gain-of-function mutant diffuse large B-cell lymphoma. Oncotarget 2015; 6(7):4863-87.
  • Clozel T, Yang S, Elstrom RL, Tam W, Martin P, Kormaksson M, Banerjee S, Vasanthakumar A, Culjkovic B, Scott DW, et al. Mechanism-based epigenetic chemosensitization therapy of diffuse large B-cell lymphoma. Cancer Discov 2013; 3:1002-19; PMID:23955273; http://dx.doi.org/10.1158/2159-8290.CD-13-0117
  • Yan Q, Dutt S, Xu R, Graves K, Juszczynski P, Manis JP, Shipp MA. BBAP monoubiquitylates histone H4 at lysine 91 and selectively modulates the DNA damage response. Mol Cell 2009; 36:110-20; PMID:19818714; http://dx.doi.org/10.1016/j.molcel.2009.08.019
  • Yan Q, Xu R, Zhu L, Cheng X, Wang Z, Manis J, Shipp MA. BAL1 and its partner E3 ligase, BBAP, link Poly(ADP-ribose) activation, ubiquitylation, and double-strand DNA repair independent of ATM, MDC1, and RNF8. Mol Cell Biol 2013; 33:845-57; PMID:23230272; http://dx.doi.org/10.1128/MCB.00990-12
  • Tang J, Cho NW, Cui G, Manion EM, Shanbhag NM, Botuyan MV, Mer G, Greenberg RA. Acetylation limits 53BP1 association with damaged chromatin to promote homologous recombination. Nat Struct Mol Biol 2013; 20:317-25; PMID:23377543; http://dx.doi.org/10.1038/nsmb.2499
  • Ye J, Ai X, Eugeni EE, Zhang L, Carpenter LR, Jelinek MA, Freitas MA, Parthun MR. Histone H4 lysine 91 acetylation a core domain modification associated with chromatin assembly. Mol Cell 2005; 18:123-30; PMID:15808514; http://dx.doi.org/10.1016/j.molcel.2005.02.031
  • Haberland M, Montgomery RL, Olson EN. The many roles of histone deacetylases in development and physiology: implications for disease and therapy. Nature reviews. Genetics 2009; 10:32-42; PMID:19065135
  • Toiber D, Erdel F, Bouazoune K, Silberman DM, Zhong L, Mulligan P, Sebastian C, Cosentino C, Martinez-Pastor B, Giacosa S, et al. SIRT6 recruits SNF2H to DNA break sites, preventing genomic instability through chromatin remodeling. Mol Cell 2013; 51:454-68; PMID:23911928; http://dx.doi.org/10.1016/j.molcel.2013.06.018
  • Yuan Z, Seto E. A functional link between SIRT1 deacetylase and NBS1 in DNA damage response. Cell Cycle 2007; 6:2869-71; PMID:18156798; http://dx.doi.org/10.4161/cc.6.23.5026
  • Jeong J, Juhn K, Lee H, Kim SH, Min BH, Lee KM, Cho MH, Park GH, Lee KH. SIRT1 promotes DNA repair activity and deacetylation of Ku70. Exp Mol Med 2007; 39:8-13; PMID:17334224; http://dx.doi.org/10.1038/emm.2007.2
  • Kaidi A, Weinert BT, Choudhary C, Jackson SP. Human SIRT6 promotes DNA end resection through CtIP deacetylation. Science 2010; 329:1348-53; PMID:20829486; http://dx.doi.org/10.1126/science.1192049
  • Sims JK, Wade PA. Mi-2/NuRD complex function is required for normal S phase progression and assembly of pericentric heterochromatin. Mol Biol Cell 2011; 22:3094-102; PMID:21737684; http://dx.doi.org/10.1091/mbc.E11-03-0258
  • Knutson SK, Chyla BJ, Amann JM, Bhaskara S, Huppert SS, Hiebert SW. Liver-specific deletion of histone deacetylase 3 disrupts metabolic transcriptional networks. The EMBO journal 2008; 27(7):1017-28; PMID:18354499; http://dx.doi.org/10.1038/emboj.2008.51