Metal-dependent Deacetylases: Cancer and Epigenetic Regulators

Epigenetic regulation is a key factor in cellular homeostasis. Post-translational modifications (PTMs) are a central focus of this regulation as they ...
0 downloads 0 Views 877KB Size
Subscriber access provided by University of South Dakota

Review

Metal-dependent deacetylases: Cancer and epigenetic regulators Jeffrey E Lopez, Eric D Sullivan, and Carol A. Fierke ACS Chem. Biol., Just Accepted Manuscript • DOI: 10.1021/acschembio.5b01067 • Publication Date (Web): 24 Feb 2016 Downloaded from http://pubs.acs.org on February 29, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Chemical Biology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 23

1

1 22 3 43 5 64 7 5 8 96 10 7 11 12 8 13 14 9 15 10 16 11 17 12 18 19 13 20 14 21 15 22 16 23 24 17 25 26 18 27 28 19 29 20 30 31 21 32 22 33 34 23 35 24 36 37 25 38 39 26 40 41 27 42 28 43 44 29 45 46 30 47 48 31 49 32 50 33 51 52 34 53 35 54 36 55 56 37 57 38 58 59 60

ACS Chemical Biology

Metal-dependent deacetylases: Cancer and epigenetic regulators Jeffrey E. López†, Eric D. Sullivan†, Carol A. Fierke†,‡. †

Interdepartmental Program in Chemical Biology, University of Michigan, 210 Washtenaw Avenue, Ann Arbor, MI. 48109-2216, United States. ‡

Departments of Chemistry and Biological Chemistry, University of Michigan, 920 North University Avenue, Ann Arbor, MI, 48109-2216, United States Abstract Epigenetic regulation is a key factor in cellular homeostasis. Post-translational modifications (PTMs) are a central focus of this regulation as they function as signaling markers within the cell. Lysine acetylation is a dynamic, reversible PTM that has garnered recent attention due to alterations in various types of cancer. Acetylation levels are regulated by two opposing enzyme families: lysine acetyltransferases (KATs) and histone deacetylases (HDACs). HDACs are key players in epigenetic regulation and have a role in the silencing of tumor suppressor genes. The dynamic equilibrium of acetylation makes HDACs attractive targets for drug therapy. However, substrate selectivity and biological function of HDAC isozymes is poorly understood. This review outlines the current understanding of the roles and specific epigenetic interactions of the metal-dependent HDACs in addition to their roles in cancer. Keywords Post-translational modifications: The enzyme-catalyzed changes that occur to proteins following translation. These include the covalent addition of functional groups. Nε-Acetylation: The post-translational modification of lysine residues by enzymatic addition of an acetyl group. Histone Deacetylase: Family of enzymes responsible for catalyzing the hydrolysis of Nε-acetyl lysine residues. HDACi: Histone deacetylase inhibitors, primarily small molecules that chelate active site divalent metal ions. SAHA: Suberoylanilide hydroxamic acid, a potent pan-HDAC inhibitor currently used as treatment for T-cell lymphoma. Cancer: A disease involving uncontrolled cell growth. Introduction Regulation of gene expression by post-translational modifications (PTMs) is important. Many of these modifications are dynamic and reversible. PTMs include methylation, phosphorylation, ubiquitination, acetylation and biotinylation, among others. Some of these PTMs, such as acetylation and methylation, were originally identified as modifications on the tails of core histones to regulate access to DNA by alterations of the chromatin structure. (1) Lysine acetylation has garnered increasing interest in recent years, with a trend in publication rate that rivals that of phosphorylation. (2) A major advance in the acetylation field has been the transition from analysis of acetylation (hyperand hypoacetylation) of specific sites in histone tails to defining and understanding the numerous 1 ACS Paragon Plus Environment

ACS Chemical Biology

39 1 40 2 41 3 4 42 5 43 6 44 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 45 37 38 46 39 47 40 41 48 42 49 43 50 44 51 45 46 52 47 53 48 54 49 50 55 51 56 52 57 53 58 54 55 59 56 57 60 58 61 59 60

Page 2 of 23

proteins that are modified through acetylation events. Acetylation has been shown to compete and cooperate with other PTMs, such as ubiquitination,(3) and affect specific protein-protein interactions, protein stability and protein-DNA interactions. (4) Identifying these interactions will lead to a better understanding of the acetylome, the collection of non-histone proteins that undergo acetylation/deacetylation, and the role of acetylation in cell regulation, growth and homeostasis (Figure 1).

Figure 1 – Acetylation is critical for regulation and proper cellular function. Lysine acetyltransferases (KATs) and histone deacetylases (HDACs) maintain acetylation at an optimal level. HATs and HDACs regulate essential processes such as DNA repair, chromatin and actin remodeling and proteins that serve as checkpoints during the cell cycle.

Acetylation is an enzymatically-catalyzed and reversible PTM in which an acetyl group is attached to the Nε-position of a lysine side chain. Lysine acetyltransferases catalyze acetylation of lysine side chains using acetyl-CoA as a cofactor, and metal-dependent histone deacetylases (HDACs) catalyze hydrolysis of the acetyl moiety to regenerate lysine and free acetate. Traditionally, HDACs have been described as transcriptional repressors since they change the recruitment and intramolecular interactions of many proteins, from bromodomain-containing proteins, MEF2-binding proteins and domains to histone tails, in addition to aiding in local chromatin compaction. (1) However, acetylation and deacetylation of many non-histone proteins has now been observed. In fact, currently, over 3600 acetylation sites have been identified in mammalian proteins, with many more being discovered through proteomics and computational and modeling analyses. (5) In light of these many non-histone HDAC targets, a more suitable name for these enzymes is acetyl-lysine deacetylases, or acKDACs. Aberrant regulation of protein acetylation has been observed in various types of cancers including prostate, (8) breast, (9) and colon (10) among others, in addition to a variety of diseases, such as Cornelia de Lange Syndrome (CdLS), (6) Huntington’s disease (7) and inflammation. (17) The increasingly evident role and overexpression of HDACs in cancer has made them an interesting anticancer target. Current research shows that there are multiple mechanisms by which HDACs affect 2 ACS Paragon Plus Environment

Page 3 of 23

62 1 63 2 64 3 4 65 5 66 6 67 7 8 68

ACS Chemical Biology

cancer development, such as inducing growth arrest, differentiation, senescence and death of cancerous cells. (14) (19) Several HDAC inhibitors (HDACi) have been developed to combat these mechanisms, however, the current clinically used compounds do not possess isozyme selectivity. Three pan-HDAC inhibitors – suberoylanilide hydroxamic acid (SAHA), (11) Romidepsin (cyclic peptide), (11) and belinostat (hydroxamic acid) (12) – have been approved by the FDA for the treatment of T-cell lymphomas and a fourth inhibitor, Panobinostat (hydroxamic acid), has recently been approved for multiple myeloma treatment. (13)

9 10 69 11 70 12 71 13 72 14 15 73 16 74 17 75 18 19

76 20 77 21 22 78 23 79 24 80 25 26 81 27 82 28 83 29 84 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Studies have recently demonstrated that pan-HDACi can also be used to increase the effectiveness of anti-cancer immunotherapy treatments. In one mechanism T-cell survival is enhanced due to prevention of activation-induced cell death by lymphocytes. (14) (15) However, these effects vary significantly and can produce a variety of non-desirable side effects as a consequence. Additionally, HDACis have been used to enhance vaccine strategies; namely, mice vaccinated with melanoma cells that have been pre-treated with trichostatin A (TSA) show an increase in immune response towards additional tumors, effectively enhancing their tumor specific immunity mechanisms. (16) (17) HDACs are divided into four different classes based on their sequence homology to yeast orthologs. (18) Class I, which shares homology with Rpd3, consists of HDACs 1, 2, 3 and 8. Class II, with homology to Hda1, can be divided into two subclasses – IIa (HDACs 4, 7 and 9) and IIb (HDACs 6 and 10). Class III, with homology to the Sir2 family, is known as the sirtuins and utilizes NAD+ as a cofactor. Class IV, which shares homology with both class I and class II, consists of HDAC11. Classes I, II and IV are metal-dependent HDACs that use a metal-water as the nucleophile during catalysis, which is activated via a general acid-base mechanism (Figure 2). (19) In addition, the activity of HDACs can be regulated by monovalent metal cations. (20) Metal-dependent HDACs are the focus of this review.

3 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 85 35 36 86 37 87 38 88 39 40 89 41 90 42 91 43 92 44 45 93 46 94 47 95 48 49 96 50 97 51 52 98 53 54 99 55 56 57 58 59 60

Page 4 of 23

Figure 2 – General acid/base mechanism of catalysis utilized by metal-dependent deacetylases. (19) The conserved Asp-Asp-His triad coordinates a divalent metal ion that coordinates the metal-water nucleophile. His143 acts as both a general acid and a general base while Tyr306 and His142 stabilize the oxyanion intermediate.

Metal-dependent HDACs share common sequence motifs (Figure 3), including a deacetylase domain that is comprised of an arginase-deacetylase fold consisting of a multi-strand β-sheet surrounded by α-helices and a divalent metal ion cofactor coordinated by an Asp-Asp-His triad. (23) Class I HDACs possess a deacetylase domain with little sequence variation and are localized mainly in the nucleus, (18) with the exception of HDAC8 which has been observed in the cytoplasm of smooth muscle cells. (21) Class II HDACs are shuttled between the nucleus and the cytoplasm and possess additional domains, such as MEF2 binding domains. HDAC6 has the largest array of domains with two deacetylase domains and a zinc finger protein binding domain. (18) Class II HDACs are expressed ubiquitously through the cell and generally have lower catalytic activity in vitro when compared to class I enzymes. Finally, HDAC11, the only class IV metal-dependent deacetylase, possesses characteristics from both class I and II enzymes, and is expressed in higher abundance in specific tissues such as brain, heart and kidney. (22)

4 ACS Paragon Plus Environment

Page 5 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 100 40 41 101 42 102 43 103 44 45 104 46 47 105 48 106 49 107 50 108 51 52 109 53 110 54 111 55 56 112 57 113 58 59 60

ACS Chemical Biology

Figure 3 – Schematic comparison of the metal-dependent deacetylases. All isozymes possess a common deacetylase domain. Class I isozymes are highly conserved and small. Class IIa isozymes have specific MEF2-binding domains in addition to their conserved deacetylase domains. Class IIb contain unique domains unlike the other classes. Only the deacetylase domain has been identified in class IV.

Crystal structures have been solved for HDAC1 (PDB: 4BKX), HDAC2 (PDB: 3MAX), HDAC3 (PDB: 4A69), HDAC4 (PDB: 2VQM), HDAC6 (PDB: 3PHD), HDAC7 (PDB: 3C0Y; catalytic domain only), HDAC8 (PDB: 2V5W), and HDAC9 (PDB: 1TQE). The isozymes lacking crystal structures are HDAC5, HDAC10 and HDAC11. One of the most prevalent questions in the HDAC field is the substrate selectivity of each isozyme. There are over 3600 validated mammalian acetylation sites (5) and 18 deacetylase isozymes. Thus, defining the substrate pool for each isozyme is essential for understanding the biological functions of acetylation. Additionally, the substrate specificity of HDACs might be regulated by oxidative stress, protein-protein interactions and other PTMs. Elucidating HDAC substrate specificity and regulation will provide insight into the function and control of acetylation sites in proteins. Much research in the field has focused on class I HDACs, thus, only the most recent work on that class is cited in this review. Class IIa, IIb and IV have fewer published studies and the literature is therefore covered more extensively. 5 ACS Paragon Plus Environment

ACS Chemical Biology

114 1 115 2 116 3 4

117 5 6

118 7 8 119 9 120 10 121 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 23

In this review, we highlight functions of metal-dependent deacetylases with regard to epigenetic regulation and homeostasis and how these modifications play a role in cell proliferation and growth in various cancers, in addition to other, as of now, unknown roles. Class I HDACs The class I HDAC subfamily is disregulated in cancers and is the best studied subfamily of the metaldependent deacetylases. Overexpression of this subclass has been observed in a variety of cancers such as gastric, (24) breast, (9) prostate, (8) and colon, (10) as well as T-cell (25) and Hodgkin’s lymphoma. (26) Class I protein-protein interactions are currently the most well understood (Table 1). HDAC Potential interactions Disease/condition isozyme HDAC1 Plasmogenic activators Breast cancer (27) (28)

Observable phenotype Cell cycle arrest, inhibition and apoptosis.

growth

IKF2 (29) Transcription BBX (29)

factor

ADNP (29) ARID5B (29) C16orf87 (29) Estrogen and progesterone receptors (30) Acute promyelocytic Lower genetic stability and lymphoma (30) increase of progenitor cells. PML-RAR (30) Carcinomas (31) HDAC2 (31) Leukemogenesis acceleration SNAIL1 (31) Leukemia (31) rate. E-cadherin (31)

Thymocyte accumulation.

P53 (30) c-myc (30) HDAC2

P53/DNA (30)

Apoptosis induction in cells. Blocking of cell proliferation

Bax (36) Bcl2 (36) 6 ACS Paragon Plus Environment

Page 7 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 122 56 57 58 59 60

ACS Chemical Biology

Cyclin E2 (37) Cyclin D1 (37) CDK2 (37) HDAC1 (39)

Insulin-like factors (34) HDAC3

chronic lymphocytic TRAIL-induced apoptosis. leukemia (39) Regression of fibril formation. Peryone’s disease (40) Growth inhibition growth Decreased body size in mice

P53 (44)

Pancreatic cancer (44)

P27 (44) Bax (44) NCoR (45)

Promyelocytic (45)

leukemia Restoration of retinoic acid gene expression

PML-RARalpha (45) T-cell lymphoma (25)

HDAC8

Affects skull morphology and growth.

Childhood neuroblastoma (50) Cell apoptosis. SMC3 (6)

Cell cycle inhibition. Cornelia de syndrome (6)

Lange Improper chromatid separation.

ARID1A (52) Stops cell proliferation CSRP2BP (52)

Lung cancer (50) Cervical cancer (50) Colon cancer (50)

HDAC11 Cytokine-interleukin 10 Inflammation (105) (105) Ductal breast carcinoma (100) Cell cycle affector. Cdt1(102)

7 ACS Paragon Plus Environment

ACS Chemical Biology

1 2 3 123 4 124 5 6 125 7 126 8 127 9 128 10 11 129 12 130 13 131 14 15 132 16 133 17 134 18 135 19 20 136 21 22 137 23 138 24 139 25 140 26 27 141 28 142 29 143 30 31 144 32 33 145 34 146 35 147 36 148 37 38 149 39 150 40 41 151 42 152 43 153 44 45 154 46 155 47 156 48 49 157 50 158 51 159 52 160 53 54 161 55 162 56 163 57 58 59 60

Page 8 of 23

Table 1 – Summary of cancer and disease related HDAC-protein interactions and associated phenotypes for class I (1, 2, 3, and 8) and class IV (11) HDACs.

In the majority of cases, upregulation of HDAC1 is associated with poor cancer prognosis. (91) Silencing of HDAC1 using siRNA knockouts results in cell cycle arrest, growth inhibition and induction of apoptosis in breast cancer cells (27) and induction of a plasminogen activator in neuroblastoma cells, increasing their invasive capacity. (28) Mass proteomic analyses have revealed additional HDAC1 protein-protein interactions, ranging from short-lived interaction proteins, such as IKF2, HMG box transcription factor BBX and activity-dependent neuroprotector homeobox protein ADNP to proteins with methylation-related functions such as ARID5B to previously uncharacterized zincbinding proteins and domains, such as C16orf87. (29) Additionally, HDAC1 has been demonstrated to interact with the oncogene fusion protein PML-RAR, a protein involved in the pathogenesis of acute promyelocytic lymphoma (APL); in particular, HDAC1 diminishes the tumorigenic activity of PML-RAR by blocking differentiation, impairing genetic stability and increasing the renewal of progenitor cells. However, HDAC1 expression enhances cell survival once they are differentiated, thus leading to a dual role in cancerous tissue. (30) Immunodeficient mice have been used to evaluate the role of HDAC1 in tumor formation using teratomas. In these models, HDAC1 deficiency leads to partially undifferentiated carcinomas, upregulation of HDAC2, elevated levels of SNAIL1 expression and delocalization of E-cadherin. (31) Knockouts of HDAC1 and 2 show dramatic acceleration of leukemogenesis in preleukemic mice. HDAC1 knockouts also led to deletion of p53 and c-myc overexpression. (30) Additionally, Dovey and collaborators demonstrated that knockouts of key components of the HDAC1/2 deacetylase complex (Sin3A and Mi2) that decrease HDAC activity in T-cells perturb the differentiation of thymocytes into mature T lymphocytes. (32) Similarly, mice knockouts of HDAC1/2 demonstrate that the loss of HDAC activity leads to the accumulation of thymocytes in addition to blocking early thymic development. (33) The previously described protein-protein interactions have led to the proposal that some HDACs function in large deacetylase complexes. HDAC1 and HDAC2, together with histone binding proteins RBBP4 and RBBP7, DNA/chromatin recognition motifs and transcription factors form the core deacetylase complexes that help localize HDACs 1 and 2 to chromatin (35) Expression of HDAC1 and HDAC3 correlate with both estrogen and progesterone receptor expression and have been proposed as prognostic markers in breast cancer tumors. (35) HDAC2 is overexpressed in lung cancer tissue and mesenchymal tumors, suggesting that it is an effector for the disease. Silencing of HDAC2 via siRNA leads to an increase in p53 DNA binding activity, Bax activation and Bcl2 suppression (36) These changes in Bax activation and Bcl2 suppression are consistent with suppressed expression of cyclin E2, cyclin D1 and CDK2, blocking cell proliferation and inducing apoptosis. (37) Truncations of HDAC2 have been detected in a large number of cancers (38) and knockouts of both HDAC1 and HDAC2 prompt TRAIL-induced apoptosis in chronic lymphocytic leukemia (CLL), indicating a possible level of cooperativity between these two isozymes. (39) Recent studies have also shown that both HDAC2 silencing and inhibition induce regression of fibrotic formation in Peyronie’s disease models. (40) (41) Using mutant fibroblasts that are HDAC2-deficient, Zimmerman and collaborators demonstrated a lack of response to insulin-like growth factors (IGFs) when compared to wild type cells, showing a potential link between HDACs and IGFs. (34) Mice models lacking HDAC1 and with a single HDAC2 allele develop a lethal pathology within 3 months, likely due to neoplastic transformation of immature T cells. (32) Additionally, mutant 8 ACS Paragon Plus Environment

Page 9 of 23

164 1 165 2 3

166 4 167 5 6 168 7 169 8 170 9 10 171 11 172 12 173 13 174 14 15 175 16 176 17 177 18 178 19 20 179 21 180 22 181 23 24 182

ACS Chemical Biology

mice with an inactive HDAC2 mutant exhibit a 25 percent decrease in body size and reduced cell number and thickness of intestinal mucosa. (34) HDAC3, along with HDAC1 and HDAC2, is often expressed in high levels in renal, colorectal and gastric cancer. (42) (43) High expression of HDAC3 has also been observed in eight different pancreatic cancer cell lines and potentially generates a post-induction repression of p53, p27 and Bax genes through deacetylation of K9 of histone H3. (44) Knockouts of HDAC3 in promyelocytic leukemia cells restores retinoic acid dependent gene expression, primarily due to loss of the interactions between HDAC3, the nuclear co-repressor NCoR and PML-RARalpha fusion protein. HDAC3 interactions with the nuclear co-repressor NCoR to block transcription are enhanced by PMLRARalpha binding to DNA. (45) The best understood example of HDAC3 function is repression of retinoic acid and thyroid hormone receptors, which can modulate p53 expression. (44) Additionally, HDAC3 depletion in mouse liver upregulates lipogenic genes and causes histone hyperacetylation, leading to hepatostaetosis. (46) However, expression of inactive HDAC3 mutant proteins in these knockout mice almost completely rescues the metabolic and gene transcription alterations, suggesting that HDAC3 plays important non-catalytic roles, such as protein-protein interactions. Consistent with this, mice knockouts of the nuclear co-repressor NcoR, an essential part of the HDAC3 deacetylase complex, exhibit metabolic and transcriptional effects resembling those of mice without hepatic HDAC3, demonstrating that interaction with NcoR is essential for the deacetylaseindependent function of HDAC3. (46)

25 26 183 27 184 28 185 29 186 30 31 187 32 188 33 189 34 190 35 36 191 37 192 38 193 39 40 194 41 195 42 196 43 197 44 45 198 46 199 47 200 48 49

201 50 202 51 52

203 53 54

204 55 56 205 57 206 58 59 60

HDAC8 is the best biochemically characterized HDAC isozyme to date. (19) (20) (47) (48) HDAC8 is the only class I isozyme that is localized to both the cytoplasm and the nucleus and that is not observed in large, multi-protein complexes in vivo (49) HDAC8 is overexpressed in childhood neuroblastoma (50) and T-cell lymphoma. (25) Knockouts of HDAC8 produce skull morphology and growth complications in mice models (18) and stop cell proliferation in lung, colon and cervical cancer cells. (50) Point mutations in HDAC8 have been observed in patients with symptoms similar to the Cornelia de Lange Syndrome (CdLS). Lack of deacetylation of SMC3 in the cohesin complex has been implicated as a contributor to this disease, inhibiting the cell cycle, disrupting proper chromatid separation and causing debilitating mental and physical abnormalities. (6) Currently, work on HDAC8 has focused on mass spectrometry and co-immunoprecipitation studies using the HDAC8 specific inhibitor - PCI-34051 – and have provided insight into potential protein substrates and interaction partners. (29) (52) Knockouts and inhibition of HDAC8 have been shown to induce apoptosis in T-cell lymphoma and leukemia cell lines. (25) Using mice xenograft models of oncogene-amplified neuroblastoma, Rettig and collaborators demonstrated that selective inhibition of HDAC8 shows antineuroblastoma activity without significant toxicity and induces cell cycle arrest and differentiation both in vivo and in vitro. (53) Additionally, the combined treatment with HDACi and retinoic acid enhanced cell differentiation, demonstrating that inhibition of HDAC isozymes can be combined with differentiation-inducing agents to target tumors. (53) Investigations of HDAC8 specificity are poised to provide insight into the role of protein-protein interactions in determining substrate specificity of metal-dependent deacetylases. Class II HDACs Class II HDACs were discovered in the early 2000s. (54) (55) (56) (57) These proteins are significantly larger than both class I and class IV HDACs due to N-terminal and C-terminal tails and/or domains attached to the canonical deacetylase domain. This class is subdivided into two subfamilies: 9 ACS Paragon Plus Environment

ACS Chemical Biology

207 1 208 2 209 3 4 210 5 211 6 212 7 8 213 9 214 10 11 215 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 23

Class IIa and IIb. Class IIa consists of HDAC 4, 5, 7 and 9 and class IIb is comprised of HDAC 6, and 10. These isozymes differ from class I HDACs in that the additional N-terminal domains interact with transcription factors and target genes, such as the MEF2 proteins, a family of transcription factors that are key regulators of cellular differentiation. (56) Recruitment of class II HDACs by MEF2 proteins to protein complexes can heavily alter the protein acetylation landscape due to blocking interactions with acetylation complexes such as CREBBP/EP300 in non-Hodgkin lymphoma. (58) (59) Class II HDACs are localized to both the cytoplasm and the nucleus. Additionally, both upregulation and downregulation of these enzymes have severe repercussions in various types of cancers. (Table 2)

HDAC Potential interactions isozyme HDAC4 p21WAF/Cip1 (62)

Disease/condition

Observable phenotype

Breast cancer (68)

Affects cell proliferation.

miR-1 and miR-22 (70) (71) Waldenstrom’s macroglobulinemia (68) Melanoma (69) STAT1 (64) Ovarian cancer (64)

Increased cellular survival.

Chondrosarcoma (74)

Growth and metastasis of cells.

HIF-1 (73) VEGF (74) Cell growth regulation Androgen receptor (75)

B-cell differentiation is reduced. Solid and hematological malignancies (76)

miR-155 (76) Bcl-6 (76) MEF2 (76) HDAC5

Caspase-3-like (77)

protein High-risk medulloblastoma Cell death (77) Hepatocellular carcinoma (78) Induces cell proliferation

P14/TBX3 (78) Breast cancer (78)

Transport of HDAC5

Protein kinase C (79) Dephosphorylation of HDAC5 Protein phosphatase PP2A (83) MEF2 (76)

Chromatin remodeling 10 ACS Paragon Plus Environment

Page 11 of 23

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 216 46 47 217 48 49 218 50 51 219 52 220 53 54 221 55 222 56 223 57 224 58 59 60

ACS Chemical Biology

14-3-3 cytoplasmic anchoring proteins (83) HDAC7

Estrogen receptor alpha Breast cancer (84) (84)

Cell growth.

Reprimo (84) Increased estrogen alpha activity.

ACTN4 (87)

Nur77 (87) HDAC9

T-cell and programing (87)

receptor

B-cell Inhibition of apoptosis.

T-cell programing (90)

Limits T-cell response activity. Decreased T-cell proliferation

ATCD/TRI29 (91) HDAC6

HSP90 (91)

Acute myeloma leukemia Increased cell proliferation (91) Prostate cancer (91) Growth and transcriptional activation. Multiple myeloma (91) Acute myeloma leukemia (93) Breast cancer (94)

Nitric Oxide synthases (95) Alzheimer’s disease (97)

Protein aggregation misfolding

and

α-tubulin (95) HDAC10 TXNIP (98)

Gastric cancer (98)

Regulation of reactive oxygen species and apoptosis trigger.

Table 2 – Summary of cancer and disease related HDAC-protein interactions and associated phenotypes for class IIa (4,5,7, and 9) and class IIb (6 and 10) HDACs

Class IIa Class IIa HDACs are expressed in specific tissues and organs, demonstrating their importance in growth and development. (18) The N-terminal domains of these proteins have conserved serine resides that can undergo phosphorylation by kinases and regulate their cellular location. (60) A unique feature of class IIa HDACs is their lack of measurable catalytic activity toward acetylated histones, despite their highly conserved deacetylase domain. The enzymatic activity of these HDACs can be enhanced through recruitment into multi-protein complexes, some containing HDAC3. (61) 11 ACS Paragon Plus Environment

ACS Chemical Biology

225 1 226 2 227 3 4 228 5 229

6 7 230 8 231 9 10 232 11 233 12 234 13 235 14 15 236 16 237 17 238 18 239 19 20 240 21 241 22 242 23 24 243 25 244 26 245 27 246 28 29 247 30 248 31 32 249 33 250 34 251 35 36 252 37 253 38 254 39 40 255 41 256 42 257 43 258 44 45 259 46 260 47 261 48 262 49 50 263 51 264 52 265 53 54 266 55 56 267 57 268 58 269 59 60

Page 12 of 23

Class IIa HDACs have been proposed to be regulators and adaptors of repressor proteins and several studies have established a link between class IIa HDAC expression and various types of cancers. (62) (63) (64) Additionally, HDACs 4, 5, and 7 promote protein sumoylation and these isozymes have been suggested to have small ubiquitin-like modifier (SUMO) E3 ligase activity. (65) (66) (67) HDAC4 expression is highly upregulated in breast cancer samples and, along with HDAC9, overexpressed in patients with Waldenstrom’s macroglobulinemia. (68) However, downregulation and misfunction of HDAC4 has also been related to cancer development; deletions of HDAC4 have been observed in melanoma cell lines. (69) HDAC4 has been proposed as a target for microRNAs (miRNAs) due to the overexpression in the absence of miR-1 and miR-22. (70) (71) HDAC4, along with other class I and class II HDACs, has also been shown to help to directly regulate the expression of cyclin-dependent kinase inhibitors p21WAF/Cip1 (62) and affect cell proliferation. In ovarian cancer cells, HDAC4 is overexpressed and catalyzes deacetylation of the transcription factor STAT1, leading to increased cell survival. (64) HDAC4 overexpression also increases survival of prostate cancer cells by stabilizing HIF-1, a transcription factor that responds to decreases in oxygen and mediates the cellular response to hypoxia. (73) Downregulation of HDAC4 has been linked to increased levels of VEGF expression in chondrosarcoma cells, enabling growth and metastasis of these cells. (74) Additionally, HDAC4 repression has been linked to repression of transcription of the androgen receptor, which inhibits the induction of prostate specific antigens that regulate cell growth. (75) Also, studies of transgenic mice models with deletion of miRNA155, one of the most overexpressed miRNAs in solid and hematological malignancies, have shown that miR-155 targets HDAC4. Particularly, miR-155 blocks the interaction between HDAC4 and Bcl6, a proto-oncogene and transcriptional suppressor that downregulates miR-155 in mice. In support of this, B-cell differentiation is highly reduced when HDAC4 is expressed. (76) HDAC5 and HDAC9 are both overexpressed in high-risk medulloblastoma patients, showing a likely relationship between their expression levels and patient survival. (77) Knockouts of both HDAC5 and HDAC9 promote cell death by increasing the activity of caspase-3-like. (77) HDAC3 and HDAC5 have been linked to hepatocellular carcinoma by inhibition of these enzymes with the HDAC pan-inhibitor panobinostat. (78) In breast cancer cells, HDAC5 regulates the expression of p14 in association with TBX3, which in turn induces cell proliferation. (78) HDAC5 shuttling between the cytoplasm and the nucleus via phosphorylation catalyzed by protein kinase C has been implicated in axon regeneration. Export out of the nucleus allows for increased nuclear acetylation and activation of pro-regenerative genes. (79) (80) HDAC5 has also been linked to the behavioral response to cocaine in mice models; activation of the protein phosphatase PP2A leads to dephosphorylation of HDAC5 at S279 creating a transient accumulation of HDAC5 in the nucleus. The development of the cocaine reward behavior in animal models is suppressed by chromatin remodeling caused by HDAC5 enhanced interaction with 14-3-3 cytoplasmic anchoring proteins and disrupted binding interaction with MEF2 transcription factors. (83) In addition, HDAC5 has been demonstrated to be necessary for replication fork formation as well as maintaining and assembling heterochromatin structure in cancer cells. (82) Finally, loss of HDAC5 activity impairs memory function but has little to no effect on Alzheimer’s disease pathogenesis in mouse models for amyloid pathology. (83) HDAC7 contributes to cell growth by associating with the estrogen receptor alpha (ERα) and inhibiting the expression of Reprimo, a cell cycle inhibitor in breast cancer cells. (84) Studies have shown that siRNA silencing of HDAC7 enhanced the expression of Reprimo and induced apoptosis leading to the proposal that interaction of ERα with HDAC7 promotes cancer cell proliferation and growth. Other 12 ACS Paragon Plus Environment

Page 13 of 23

270 1 271 2 272 3 4 273 5 274 6 275 7 8 276 9 277 10 278 11 279 12 13

280 14 15 281 16 282 17 283 18 284 19 20 285 21 286 22 287 23 24 288

ACS Chemical Biology

studies have reported association of HDAC7 with actinin alpha 4 (ACTN4), a nuclear receptor gene coactivator, increasing ERα transcriptional activity. Phosphorylation and localization of HDAC7 are critical for activation and regulation of the transcriptional machinery activation and regulation. (87) HDAC7 has also been linked to B-cell and T-cell programming, as well as autoimmune disease in humans. (85) HDAC7 is highly expressed in the nucleus of thymocytes at the CD4+ and CD8+ dual stage, altering both positive and negative growth regulators. (86) For example, HDAC7 represses transcription of Nur77, an orphan nuclear receptor responsible for apoptosis activation due to recruitment of HDAC7 by MEF2 proteins. (87) HDAC7 is highly abundant in mature B-cells but significantly downregulated upon conversion into macrophages. (88) The mechanism of HDAC7 regulation by protein-protein interactions remains to be established. HDAC9 has also been linked to the immune system. In particular, it is highly overexpressed in regulatory T-cells, which limit the T-cell immune response. Regulatory T-cells from HDAC9-deficient mice have more repressive activity than wild type. (90) Additionally, HDAC9-deficient mice exhibit decreased T-cell proliferation and higher inflammation. HDAC9 has also been implicated in obesity and adipogenesis, as expression is downregulated in fat storing cells and tissues. Finally, HDAC9 catalyzes deacetylation of ATDC/TRI29 protein, an ataxia telangiectasia group D-complementing protein in acute myeloma patients which inhibits p53 activity, leading to the upregulation of genes implicated in cell proliferation. (91) In this case, HDAC9 acts as a tumor suppressor though the full extent of its function is not fully understood.

25 26 289 27 290 28 29 291 30 292 31 293 32 33 294 34 295 35 296 36 297 37 38 298 39 299 40 300 41 42 301 43 302 44 303 45 304 46 47 305 48 306 49 50 307 51 308 52 309 53 310 54 311 55 56 312 57 313 58 314 59 60

Class IIb HDAC6 contains two deacetylase catalytic domains and a C-terminal zinc finger and is located exclusively in the cytoplasm. HDAC6 has been implicated in catalysis of the deacetylation of Hsp90 leading to inhibition of transcriptional activation and growth in prostate cancer cells. (91) Inhibition of HDAC6 has anti-proliferation effects in a variety of cancerous tissues, including multiple myeloma, (91) and pediatric acute myeloid leukemia. (93) Increased expression of HDAC6 has been observed in advanced stage cancers as compared to early stage (94), particularly in breast cancer. This suggests that HDAC6 levels could be used as an indicator for breast cancer progression and prognosis. Recently, HDAC6 has been observed to be a target for nitrosylation catalyzed by nitric oxide synthases. Nitrosylation repressed the catalytic activity of HDAC6, leading to increased acetylation of α-tubulin, the best validated HDAC6 substrate. (95) Deacetylation of α-tubulin affects protein-protein interactions and has been implicated in the cellular response to protein aggregation and misfolding. (96) Interestingly, mice models lacking HDAC6 are cognitively normal. In Alzheimer’s disease models reduction of endogenous HDAC6 levels leads to a restoration of learning and memory along with α-tubulin acetylation. Loss of HDAC6 renders neurons resistant to amyloid-βmediated impairment. (97) Additionally, our current level of knowledge about HDACs and inhibitors has allowed researchers to begin using targeted approaches in developing new HDACi based therapies. A recent example of this is seen in the coupling of HDACi’s to genetically engineered viruses in an effort to target and attack tumor cells. (107) The use of engineered viruses, known as oncolytic viral (OV) therapy has shown potential in FDA trials (107), however, such an approach inherently relies on the effectiveness with which the virus is able to enter host cells and replicate. HDACi coupled to OVs have demonstrated an increase in both the initial phase of OV infection and subsequent viral replication. (107) This has been taken a step further with researchers linking HDAC6, known to interact with tubulin, to the success of this therapy. (108) Inhibition of HDAC6 leads to an increase in the acetylation profile of tubulin and down-regulation of beta interferon (IFN-β), potentially hindering the 13 ACS Paragon Plus Environment

ACS Chemical Biology

Page 14 of 23

315 1 316 2

cell’s antiviral machinery. (108) (109) Indeed, upon inhibition of HDAC6, OV trials on glioma cells showed enhanced replication and entry of the virus into the nucleus. (108)

3 317 4 318 5 6 319 7 320 8 321 9 322 10 11 323 12 324

HDAC10, one of the least studied deacetylases, can be shuttled between the nucleus and the cytoplasm, (18) however, its substrate set remains unknown. Recent studies show that HDAC10 is important for the regulation of reactive oxygen species in gastric cancer cells; inhibition of HDAC10 using siRNA causes an increase in reactive oxygen species, triggering the apoptotic pathway via the induced released of cytochrome c. (98) Additionally, knockouts of HDAC10 in human gastric cancer cells significantly increase mRNA expression levels of thioredoxin-interaction protein (TXNIP) which acts as a cellular antioxidant in gastric cancer cells. However, the mechanism of TXNIP regulation by HDAC10 requires further investigation.

13 14 325 15 16 326 17 327 18 328 19 20 329 21 330 22 331 23 332 24 25 333 26 27 334 28 335 29 336 30 31 337 32 338 33 339 34 340 35 36 341 37 342 38 343 39 344 40 41 345 42 43 346 44 45 347 46 47 348 48 349 49 350 50 351 51 352 52 353 53 54 354 55 355 56 356 57 357 58 59 60

Class IV HDACs HDAC11, the most recently discovered isozyme, is the sole member of the Class IV HDAC subfamily. (22) At 39 KDa, HDAC11 is the smallest isozyme. Sequence alignments suggest that HDAC11 is more closely aligned with Class I HDACs (28% sequence identity to HDAC8) than Class II with retention of highly conserved residues in the active site and in the mono- and divalent metal binding sites seen in other metal-dependent deacetylases. HDAC11 sequence alignments identify one significant sequence change, an aspartate (D101 in HDAC8) to asparagine, in a residue located on the flexible L2 loop, near the entrance to the active site tunnel. (104) HDAC11 expression is tissue specific, with the greatest expression occurring in the brain, heart, kidneys, skeletal muscle, and testis. (22) Studies of murine brain development suggest a role for HDAC11 in the formation of mature oligodendrocytes. (99) Overexpression of HDAC11 in RAW264.7 cells is associated with a decrease in mRNA levels of the anti-inflammatory cytokine interleukin-10, indicating a possible role for HDAC11 in inflammatory and autoimmune diseases. (105) Furthermore, mRNA analysis uncovered a link between HDAC11 and cancer; mRNA levels for HDAC11 are in the top 1% of differentially overexpressed genes in ductal breast carcinoma when compared to healthy breast tissue. (100) Finally, the DNA replication factor Cdt1 is a potential HDAC11 substrate. Cdt1 is integral in recruiting mini-chromosome maintenance (MCM) helicase to DNA, which is required for DNA replications during the cell cycle. To maintain a single copy of DNA per cell, Cdt1 must be inhibited after loading MCM in the G1 phase. (102) Cdt1 is an acetylated protein that coimmunoprecipitates with HDAC11. (102) Finally, HDAC11 knockout mice are viable, but they exhibit increased cell proliferation and secrete higher levels of IL-2, TNF, and IFN-y than WT mice. (106) Conclusions Lysine acetylation/deacetylation is a dynamic, reversible post-translational modification with a defined role in histone modification and a growing pool of non-histone substrates that are critical for epigenetic regulation, DNA repair, cell cycle regulation and cancer growth and proliferation. The HDAC family of enzymes catalyzes deacetylation of both histones and non-histone proteins to maintain acetylation homeostasis that is critical for cell regulation and survival. Changes in acetylation patterns have become a key feature of various types of cancer and many of these changes in acetylation levels are due to increased expression and/or misregulation of HDACs. Due to the link between HDACs and cancer prognosis and survival, these enzymes have become an attractive target for drug development, as shown by the development and FDA approval of pan-HDAC inhibitors and the current development of novel immuno- and OV therapies. To further our understanding of the cellular function of HDACs, huge strides are currently being made in discovering HDAC specific 14 ACS Paragon Plus Environment

Page 15 of 23

358 1 359 2 360 3 361 4 5 362

ACS Chemical Biology

binding partners and substrates using a variety of methods, from knockouts and mass spectrometric techniques to in vivo mammalian cell work. The current body of literature on HDACs has shed light on the multiple roles of these enzymes in both transcriptional regulation and in protein-protein interactions, particularly with respect to their roles in disease, including multiple cancers, Alzheimer’s disease, and many others.

6

363 7 8 364 9 365 10 366 11 367 12 368 13 14 369 15 370 16 17 371 18 19 372 20 373 21 374 22 375 23 24 376 25 26 377 27 28 378 29 30 379 31 380 32 381 33 382 34 35 383 36 384 37 385 38 39 386 40 387 41 388 42 389 43 44 390 45 391 46 392 47 393 48 49 394 50 395 51 396 52 53 397 54 398 55 399 56 57 58 59 60

However, a major limitation with the current understanding of HDACs is still the lack of a full profile of specific inhibitors for each isozyme. New isozyme-specific inhibitors will continue to provide insight into the regulation and cellular role of each isozyme, potentially without compromising cellular viability. We predict that future work will incorporate varied approaches such as mass spectrometry, co-immunoprecipitation, in vivo perturbation, in vitro functional studies, etc., to advance our understanding of the cellular role of each isozyme. Understanding the specificity of each HDAC will provide further insight into both their role in various diseases as well as their function as epigenetic regulators and markers. Acknowledgements We would like to acknowledge Dr. Carol Ann Pitcairn and Katherine R. Leng for their input in the text and context of this review. This review was supported by the National Institutes of Health (Grants NIGMS GM40602 (C.A.F.), F31-GM-116619 (J.E.L.)) and the Rackham Graduate School (Rackham Merit Fellowship (J.E.L.).

Bibliography (1) Parbin, S., Kar, S., Shilpi, A., Sengupta, D., Deb, M., Rath, S. K., and Patra, S. K. (2014) Histone deacetylases: a saga of perturbed acetylation homeostasis in cancer. J. Histochem. Cytochem. 62, 11–33. (2) Verdin, E., and Ott, M. (2014) 50 years of protein acetylation: from gene regulation to epigenetics, metabolism and beyond. Nat. Rev. Mol. Cell Biol. 16, 258–64.. (3) Caron, C., Boyault, C., and Khochbin, S. (2005) Regulatory cross-talk between lysine acetylation and ubiquitination: role in the control of protein stability. Bioessays 27, 408–15. (4) Glozak, M. A., Sengupta, N., Zhang, X., and Seto, E. (2005) Acetylation and deacetylation of non-histone proteins. Gene 363, 15–23. (5) Khoury, G. A., Baliban, R. C., and Floudas, C. A. (2011) Proteome-wide post-translational modification statistics: frequency analysis and curation of the swiss-prot database. Sci. Rep. 1. (6) Deardorff, M. A., Bando, M., Nakato, R., Watrin, E., Itoh, T., Minamino, M., Saitoh, K., Komata, M., Katou, Y., Clark, D., Cole, K. E., De Baere, E., Decroos, C., Di Donato, N., Ernst, S., Francey, L. J., Gyftodimou, Y., Hirashima, K., Hullings, M., Ishikawa, Y., Jaulin, C., Kaur, M., Kiyono, T., Lombardi, P. M., Magnaghi-Jaulin, L., Mortier, G. R., Nozaki, N., Petersen, M. B., Seimiya, H., Siu, V. M., Suzuki, Y., Takagaki, K., Wilde, J. J., Willems, P. J., Prigent, C., Gillessen-Kaesbach, G., Christianson, D. W., Kaiser, F. J., Jackson, L. G., Hirota, T., Krantz, I. D., and Shirahige, K. (2012) HDAC8 mutations in Cornelia de Lange syndrome affect the cohesin acetylation cycle. Nature 489, 313–7. (7) Jia, H., Morris, C. D., Williams, R. M., Loring, J. F., and Thomas, E. A. (2015) HDAC inhibition imparts beneficial transgenerational effects in Huntington’s disease mice via altered DNA and histone methylation. Proc. Natl. Acad. Sci. 112, E56–E64.

15 ACS Paragon Plus Environment

ACS Chemical Biology

400 1 401 2 402 3 4 403 5 404 6 405 7 8 406 9 407 10 408 11 409 12 13 410 14 411 15 412 16 413 17 18 414 19 415 20 416 21 22 417 23 418 24 419 25 420 26 27 421 28 422 29 423 30 31 424 32 425 33 426 34 427 35 428 36 37 429 38 430 39 431 40 432 41 42 433 43 434 44 435 45 46 436 47 437 48 438 49 439 50 51 440 52 441 53 442 54 443 55 56 57 58 59 60

Page 16 of 23

(8) Halkidou, K., Gaughan, L., Cook, S., Leung, H. Y., Neal, D. E., and Robson, C. N. (2004) Upregulation and nuclear recruitment of HDAC1 in hormone refractory prostate cancer. Prostate 59, 177–89. (9) Zhang, Z., Yamashita, H., Toyama, T., Sugiura, H., Omoto, Y., Ando, Y., Mita, K., Hamaguchi, M., Hayashi, S. I., and Iwase, H. (2004) HDAC6 expression is correlated with better survival in breast cancer. Clin. Cancer Res. (10) Wilson, A. J., Byun, D.-S., Popova, N., Murray, L. B., L’Italien, K., Sowa, Y., Arango, D., Velcich, A., Augenlicht, L. H., and Mariadason, J. M. (2006) Histone deacetylase 3 (HDAC3) and other class I HDACs regulate colon cell maturation and p21 expression and are deregulated in human colon cancer. J. Biol. Chem. 281, 13548–58. (11) Mottamal, M., Zheng, S., Huang, T. L., and Wang, G. (2015) Histone deacetylase inhibitors in clinical studies as templates for new anticancer agents. Molecules 20, 3898–941. (12) Poole, R. M. (2014) Belinostat: first global approval. Drugs 74, 1543–54. (13) Fenichel, M. P. (2015) FDA approves new agent for multiple myeloma. J. Natl. Cancer Inst. 107, djv165. (14) Gallagher, S., Tiffen, J., and Hersey, P. (2015) Histone Modifications, Modifiers and Readers in Melanoma Resistance to Targeted and Immune Therapy. Cancers (Basel). 7, 1959–1982. (15) Cao, K., Wang, G., Li, W., Zhang, L., Wang, R., Huang, Y., Du, L., Jiang, J., Wu, C., He, X., Roberts, A. I., Li, F., Rabson, A. B., Wang, Y., and Shi, Y. (2015) Histone deacetylase inhibitors prevent activation-induced cell death and promote anti-tumor immunity. Oncogene 34, 5960–70. (16) Khan, A.N.; Magner, W.J.; Tomasi, T.B. (2004) An epigenetically altered tumor cell vaccine. Cancer Immunol. Immunother., 53, 748–754. (17) Khan, A.N.; Magner, W.J.; Tomasi, T.B. (2007) An epigenetic vaccine model active in the prevention and treatment of melanoma. J. Transl. Med., 5, 64. (18) Haberland, M., Montgomery, R. L., and Olson, E. N. (2009) The many roles of histone deacetylases in development and physiology: implications for disease and therapy. Nat. Rev. Genet. 10, 32–42. (19) Gantt, S. M. L., Decroos, C., Lee, M. S., Gullett, L. E., Bowman, C. M., Christianson, D. W., and Fierke, C. A. (2016) General Base-General Acid Catalysis in Human Histone Deacetylase 8. Biochemistry. 55: 820-832 (20) Gantt, S. L., Joseph, C. G., and Fierke, C. A. (2009) Activation and Inhibition of Histone Deacetylase 8 by Monovalent Cations. J. Biol. Chem. 285, 6036–6043. (21) Waltregny, D., Glénisson, W., Tran, S. L., North, B. J., Verdin, E., Colige, A., and Castronovo, V. (2005) Histone deacetylase HDAC8 associates with smooth muscle alpha-actin and is essential for smooth muscle cell contractility. FASEB J. 19, 966–8. (22) Gao, L., Cueto, M. A., Asselbergs, F., and Atadja, P. (2002) Cloning and functional characterization of HDAC11, a novel member of the human histone deacetylase family. J. Biol. Chem. 277, 25748–55. (23) Smith, B. C., and Denu, J. M. (2009) Chemical mechanisms of histone lysine and arginine modifications. Biochim. Biophys. Acta 1789, 45–57. (24) Spiegel, S., Milstien, S., and Grant, S. (2012) Endogenous modulators and pharmacological inhibitors of histone deacetylases in cancer therapy. Oncogene 31, 537–51.

16 ACS Paragon Plus Environment

Page 17 of 23

444 1 445 2 446 3 4 447 5 448 6 449 7 8 450 9 451 10 452 11 453 12 13 454 14 455 15 16 456 17 457 18 458 19 459 20 21 460 22 461 23 462 24 463 25 464 26 27 465 28 466 29 467 30 468 31 469 32 470 33 34 471 35 472 36 473 37 474 38 475 39 40 476 41 477 42 478 43 44 479 45 480 46 481 47 482 48 49 483 50 484 51 485 52 53 486 54 487 55 488 56 489 57 58 490 59 60

ACS Chemical Biology

(25) Balasubramanian, S., Ramos, J., Luo, W., Sirisawad, M., Verner, E., and Buggy, J. J. (2008) A novel histone deacetylase 8 (HDAC8)-specific inhibitor PCI-34051 induces apoptosis in T-cell lymphomas. Leukemia 22, 1026–34. (26) Adams, H., Fritzsche, F. R., Dirnhofer, S., Kristiansen, G., and Tzankov, A. (2010) Class I histone deacetylases 1, 2 and 3 are highly expressed in classical Hodgkin’s lymphoma. Expert Opin. Ther. Targets 14, 577–84. (27) Senese, S., Zaragoza, K., Minardi, S., Muradore, I., Ronzoni, S., Passafaro, A., Bernard, L., Draetta, G. F., Alcalay, M., Seiser, C., and Chiocca, S. (2007) Role for histone deacetylase 1 in human tumor cell proliferation. Mol. Cell. Biol. 27, 4784–95. (28) Pulukuri, S. M. K., Gorantla, B., and Rao, J. S. (2007) Inhibition of histone deacetylase activity promotes invasion of human cancer cells through activation of urokinase plasminogen activator. J. Biol. Chem. 282, 35594–603. (29) Joshi, P., Greco, T. M., Guise, A. J., Luo, Y., Yu, F., Nesvizhskii, A. I., and Cristea, I. M. (2013) The functional interactome landscape of the human histone deacetylase family. Mol. Syst. Biol. 9, 672. (30) Santoro, F., Botrugno, O. A., Dal Zuffo, R., Pallavicini, I., Matthews, G. M., Cluse, L., Barozzi, I., Senese, S., Fornasari, L., Moretti, S., Altucci, L., Pelicci, P. G., Chiocca, S., Johnstone, R. W., and Minucci, S. (2013) A dual role for Hdac1: oncosuppressor in tumorigenesis, oncogene in tumor maintenance. Blood 121, 3459–68. (31) Lagger, S., Meunier, D., Mikula, M., Brunmeir, R., Schlederer, M., Artaker, M., Pusch, O., Egger, G., Hagelkruys, A., Mikulits, W., Weitzer, G., Muellner, E. W., Susani, M., Kenner, L., and Seiser, C. (2010) Crucial function of histone deacetylase 1 for differentiation of teratomas in mice and humans. EMBO J. 29, 3992–4007. (32) Dovey, O. M., Foster, C. T., Conte, N., Edwards, S. A., Edwards, J. M., Singh, R., Vassiliou, G., Bradley, A., and Cowley, S. M. (2013) Histone deacetylase 1 and 2 are essential for normal T-cell development and genomic stability in mice. Blood 121, 1335–44. (33) Heideman, M. R., Wilting, R. H., Yanover, E., Velds, A., de Jong, J., Kerkhoven, R. M., Jacobs, H., Wessels, L. F., and Dannenberg, J.-H. (2013) Dosage-dependent tumor suppression by histone deacetylases 1 and 2 through regulation of c-Myc collaborating genes and p53 function. Blood 121, 2038–50. (34) Zimmermann, S., Kiefer, F., Prudenziati, M., Spiller, C., Hansen, J., Floss, T., Wurst, W., Minucci, S., and Göttlicher, M. (2007) Reduced body size and decreased intestinal tumor rates in HDAC2-mutant mice. Cancer Res. 67, 9047–54. (35) Yao, Y.-L., and Yang, W.-M. (2003) The metastasis-associated proteins 1 and 2 form distinct protein complexes with histone deacetylase activity. J. Biol. Chem. 278, 42560–8. (36) Krusche, C. A., Wülfing, P., Kersting, C., Vloet, A., Böcker, W., Kiesel, L., Beier, H. M., and Alfer, J. (2005) Histone deacetylase-1 and -3 protein expression in human breast cancer: a tissue microarray analysis. Breast Cancer Res. Treat. 90, 15–23. (37) Jung, K. H., Noh, J. H., Kim, J. K., Eun, J. W., Bae, H. J., Xie, H. J., Chang, Y. G., Kim, M. G., Park, H., Lee, J. Y., and Nam, S. W. (2012) HDAC2 overexpression confers oncogenic potential to human lung cancer cells by deregulating expression of apoptosis and cell cycle proteins. J. Cell. Biochem. 113, 2167–77. (38) Ropero, S., Fraga, M. F., Ballestar, E., Hamelin, R., Yamamoto, H., Boix-Chornet, M., Caballero, R., Alaminos, M., Setien, F., Paz, M. F., Herranz, M., Palacios, J., Arango, D., Orntoft, T. F., Aaltonen, L. A., Schwartz, S., and Esteller, M. (2006) A truncating mutation of HDAC2 in human cancers confers resistance to histone deacetylase inhibition. Nat. Genet. 38, 566–9. 17 ACS Paragon Plus Environment

ACS Chemical Biology

491 1 492 2 493 3 4 494 5 495 6 496 7 8 497 9 498 10 499 11 500 12 13 501 14 502 15 503 16 504 17 18 505 19 506 20 507 21 22 508 23 509 24 510 25 511 26 27

512 28 29 513 30 514 31 515 32 516 33 517 34 35 518 36 519 37 520 38 39 521 40 522 41 523 42 524 43 44 525 45 526 46 527 47 48 528 49 529 50 530 51 531 52 53 532 54 533 55 56 534 57 535 58 59 60

Page 18 of 23

(39) Inoue, S., Mai, A., Dyer, M. J. S., and Cohen, G. M. (2006) Inhibition of histone deacetylase class I but not class II is critical for the sensitization of leukemic cells to tumor necrosis factor-related apoptosis-inducing ligand-induced apoptosis. Cancer Res. 66, 6785– 92. (40) Kwon, K.-D., Choi, M. J., Park, J.-M., Song, K.-M., Kwon, M.-H., Batbold, D., Yin, G. N., Kim, W. J., Ryu, J.-K., and Suh, J.-K. (2014) Silencing histone deacetylase 2 using small hairpin RNA induces regression of fibrotic plaque in a rat model of Peyronie’s disease. BJU Int. 114, 926–36. (41) Ryu, J.-K., Kim, W.-J., Choi, M.-J., Park, J.-M., Song, K.-M., Kwon, M.-H., Das, N.-D., Kwon, K.-D., Batbold, D., Yin, G.-N., and Suh, J.-K. (2013) Inhibition of histone deacetylase 2 mitigates profibrotic TGF-β1 responses in fibroblasts derived from Peyronie’s plaque. Asian J. Androl. 15, 640–5. (42) Adams, H., Fritzsche, F. R., Dirnhofer, S., Kristiansen, G., and Tzankov, A. (2010) Class I histone deacetylases 1, 2 and 3 are highly expressed in classical Hodgkin’s lymphoma. Expert Opin. Ther. Targets 14, 577–84. (43) Fritzsche, F. R., Weichert, W., Röske, A., Gekeler, V., Beckers, T., Stephan, C., Jung, K., Scholman, K., Denkert, C., Dietel, M., and Kristiansen, G. (2008) Class I histone deacetylases 1, 2 and 3 are highly expressed in renal cell cancer. BMC Cancer 8, 381. (44) Jiao, F., Hu, H., Yuan, C., Jin, Z., Guo, Z., Wang, L., and Wang, L. (2014) Histone deacetylase 3 promotes pancreatic cancer cell proliferation, invasion and increases drugresistance through histone modification of P27, P53 and Bax. Int. J. Oncol. 45, 1523–30. (45) Atsumi, A., Tomita, A., Kiyoi, H., and Naoe, T. (2006) Histone deacetylase 3 (HDAC3) is recruited to target promoters by PML-RARalpha as a component of the N-CoR co-repressor complex to repress transcription in vivo. Biochem. Biophys. Res. Commun. 345, 1471–80. (46) Sun, Z., Feng, D., Fang, B., Mullican, S. E., You, S.-H., Lim, H.-W., Everett, L. J., Nabel, C. S., Li, Y., Selvakumaran, V., Won, K.-J., and Lazar, M. A. (2013) Deacetylase-independent function of HDAC3 in transcription and metabolism requires nuclear receptor corepressor. Mol. Cell 52, 769–82. (47) Gantt, S. L., Gattis, S. G., and Fierke, C. A. (2006) Catalytic activity and inhibition of human histone deacetylase 8 is dependent on the identity of the active site metal ion. Biochemistry 45, 6170–8. (48) Dowling, D. P., Gattis, S. G., Fierke, C. A., and Christianson, D. W. (2010) Structures of metal-substituted human histone deacetylase 8 provide mechanistic inferences on biological function. Biochemistry 49, 5048–56. (49) Yang, X.-J., and Seto, E. (2008) The Rpd3/Hda1 family of lysine deacetylases: from bacteria and yeast to mice and men. Nat. Rev. Mol. Cell Biol. 9, 206–18. (50) Oehme, I., Deubzer, H. E., Wegener, D., Pickert, D., Linke, J.-P., Hero, B., KoppSchneider, A., Westermann, F., Ulrich, S. M., von Deimling, A., Fischer, M., and Witt, O. (2009) Histone deacetylase 8 in neuroblastoma tumorigenesis. Clin. Cancer Res. 15, 91–9. (51) Vannini, A., Volpari, C., Filocamo, G., Casavola, E. C., Brunetti, M., Renzoni, D., Chakravarty, P., Paolini, C., De Francesco, R., Gallinari, P., Steinkühler, C., and Di Marco, S. (2004) Crystal structure of a eukaryotic zinc-dependent histone deacetylase, human HDAC8, complexed with a hydroxamic acid inhibitor. Proc. Natl. Acad. Sci. U. S. A. 101, 15064–9. (52) Olson, D. E., Udeshi, N. D., Wolfson, N. A., Pitcairn, C. A., Sullivan, E. D., Jaffe, J. D., Svinkina, T., Natoli, T., Lu, X., Paulk, J., McCarren, P., Wagner, F. F., Barker, D., Howe, E., 18 ACS Paragon Plus Environment

Page 19 of 23

536 1 537 2 538 3 539 4 5 540 6 541 7 542 8 543 9 544 10 11 545 12 546 13 547 14 548 15 549 16 17 550 18 551 19 20 552 21 553 22 554 23 555 24 25 556 26 557 27 558 28 559 29 30 560 31 561 32 562 33 34 563 35 564 36 565 37 566 38 39 567 40 568 41 569 42 570 43 44 571 45 572 46 573 47 48 574 49 575 50 576 51 577 52 53 578 54 579 55 580 56 57 581 58 59 60

ACS Chemical Biology

Lazzaro, F., Gale, J. P., Zhang, Y.-L., Subramanian, A., Fierke, C. A., Carr, S. A., and Holson, E. B. (2014) An unbiased approach to identify endogenous substrates of “histone” deacetylase 8. ACS Chem. Biol. 9, 2210–6. (53) Rettig, I., Koeneke, E., Trippel, F., Mueller, W. C., Burhenne, J., Kopp-Schneider, A., Fabian, J., Schober, A., Fernekorn, U., von Deimling, A., Deubzer, H. E., Milde, T., Witt, O., and Oehme, I. (2015) Selective inhibition of HDAC8 decreases neuroblastoma growth in vitro and in vivo and enhances retinoic acid-mediated differentiation. Cell Death Dis. 6, 1657. (54) Grozinger, C. M., Hassig, C. A., and Schreiber, S. L. (1999) Three proteins define a class of human histone deacetylases related to yeast Hda1p. Proc. Natl. Acad. Sci. U. S. A. 96, 4868–73. (55) Wang, A. H., Bertos, N. R., Vezmar, M., Pelletier, N., Crosato, M., Heng, H. H., Th’ng, J., Han, J., and Yang, X. J. (1999) HDAC4, a human histone deacetylase related to yeast HDA1, is a transcriptional corepressor. Mol. Cell. Biol. 19, 7816–27. (56) Dequiedt, F., Kasler, H., Fischle, W., Kiermer, V., Weinstein, M., Herndier, B. G., and Verdin, E. (2003) HDAC7, a thymus-specific class II histone deacetylase, regulates Nur77 transcription and TCR-mediated apoptosis. Immunity 18, 687–98. (57) Mahlknecht, U., Schnittger, S., Ottmann, O. G., Schoch, C., Mosebach, M., Hiddemann, W., and Hoelzer, D. (2000) Chromosomal organization and localization of the human histone deacetylase 5 gene (HDAC5). Biochim. Biophys. Acta 1493, 342–8. (58) Morin, R. D., Mendez-Lago, M., Mungall, A. J., Goya, R., Mungall, K. L., Corbett, R. D., Johnson, N. A., Severson, T. M., Chiu, R., Field, M., Jackman, S., Krzywinski, M., Scott, D. W., Trinh, D. L., Tamura-Wells, J., Li, S., Firme, M. R., Rogic, S., Griffith, M., Chan, S., Yakovenko, O., Meyer, I. M., Zhao, E. Y., Smailus, D., Moksa, M., Chittaranjan, S., Rimsza, L., BrooksWilson, A., Spinelli, J. J., Ben-Neriah, S., Meissner, B., Woolcock, B., Boyle, M., McDonald, H., Tam, A., Zhao, Y., Delaney, A., Zeng, T., Tse, K., Butterfield, Y., Birol, I., Holt, R., Schein, J., Horsman, D. E., Moore, R., Jones, S. J. M., Connors, J. M., Hirst, M., Gascoyne, R. D., and Marra, M. A. (2011) Frequent mutation of histone-modifying genes in non-Hodgkin lymphoma. Nature 476, 298–303. (59) Han, A., He, J., Wu, Y., Liu, J. O., and Chen, L. (2005) Mechanism of recruitment of class II histone deacetylases by myocyte enhancer factor-2. J. Mol. Biol. 345, 91–102. (60) Parra, M., and Verdin, E. (2010) Regulatory signal transduction pathways for class IIa histone deacetylases. Curr. Opin. Pharmacol. 10, 454–60. (61) Fischle, W., Dequiedt, F., Hendzel, M. J., Guenther, M. G., Lazar, M. A., Voelter, W., and Verdin, E. (2002) Enzymatic activity associated with class II HDACs is dependent on a multiprotein complex containing HDAC3 and SMRT/N-CoR. Mol. Cell 9, 45–57. (62) Mottet, D., Pirotte, S., Lamour, V., Hagedorn, M., Javerzat, S., Bikfalvi, A., Bellahcène, A., Verdin, E., and Castronovo, V. (2009) HDAC4 represses p21(WAF1/Cip1) expression in human cancer cells through a Sp1-dependent, p53-independent mechanism. Oncogene 28, 243–56. (63) Wilson, A. J., Byun, D.-S., Nasser, S., Murray, L. B., Ayyanar, K., Arango, D., Figueroa, M., Melnick, A., Kao, G. D., Augenlicht, L. H., and Mariadason, J. M. (2008) HDAC4 promotes growth of colon cancer cells via repression of p21. Mol. Biol. Cell 19, 4062–75. (64) Stronach, E. A., Alfraidi, A., Rama, N., Datler, C., Studd, J. B., Agarwal, R., Guney, T. G., Gourley, C., Hennessy, B. T., Mills, G. B., Mai, A., Brown, R., Dina, R., and Gabra, H. (2011) HDAC4-regulated STAT1 activation mediates platinum resistance in ovarian cancer. Cancer Res. 71, 4412–22. 19 ACS Paragon Plus Environment

ACS Chemical Biology

582 1 583 2 584 3 4 585 5 586 6 587 7 8 588 9 589 10 590 11 591 12 13 592 14 593 15 594 16 595 17 18 596 19 597 20 598 21 22 599 23 600 24 601 25 602 26 27 603 28 604 29 605 30 31 606 32 607 33 608 34 609 35 36 610 37 611 38 612 39 613 40 41 614 42 43 615 44 616 45 617 46 618 47 48 619 49 620 50 621 51 622 52 53 623 54 624 55 625 56 626 57 58 59 60

Page 20 of 23

(65) Grégoire, S., Tremblay, A. M., Xiao, L., Yang, Q., Ma, K., Nie, J., Mao, Z., Wu, Z., Giguère, V., and Yang, X.-J. (2006) Control of MEF2 transcriptional activity by coordinated phosphorylation and sumoylation. J. Biol. Chem. 281, 4423–33. (66) Zhao, X., Sternsdorf, T., Bolger, T. A., Evans, R. M., and Yao, T.-P. (2005) Regulation of MEF2 by histone deacetylase 4- and SIRT1 deacetylase-mediated lysine modifications. Mol. Cell. Biol. 25, 8456–64. (67) Gao, C., Ho, C.-C., Reineke, E., Lam, M., Cheng, X., Stanya, K. J., Liu, Y., Chakraborty, S., Shih, H.-M., and Kao, H.-Y. (2008) Histone Deacetylase 7 Promotes PML Sumoylation and Is Essential for PML Nuclear Body Formation. Mol. Cell. Biol. 28, 5658–5667. (68) Sun, J. Y., Xu, L., Tseng, H., Ciccarelli, B., Fulciniti, M., Hunter, Z. R., Maghsoudi, K., Hatjiharissi, E., Zhou, Y., Yang, G., Zhu, B., Liu, X., Gong, P., Ioakimidis, L., Sheehy, P., Patterson, C. J., Munshi, N. C., O’Connor, O. A., and Treon, S. P. (2011) Histone deacetylase inhibitors demonstrate significant preclinical activity as single agents, and in combination with bortezomib in Waldenström’s macroglobulinemia. Clin. Lymphoma. Myeloma Leuk. 11, 152–6. (69) Stark, M., and Hayward, N. (2007) Genome-wide loss of heterozygosity and copy number analysis in melanoma using high-density single-nucleotide polymorphism arrays. Cancer Res. 67, 2632–42. (70) Datta, J., Kutay, H., Nasser, M. W., Nuovo, G. J., Wang, B., Majumder, S., Liu, C.-G., Volinia, S., Croce, C. M., Schmittgen, T. D., Ghoshal, K., and Jacob, S. T. (2008) Methylation mediated silencing of MicroRNA-1 gene and its role in hepatocellular carcinogenesis. Cancer Res. 68, 5049–58. (71) Zhang, J., Yang, Y., Yang, T., Liu, Y., Li, A., Fu, S., Wu, M., Pan, Z., and Zhou, W. (2010) microRNA-22, downregulated in hepatocellular carcinoma and correlated with prognosis, suppresses cell proliferation and tumourigenicity. Br. J. Cancer 103, 1215–20. (72) Song, B., Wang, Y., Xi, Y., Kudo, K., Bruheim, S., Botchkina, G. I., Gavin, E., Wan, Y., Formentini, A., Kornmann, M., Fodstad, O., and Ju, J. (2009) Mechanism of chemoresistance mediated by miR-140 in human osteosarcoma and colon cancer cells. Oncogene 28, 4065–74. (73) Geng, H., Harvey, C. T., Pittsenbarger, J., Liu, Q., Beer, T. M., Xue, C., and Qian, D. Z. (2011) HDAC4 protein regulates HIF1α protein lysine acetylation and cancer cell response to hypoxia. J. Biol. Chem. 286, 38095–102. (74) Sun, X., Wei, L., Chen, Q., and Terek, R. M. (2009) HDAC4 represses vascular endothelial growth factor expression in chondrosarcoma by modulating RUNX2 activity. J. Biol. Chem. 284, 21881–90. (75) Yang, Y., Tse, A. K.-W., Li, P., Ma, Q., Xiang, S., Nicosia, S. V, Seto, E., Zhang, X., and Bai, W. (2011) Inhibition of androgen receptor activity by histone deacetylase 4 through receptor SUMOylation. Oncogene 30, 2207–2218. (76) Sandhu, S. K., Volinia, S., Costinean, S., Galasso, M., Neinast, R., Santhanam, R., Parthun, M. R., Perrotti, D., Marcucci, G., Garzon, R., and Croce, C. M. (2012) miR-155 targets histone deacetylase 4 (HDAC4) and impairs transcriptional activity of B-cell lymphoma 6 (BCL6) in the Eµ-miR-155 transgenic mouse model. Proc. Natl. Acad. Sci. U. S. A. 109, 20047–52. (77) Milde, T., Oehme, I., Korshunov, A., Kopp-Schneider, A., Remke, M., Northcott, P., Deubzer, H. E., Lodrini, M., Taylor, M. D., von Deimling, A., Pfister, S., and Witt, O. (2010) HDAC5 and HDAC9 in medulloblastoma: novel markers for risk stratification and role in tumor cell growth. Clin. Cancer Res. 16, 3240–52. 20 ACS Paragon Plus Environment

Page 21 of 23

627 1 628 2 629 3 4 630 5 631 6 632 7 8 633 9 634 10 635 11 636 12 13 637

ACS Chemical Biology

(78) Lachenmayer, A., Toffanin, S., Cabellos, L., Alsinet, C., Hoshida, Y., Villanueva, A., Minguez, B., Tsai, H.-W., Ward, S. C., Thung, S., Friedman, S. L., and Llovet, J. M. (2012) Combination therapy for hepatocellular carcinoma: additive preclinical efficacy of the HDAC inhibitor panobinostat with sorafenib. J. Hepatol. 56, 1343–50. (79) Yarosh, W., Barrientos, T., Esmailpour, T., Lin, L., Carpenter, P. M., Osann, K., AntonCulver, H., and Huang, T. (2008) TBX3 is overexpressed in breast cancer and represses p14 ARF by interacting with histone deacetylases. Cancer Res. 68, 693–9. (80) Cho, Y., and Cavalli, V. (2012) HDAC5 is a novel injury-regulated tubulin deacetylase controlling axon regeneration. EMBO J. 31, 3063–78. (81) Cho, Y., Sloutsky, R., Naegle, K. M., and Cavalli, V. (2013) Injury-Induced HDAC5 Nuclear Export Is Essential for Axon Regeneration. Cell 155, 894–908.

14 15 638 16 639 17 640 18 641 19 642 20 21 643 22 644 23 645 24 25 646 26 647 27 648 28 649 29 30 650 31 651 32 652 33 653 34 35 654 36 655 37 656 38 39 657 40 658 41 659 42 660 43 44 661 45 662 46 663 47 48 664 49 665 50 666 51 667 52 668 53 54 669 55 670 56 671 57 672 58 59 60

(82) Taniguchi, M., Carreira, M. B., Smith, L. N., Zirlin, B. C., Neve, R. L., and Cowan, C. W. (2012) Histone deacetylase 5 limits cocaine reward through cAMP-induced nuclear import. Neuron 73, 108–20. (83) Agis-Balboa, R. C., Pavelka, Z., Kerimoglu, C., and Fischer, A. (2013) Loss of HDAC5 impairs memory function: implications for Alzheimer’s disease. J. Alzheimers. Dis. 33, 35–44. (84) Peixoto, P., Castronovo, V., Matheus, N., Polese, C., Peulen, O., Gonzalez, A., Boxus, M., Verdin, E., Thiry, M., Dequiedt, F., and Mottet, D. (2012) HDAC5 is required for maintenance of pericentric heterochromatin, and controls cell-cycle progression and survival of human cancer cells. Cell Death Differ. 19, 1239–52. (85) Malik, S., Jiang, S., Garee, J. P., Verdin, E., Lee, A. V, O’Malley, B. W., Zhang, M., Belaguli, N. S., and Oesterreich, S. (2010) Histone deacetylase 7 and FoxA1 in estrogenmediated repression of RPRM. Mol. Cell. Biol. 30, 399–412. (86) Liu, J. Z., Hov, J. R., Folseraas, T., Ellinghaus, E., Rushbrook, S. M., Doncheva, N. T., Andreassen, O. A., Weersma, R. K., Weismüller, T. J., Eksteen, B., Invernizzi, P., Hirschfield, G. M., Gotthardt, D. N., Pares, A., Ellinghaus, D., Shah, T., Juran, B. D., Milkiewicz, P., Rust, C., Schramm, C., Müller, T., Srivastava, B., Dalekos, G., Nöthen, M. M., Herms, S., Winkelmann, J., Mitrovic, M., Braun, F., Ponsioen, C. Y., Croucher, P. J. P., Sterneck, M., Teufel, A., Mason, A. L., Saarela, J., Leppa, V., Dorfman, R., Alvaro, D., Floreani, A., OnengutGumuscu, S., Rich, S. S., Thompson, W. K., Schork, A. J., Næss, S., Thomsen, I., Mayr, G., König, I. R., Hveem, K., Cleynen, I., Gutierrez-Achury, J., Ricaño-Ponce, I., van Heel, D., Björnsson, E., Sandford, R. N., Durie, P. R., Melum, E., Vatn, M. H., Silverberg, M. S., Duerr, R. H., Padyukov, L., Brand, S., Sans, M., Annese, V., Achkar, J.-P., Boberg, K. M., Marschall, H.-U., Chazouillères, O., Bowlus, C. L., Wijmenga, C., Schrumpf, E., Vermeire, S., Albrecht, M., Rioux, J. D., Alexander, G., Bergquist, A., Cho, J., Schreiber, S., Manns, M. P., Färkkilä, M., Dale, A. M., Chapman, R. W., Lazaridis, K. N., Franke, A., Anderson, C. A., and Karlsen, T. H. (2013) Dense genotyping of immune-related disease regions identifies nine new risk loci for primary sclerosing cholangitis. Nat. Genet. 45, 670–5. (87) Youn, H. D., and Liu, J. O. (2000) Cabin1 represses MEF2-dependent Nur77 expression and T cell apoptosis by controlling association of histone deacetylases and acetylases with MEF2. Immunity 13, 85–94. (88) Kasler, H. G., and Verdin, E. (2007) Histone deacetylase 7 functions as a key regulator of genes involved in both positive and negative selection of thymocytes. Mol. Cell. Biol. 27, 5184–200. (89) Barneda-Zahonero, B., Román-González, L., Collazo, O., Rafati, H., Islam, A. B. M. M. K., Bussmann, L. H., di Tullio, A., De Andres, L., Graf, T., López-Bigas, N., Mahmoudi, T., and 21 ACS Paragon Plus Environment

ACS Chemical Biology

673 1 674 2 675 3 4 676 5 677 6 678 7 8 679 9 680 10 681 11 682 12 13 683 14 684 15 685 16 686 17 18 687 19 688 20 689 21 22 690 23 691 24 692 25 26

693 27 694 28 695 29 30 696 31 697 32 698 33 34 699 35 700 36 701 37 702 38 39 703 40 704 41 705 42 43 706 44 707 45 708 46 709 47 710 48 711 49 50 712 51 713 52 714 53 715 54 716 55 56 717 57 58 59 60

Page 22 of 23

Parra, M. (2013) HDAC7 is a repressor of myeloid genes whose downregulation is required for transdifferentiation of pre-B cells into macrophages. PLoS Genet. 9, e1003503. (90) Yuan, Z., Peng, L., Radhakrishnan, R., and Seto, E. (2010) Histone deacetylase 9 (HDAC9) regulates the functions of the ATDC (TRIM29) protein. J. Biol. Chem. 285, 39329– 38. (91) Gao, L., and Alumkal, J. (2010) Epigenetic regulation of androgen receptor signaling in prostate cancer. Epigenetics 5, 100–4. (92) Santo, L., Hideshima, T., Kung, A. L., Tseng, J.-C., Tamang, D., Yang, M., Jarpe, M., van Duzer, J. H., Mazitschek, R., Ogier, W. C., Cirstea, D., Rodig, S., Eda, H., Scullen, T., Canavese, M., Bradner, J., Anderson, K. C., Jones, S. S., and Raje, N. (2012) Preclinical activity, pharmacodynamic, and pharmacokinetic properties of a selective HDAC6 inhibitor, ACY-1215, in combination with bortezomib in multiple myeloma. Blood 119, 2579–89. (93) Xu, X., Xie, C., Edwards, H., Zhou, H., Buck, S. A., and Ge, Y. (2011) Inhibition of Histone Deacetylases 1 and 6 Enhances Cytarabine-Induced Apoptosis in Pediatric Acute Myeloid Leukemia Cells. PLoS One (Pandey, S., Ed.) 6, e17138. (94) Sakuma, T., Uzawa, K., Onda, T., Shiiba, M., Yokoe, H., Shibahara, T., and Tanzawa, H. (2006) Aberrant expression of histone deacetylase 6 in oral squamous cell carcinoma. Int. J. Oncol. 29, 117–24. (95) Okuda, K., Ito, A., and Uehara, T. (2015) Regulation of Histone Deacetylase 6 Activity via S-Nitrosylation. Biol. Pharm. Bull. 38, 1434–7. (96) Li, G., Jiang, H., Chang, M., Xie, H., and Hu, L. (2011) HDAC6 α-tubulin deacetylase: a potential therapeutic target in neurodegenerative diseases. J. Neurol. Sci. 304, 1–8. (97) Govindarajan, N., Rao, P., Burkhardt, S., Sananbenesi, F., Schlüter, O. M., Bradke, F., Lu, J., and Fischer, A. (2013) Reducing HDAC6 ameliorates cognitive deficits in a mouse model for Alzheimer’s disease. EMBO Mol. Med. 5, 52–63. (98) Lee, J.-H., Jeong, E.-G., Choi, M.-C., Kim, S.-H., Park, J.-H., Song, S.-H., Park, J., Bang, Y.-J., and Kim, T.-Y. (2010) Inhibition of histone deacetylase 10 induces thioredoxininteracting protein and causes accumulation of reactive oxygen species in SNU-620 human gastric cancer cells. Mol. Cells 30, 107–12. (99) Liu, H., Hu, Q., Kaufman, A., D’Ercole, A. J., and Ye, P. (2008) Developmental expression of histone deacetylase 11 in the murine brain. J. Neurosci. Res. 86, 537–43. (100) Deubzer, H. E., Schier, M. C., Oehme, I., Lodrini, M., Haendler, B., Sommer, A., and Witt, O. (2013) HDAC11 is a novel drug target in carcinomas. Int. J. Cancer 132, 2200–2208. (101) Wong, P. G., Glozak, M. A., Cao, T. V, Vaziri, C., Seto, E., and Alexandrow, M. (2010) Chromatin unfolding by Cdt1 regulates MCM loading via opposing functions of HBO1 and HDAC11-geminin. Cell Cycle 9, 4351–63. (102) Glozak, M. A., and Seto, E. (2009) Acetylation/deacetylation modulates the stability of DNA replication licensing factor Cdt1. J. Biol. Chem. 284, 11446–53. (103) Minamiya, Y., Ono, T., Saito, H., Takahashi, N., Ito, M., Mitsui, M., Motoyama, S., and Ogawa, J. (2011) Expression of histone deacetylase 1 correlates with a poor prognosis in patients with adenocarcinoma of the lung. Lung Cancer 74, 300–4. (104) Vannini, A., Volpari, C., Gallinari, P., Jones, P., Mattu, M., Carfí, A., De Francesco, R., Steinkühler, C., and Di Marco, S. (2007) Substrate binding to histone deacetylases as shown by the crystal structure of the HDAC8– substrate complex. EMBO Rep. 8, 879–884.

22 ACS Paragon Plus Environment

Page 23 of 23

718 1 719 2 720 3 721 4 5 722 6 723 7 724 8 725 9 726 10 11 727 12 728 13 729 14 730 15 731 16 732 17 18 733 19 734 20 21 735 22 23 736 24 25 737 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Chemical Biology

(105) Villagra, A., Cheng, F., Wang, H.-W., Suarez, I., Glozak, M., Maurin, M., Nguyen, D., Wright, K. L., Atadja, P. W., Bhalla, K., Pinilla-Ibarz, J., Seto, E., and Sotomayor, E. M. (2009) The histone deacetylase HDAC11 regulates the expression of interleukin 10 and immune tolerance. Nat. Immunol. 10, 92–100. (106) Woods, D., Woan, K., Wang, D., Yu, Y., Powers, J., Sahakian, E., Cheng, F., Wang, H., Rock-Klotz, J., Villagra, A., Pinilla-Ibarz, J., Yu, X.-Z., and Sotomayor, E. (2013) Histone deacetylase 11 is an epigenetic regulator of cytotoxic T-lymphocyte effector function and memory formation (P1404). J. Immunol. 190, 117.2. (107) Chiocca, E. A., and Rabkin, S. D. (2014) Oncolytic viruses and their application to cancer immunotherapy., Cancer Immunol Res 2, 295–300. (108) Nakashima, H., Kaufmann, J. K., Wang, P.-Y. Y., Nguyen, T., Speranza, M.-C. C., Kasai, K., Okemoto, K., Otsuki, A., Nakano, I., Fernandez, S., Goins, W. F., Grandi, P., Glorioso, J. C., Lawler, S., Cripe, T. P., and Chiocca, E. A. (2015) Histone deacetylase 6 inhibition enhances oncolytic viral replication in glioma., J. Clin. Invest. 125, 4269–80. (109) Nusinzon, I., and Horvath, C. M. (2006) Positive and negative regulation of the innate antiviral response and beta interferon gene expression by deacetylation., Mol. Cell. Biol. 26, 3106–13.

23 ACS Paragon Plus Environment