Abstract
Zinc-dependent histone deacetylases (HDACs) regulate the biological function of histone and non-histone proteins through the hydrolysis of acetyllysine side chains to yield free lysine and acetate. Certain HDAC isozymes exhibit alternative catalytic activities, such as polyamine deacetylase or lysine fatty acid deacylase activity. To date, crystal structures have been reported for class I HDACs (1, 2, 3, and 8), class IIa HDACs (4 and 7), and class IIb HDACs (6 and 10). Conserved active site residues mediate the chemistry of substrate activation and hydrolysis in these isozymes through a metal-activated water molecule assisted by general base-general acid catalysis. Upregulated HDAC activity is observed in cancer and neurodegenerative disease, and four HDAC inhibitors are currently approved for use in cancer chemotherapy. Crystal structures of HDAC-inhibitor complexes guide the design of new inhibitors with high affinity and selectivity for specific HDAC isozymes implicated in human disease.
Graphical Abstract

Introduction
The reversible acetylation of lysine side chains on the surfaces of enzymes and other proteins comprises a vital regulatory strategy in myriad cellular processes such as transcription, the cell cycle, and metabolism [1–4]. At present, the acetylome encompasses nearly 40,000 unique protein acetylation sites as classified in the PhosphoSitePlus database [5]. Accordingly, acetylation rivals phosphorylation as a ubiquitous chemical modification in the regulation of protein function [6].
The chemistry of lysine acetylation requires “writers” – acetyl transferases that utilize acetyl-CoA as a co-substrate [7, 8], and “erasers” – deacetylases that catalyze the hydrolysis of acetyllysine to yield lysine and acetate [9]. With their identification rooted in Allfrey’s pioneering discovery of histone acetylation [10], lysine deacetylases are historically referred to as histone deacetylases (HDACs). However, this name belies the wider function of these enzymes in processing both histone and non-histone protein substrates. Moreover, certain HDACs are not lysine deacetylases, but instead are lysine fatty-acid deacylases [•11, •12]; others process non-protein substrates through their function as polyamine deacetylases [••13].
The HDACs are particularly important as targets for therapeutic intervention in the treatment of cancer, neurodegenerative disease, and other disorders [14–16]. To date, four different inhibitors of metal-dependent HDACs are approved in the US for cancer chemotherapy, and each inhibitor targets the active site Zn2+ ion. Vorinostat, Belinostat, and Panobinostat contain hydroxamic acid moieties that chelate Zn2+, and Romidepsin is a cyclic depsipeptide that, in the reduced form, bears a pendant thiol group that coordinates to Zn2+. These inhibitors are relatively non-selective, i.e., they do not exhibit significant preference for the inhibition of one particular HDAC isozyme or another, so off-target side effects can result from their use [14]. Accordingly, the development of isozyme selectivity in HDAC inhibitors is a priority for current drug design efforts.
Arginase-deacetylase fold
Metal-dependent HDACs adopt an α/β fold first observed in the binuclear manganese metalloenzyme arginase [17] and subsequently observed in the histone deacetylase-like protein from Aquifex aeolicus [18]. This structural and evolutionary relationship was unexpected, since these two enzymes share insignificant amino acid sequence identity. Even so, the Zn2+ binding site of the HDACs is conserved as the Mn2+B site of the arginases [19, 20]. Thus, metal binding stoichiometry and selectivity in the arginase-deacetylase family diverged from a common metalloenzyme ancestor.
Recent phylogenetic analysis indicates 12 clades in the arginase-deacetylase family (Figure 1) [••13]. The arginase family [21] also includes agmatinases, ureohydrolases, formiminoglutamases, and even pseudo-arginases that lack metal binding sites, such as the arginase-like protein of unknown function in Trypanosoma brucei [22]. In addition to the HDAC isozymes, the deacetylase family also includes bacterial polyamine deacetylases, acetoin utilization proteins, and pseudo-deacetylases of unknown chemical function [23]. The closest relationship between the arginases and deacetylases depicted in Figure 1 is found between human agmatinase and the pseudo-deacetylase domain of guinea pig HDAC10, which share 19% amino acid sequence identity.
Figure 1: Arginase-deacetylase family of proteins.
Unrooted phylogenetic tree indicates 12 clades: arginases, pseudo-arginases (ΨARG), formiminoglutamases (FIGase) and ureohydrolases, yeast Hos3 homologues, bacterial acetylpolyamine amidohydrolases (APAH), bacterial histone deacetylase-like amidohydrolases (HDAH), class II HDACs, class I HDACs, bacterial acetoin utilization proteins (AcuC), class IV HDACs, uncharacterized protein family UPF0489, and pseudo-deacetylases (ΨDAC). Individual proteins, species, and UniProt accession numbers are listed in Supplementary Table 3 of ref. [••13]. Reprinted from ref. [••13] (Creative Commons Attribution 4.0 International License).
Metal-dependent HDAC isozymes
Since the discovery of the first mammalian HDAC [24], 11 metal-dependent HDACs have been identified and classified through phylogenetic analysis [25]: the class I isozymes HDAC1, HDAC2, HDAC3, and HDAC8; the class IIa isozymes HDAC4, HDAC5, HDAC7, and HDAC9; the class IIb isozymes HDAC6 and HDAC10; and the sole class IV isozyme, HDAC11. Parenthetically, class III HDACs are better known as sirtuins, but these are structurally and mechanistically distinct NAD+-dependent enzymes [26, 27]. X-ray crystal structures of metal-dependent HDACs reveal conservation of the α/β arginase-deacetylase fold: HDAC1 [28, ••29], HDAC2 [30], HDAC3 [••31], HDAC4 [32], HDAC6 [••33, ••34], HDAC7 [35], [HDAC8 [36, 37], and HDAC10 [••13]. Crystal structures of HDAC isozymes illuminate critical molecular features of biological function and inhibition.
To date, crystal structures have not yet been reported for HDAC5, HDAC9, or HDAC11. Among these isozymes, HDAC11 is particularly intriguing due to its role as an immunoregulator that interacts with the promoter of the gene encoding interleukin 10 (IL-10) [38]. By blocking the expression of IL-10, HDAC11 activates immune function; loss of HDAC11 leads to immune tolerance, but it is not clear how the catalytic activity of HDAC11 is involved in this process. Possibly, HDAC11 is involved in chromatin remodeling as it associates with the IL-10 promoter. Deletion of HDAC11 or treatment with small molecule inhibitors increases the immunosuppressive function of Foxp3+ T-regulatory cells, suggesting that HDAC11 may be a viable target for immunotherapy [39].
Curiously, however, HDAC11 is not a lysine deacetylase, but instead is a robust lysine fatty acid deacylase: steady-state kinetics indicate optimal catalytic efficiency for the hydrolysis of C10 decanoyllysine, C12 dodecanoyllysine, and C14 myristoyllysine [•11, •12]. Thus, the immunomodulatory function of HDAC11 may be related to its catalytic activity as a fatty acid deacylase. Homology modeling studies suggest that the long fatty acid is accommodated in an internal pocket also shared with HDAC8 [•11], consistent with the weak fatty acid deacylase activity of HDAC8 [40]. The X-ray crystal structure determination of HDAC11 will be required to fully understand these unusual structure-function relationships.
Mechanism of amide bond hydrolysis
A catalytic mechanism for the hydrolysis of acetyllysine by a metal-dependent deacetylase was first proposed with the crystal structure determination of a bacterial histone deacetylase-like protein [18]. Subsequent studies refined this mechanistic proposal, based in large part on structural and enzymological studies of HDAC8 (Figure 2). In the precatalytic Michaelis complex, the scissile carbonyl group of acetyllysine coordinates to Zn2+ and accepts a hydrogen bond from Y306. A zinc-carbonyl interaction in the Michaelis complex is consistent with the metal-dependent variation of KM when Fe2+ and Co2+ are substituted for Zn2+ [41], as well as the crystal structures of intact substrates bound to the Y306F and H143A HDAC8 variants [42, 43]. Y306 may undergo induced-fit conformational changes between “out” and “in” orientations to accommodate substrate binding, as indicated by computational and experimental studies of HDAC8 and related deacetylases [32, 44–47].
Figure 2: Proposed mechanism of acetyllysine hydrolysis catalyzed by HDAC8.
Enzymological measurements indicate that H143 serves as a single general base-general acid, while H142 remains in the positively charged imidazolium form and serves as an electrostatic catalyst. However, hydrogen bond differences for the tandem histidine pair in other HDAC isozymes may allow for dual general base-general acid function. For example, in HDAC6, H573 may serve as a general base, and H574 may serve as a general acid [••33].
Both metal coordination and hydrogen bond interactions are required to fully activate the acetyllysine carbonyl group for nucleophilic attack by a Zn2+-bound water molecule. Enzymological studies of wild-type HDAC8 and site-specific variants indicate that H143 serves as a general base to assist Zn2+ in activating the nucleophilic water molecule, the pKa of which is lowered by metal coordination; subsequently, H143 serves as a general acid to protonate the leaving amino group, thereby enabling the collapse of the tetrahedral intermediate to yield lysine and acetate [48, 49]. The imidazolium side chain of H142 serves as an electrostatic catalyst, remaining protonated throughout the catalytic cycle. Both H142 and H143 hydrogen bond with the Zn2+-bound water molecule, thereby ensuring its optimal orientation for nucleophilic attack. Product acetate may depart through a “backdoor” tunnel in the dissociation of the enzyme-product complex [50]. This tunnel may also accommodate the fatty acid product in the weak fatty acid deacylase activity measured for HDAC8 [40].
Binding interactions of the tetrahedral intermediate and its flanking transition states have been visualized through the study of reactive substrate analogues that, upon binding in active site, undergo a chemical reaction that mimics the first step of catalysis. Specifically, consider the cyclic tetrapeptide Trapoxin A (Figure 3a), the irreversible inhibitor of class I HDACs used in the isolation and identification of human HDAC1 [24]. Trapoxin A contains an amino acid bearing an unusual α,β-epoxyketone amino acid, (2S,9S)-2-amino-8-oxo-9,10-epoxydecanoic acid (L-Aoe). The ketone carbonyl group of the L-Aoe side chain is isosteric with the amide carbonyl group of acetyl-L-lysine, the HDAC substrate. The crystal structure of the HDAC8-Trapoxin A complex shows that the ketone carbonyl undergoes nucleophilic attack to yield a tetrahedral gem-diolate that mimics the tetrahedral intermediate in catalysis [••51]. The presumed oxyanion of the gem-diolate coordinates to Zn2+ and accepts a hydrogen bond from Y306, and the hydroxyl group of the gem-diolate forms hydrogen bonds with H142 and H143 (Figure 3a).
Figure 3: HDAC complexes with cyclic tetrapeptide inhibitors.
(a) Structure of the HDAC8-Trapoxin A complex determined at 1.24 Å resolution (PDB 5VI6). The α,β-epoxyketone side chain of L-Aoe binds as a gem-diolate, thereby mimicking the tetrahedral intermediate and its flanking transition states in catalysis. The flanking NH groups of L-Aoe donate hydrogen bonds to D101, a key determinant of isozyme-substrate recognition. Reproduced from ref [••51]. Copyright 2017 American Chemical Society. (b) Structure of the HDAC6-HC Toxin complex determined at 1.73 Å resolution (PDB 5EFJ). Here, too, the α,β-epoxyketone side chain binds as a gem-diolate. The backbone NH group of L-Aoe donates a hydrogen bond to S531, a key determinant of isozyme-substrate recognition. Reprinted from ref [••33].
HC Toxin is a cyclic tetrapeptide inhibitor similarly containing an L-Aoe residue like that of Trapoxin A, and the crystal structure of the HDAC6-HC Toxin complex similarly reveals the binding of the gem-diolate form of the α,β-epoxyketone side chain (Figure 3b) [••33]. Intriguingly, cocrystallization of H574A HDAC6 with an acetyllysine-containing substrate yields an electron density map showing tetrahedral instead of planar electron density for the scissile amide moiety, suggesting that the actual tetrahedral intermediate is trapped in the active site [••33]. Since the tetrahedral intermediate cannot collapse without a proton donor to the leaving amino group, the loss of general acid H574 results in the dead-end formation of the tetrahedral intermediate in much the same way as the tetrahedral gem-diolate of HC Toxin is formed and stabilized in the HDAC6 active site. Crystal structures of HDAC4 [32], HDAC6 [••33], and HDAC10 [••13] complexed with α-trifluoromethylketones similarly reveal the binding of the gem-diolate forms of these inhibitors in deacetylase active sites.
As indicated in Figure 2, key catalytic residues in the active sites of most HDACs are the tandem histidine dyad, a catalytic tyrosine, and the Zn2+ ion. However, in class IIa HDACs (4, 5, 7, and 9), the catalytic tyrosine is substituted with a histidine, and deacetylase activity is significantly compromised as a result [52]. Even so, robust lysine deacetylase activity of HDAC4 and HDAC7 can be restored by histidine-to-tyrosine substitutions [32, 35, 52]. The biological significance of this activity difference is unclear. The active sites of most zinc hydrolases contain a hydrogen bond donor in addition to the Zn2+ ion that helps polarize the scissile amide carbonyl, as first noted for the prototypical zinc hydrolases thermolysin and carboxypeptidase A [53, 54]. Possibly, the additional histidine in the active sites of class IIa HDACs confers an alternative substrate specificity that is yet to be discovered [52].
Substrate specificity
In cells, HDAC-substrate pairs can be identified using Stable Isotope Labeling of Amino Acids in Cell Culture (SILAC) proteomics coupled with the administration of a selective HDAC inhibitor [55], or the application of a genetically-encoded photocrosslinker by unnatural amino acid expression [56]. Acetylomics approaches have also been used to assess the biological effects of isozyme-selective HDAC inhibitors [57].
For enzymological studies, peptide-based assays have been developed to measure HDAC activity, allowing for analysis of substrate sequence-activity relationships. One such assay utilizes peptides of varying composition and length bearing a C-terminal acetyl-L-lysine (Aly) followed by an aminomethylcoumarin (AMC) moiety [58]. This assay exploits the fluorescence shift of the AMC moiety upon liberation from the peptide by trypsin, which only occurs after the Aly residue has been deacetylated. However, this assay only allows for variation of the peptide sequence on the amino side of the Aly residue, i.e., in the −1 position and beyond. Alternatively, peptide substrates bearing an internal Aly residue can be incubated with HDAC; after quenching, the newly liberated lysine side chain in the peptide product can be covalently modified with dansyl chloride, and the peptide product then identified and quantified using liquid chromatography-mass spectrometry [••33].
X-ray crystal structures of HDAC isozymes complexed with substrates or substrate analogues reveal the molecular basis of substrate recognition. As lysine deacetylases, HDAC6 and HDAC8 catalyze the hydrolysis of acetyllysine residues contained in polypeptide and protein substrates. The active site of each enzyme consists of a concave molecular surface surrounding a narrow cleft that accommodates the scissile acetyllysine side chain [••33, 36, 37]. Crystal structures of H143A HDAC8 and Y306F HDAC8 complexed with the p53-derived assay substrate Ac-Arg-His-Aly-Aly-AMC show that the backbone NH groups of the scissile acetyllysine residue and the AMC moiety at the +1 position donate hydrogen bonds to the carboxylate side chain of D101 (Figure 4a) [42, 43]. Corresponding hydrogen bonds are observed between HDAC8 and the NH groups flanking the α,β-epoxyketone residue of Trapoxin A (Figure 3a) [••51]. Dual hydrogen bond interactions with D101 suggest that a proline residue at the +1 position of the substrate cannot be accommodated by HDAC8, and this has been confirmed experimentally [••33]. No other hydrogen bonds are observed between the concave active site surface of HDAC8 and main chain atoms for residues in the −3 to −1 positions of the bound substrate; the guanidinium group of the arginine residue at the −3 position forms hydrogen bonds with the backbone carbonyl of I94 or the carboxylate group of E148, or makes no hydrogen bonds at all in several crystal structures [42, 43]. The relative dearth of enzyme-substrate hydrogen bond specificity ensures broad specificity for flanking amino acid sequences in protein substrates. Although no structural information is available with regard to peptide substrate binding at the +2 position or beyond, there is enzymological evidence for an exosite that recognizes the basic sequence KRHR at the +4 to +7 positions of a substrate derived from the N-terminal tail of the histone H4 protein [59].
Figure 4: Substrate binding to HDACs.
(a) Superposition of substrate complexes with Y306F HDAC8 (blue, PDB 2V5W) and H143A HDAC8 (grey, PDB 3EWF) reveals that the scissile carbonyl of acetyllysine is activated by Zn2+ coordination and a hydrogen bond with Y306. The flanking NH groups of the acetyllysine substrate donate hydrogen bonds to D101. (b) Superposition of substrate complexes with Y745F HDAC6 (light green, PDB 5EFK) and H574A HDAC6 (dark green, PDB 5EFN) similarly reveals activation of the scissile substrate carbonyl by Zn2+ coordination and a hydrogen bond with Y745. Moreover, nucleophilic attack of Zn2+-bound water at the substrate carbonyl group occurs in H574A HDAC6 to result in a tetrahedral intermediate. The backbone NH group of acetyllysine donates a hydrogen bond to S531. (c) The trifluoromethylketone analogue of N8-acetylspermidine binds as a tetrahedral gem-diolate that mimics the tetrahedral intermediate in the active site of HDAC10 (PDB 5TD7). A 310 helix unique to this isozyme (magenta) sterically constricts the active site and confers specificity for acetylpolyamine substrates.
In contrast with HDAC8, crystal structures of HDAC6 complexed with two different assay substrates, the α-tubulin-derived peptide Ac-Ser-Asp-Aly-AMC and the histone H4-derived peptide Ac-Arg-Gly-Aly-AMC, show that only a single hydrogen bond is made between the backbone NH group of the scissile acetyllysine residue and S531 (Figure 4b); the corresponding hydrogen bond is similarly observed between S531 and the NH group of the α,β-epoxyketone residue of HC Toxin [••33]. Accordingly, HDAC6 substrates can accommodate proline residues at the +1 position. In the HDAC6-substrate complexes, peptide segments preceding the scissile acetyllysine residues are disordered and hence not fully modeled in the refined crystal structures [••33]. However, the backbone carbonyl of the aspartate or glycine residue at the −1 positions of these substrates hydrogen bonds with N530. Here, too, minimal hydrogen bonding with main chain or side chain atoms ensures broad specificity for flanking amino acid sequences in protein substrates.
In contrast with other HDACs of known structure, the crystal structure of HDAC10 reveals an active site cleft sterically constricted by a 310-helix containing the highly conserved P(E,A)CE motif (Figure 4c) [••13, 60]. Consequently, peptide substrates cannot bind easily in the active site of HDAC10 due to the long, narrow active site cleft. Instead, the active site of HDAC10 readily accommodates long, slender polyamine substrates, exhibiting optimal catalytic efficiency for the deacetylation of N8-acetylspermidine. A conserved gatekeeper in the HDAC10 active site, E274, ensures electrostatic recognition of the positively charged secondary amino group of N8-acetylspermidine. Conserved glutamate and aspartate residues across the protein surface additionally contribute to the molecular recognition of the cationic substrate [••13]. HDAC10 appears to be the long-sought cytosolic N8-acetylspermidine deacetylase for which cellular activity was discovered more than 40 years ago [61, 62].
Regulation by corepressors
Several metal-dependent HDACs function in multi-protein regulatory complexes. For example, crystal structures of HDAC1 and HDAC3 complexed with corepressor proteins that activate catalysis reveal significant protein-protein interactions that sandwich inositol tetraphosphate at the interface [28, ••29, ••31]. Corepressor binding buttresses the L1 loop flanking the deacetylase active site, which reduces enzyme surface dynamics [••31]. Additionally, superposition of the structures of HDAC1 and HDAC3-corepressor complexes with the structure of an HDAC6-inhibitor complex (which lacks a corepressor) suggests that corepressor binding may slightly constrict the active site [•63]. Inositol tetraphosphate interacts with a conserved arginine side chain on the L6 loop, which forms one wall of the active site [••29]. Mutation of this arginine in the HDAC3-SMRT corepressor complex results in a dramatic loss of enzyme activity [28], suggesting that this interaction is vital for activation of these class I HDACs.
Molecular dynamics simulations of HDAC3 reveal conformational mobility for the catalytic tyrosine in the active site, which adopts the “out” conformation in the absence of corepressor and the catalytically-required “in” conformation upon the binding of corepressor [45]. Similar conformational mobility has been observed for this residue in molecular dynamics studies of HDAC8 [46, 47] and in X-ray crystallographic studies of a bacterial polyamine deacetylase [44]. Conformational mobility of the catalytic tyrosine in HDAC active sites is facilitated by its placement in a conserved glycine-rich loop [47]. Conformational control of the catalytic tyrosine in the class I HDAC active site by deacetylase-activating domains of corepressor proteins represents a simple molecular strategy for the regulation of deacetylase activity in multi-protein complexes.
Notably, HDAC8 is the only class I enzyme that does not require a corepressor for optimal activity due to structural differences in the L1 and L6 loops that allow the enzyme to adopt an active conformation without steric persuasion from a corepressor protein. Even so, active site loops in HDAC8 can adopt alternative conformations as they accommodate derivatives of the cyclic tetrapeptide inhibitor Largazole [64]; the flexibility of these loops may contribute to the significant entropic cost of inhibitor binding [65].
Inhibitor binding and isozyme selectivity
Given that HDACs are validated targets for therapeutic intervention in the treatment of cancer and other diseases, X-ray crystal structures provide valuable guidance for drug design. A key feature of an HDAC inhibitor is the functional group that coordinates to the active site Zn2+ ion. The hydroxamate group is most commonly utilized and is found in three out of four of the currently-approved HDAC-targeted drugs (Figure 5a) [14–16]. However, the hydroxamate group can be mutagenic due to degradation via the Lossen rearrangement, which generates a highly reactive isocyanate capable of alkylating cellular components (Figure 5b) [66]. This chemical reactivity can limit the utility of hydroxamate-containing drugs.
Figure 5: HDAC inhibitors.
(a) Hydroxamate-based HDAC inhibitors currently approved for cancer chemotherapy. (b) The hydroxamate moiety is susceptible to degradation through the Lossen rearrangement, which generates an electrophilic isocyanate. (c) Romidepsin is currently approved for cancer chemotherapy and is activated by reduction of the disulfide linkage. The marine natural product Largazole is structurally similar to Romidepsin and is activated by thioester hydrolysis. (d) The crystal structure of the HDAC8-Largazole complex shows that the ionized thiolate side chain coordinates to Zn2+ and accepts a hydrogen bond from Y306 (PDB 3RQD).
Macrocyclic peptides and depsipeptides comprise a fascinating class of HDAC inhibitors. These natural products contain pendant side chains that mimic the scissile acetyllysine side chain of an HDAC substrate. Romidepsin is a depsipeptide drug that, upon reduction of an intramolecular disulfide linkage, yields a side chain bearing a thiol group that coordinates to the active site Zn2+ ion. While the crystal structure of an HDAC-Romidepsin complex has not yet been reported, the crystal structures of HDAC8 complexed with Largazole and its synthetic analogues, which are structurally similar to Romidepsin, have been reported (Figure 5c,d) [64, 67].
As previously discussed, the macrocyclic tetrapeptide inhibitors HC Toxin and Trapoxin A have been studied in complexes with HDAC6 and HDAC8, respectively (Figure 3) [••33, ••51]. Hydrogen bond interactions between the NH groups flanking the L-Aoe residue are identical to those observed in structures of enzyme-substrate complexes [••33, 43] and highlight differences that contribute to inhibitor selectivity for HDAC6 over class I HDACs such as HDAC8. Specifically, D101 of HDAC8 accepts hydrogen bonds from both NH groups flanking the L-Aoe residue [••51], whereas S531 of HDAC6 accepts a hydrogen bond only from the NH group of LAoe (Figure 3) [••33]. Crystal structures of selective inhibitors bound to HDAC6 similarly reveal direct or water-mediated hydrogen bond interactions with S531, so this residue is a key determinant of isozyme selectivity [•63, 68, •69]. Additionally, calorimetric measurements indicate that inhibitor binding to HDAC6 is entropically favorable, in contrast with inhibitor binding to HDAC8 [65].
Other functional groups targeting Zn2+ coordination in HDAC active sites include α-trifluoromethylketones, which bind as gem-diolates in the active sites of HDAC4 [32], HDAC6 [••33], and HDAC10 (Figure 4c) [••13], and benzamides that chelate the Zn2+ ion of HDAC2 through the anilide nitrogen and amide carbonyl groups [30]. Most recently, structural studies of HDAC-mercaptoacetamide complexes reveal functional mimicry of hydroxamate binding to HDAC6 [70] but not HDAC8 [71]. Clearly, opportunities abound for new HDAC inhibitor designs incorporating alternative metal-binding groups, such as those exemplified in a recent review on bioinorganic fragment-based drug discovery [•72].
Concluding remarks
Recent years have witnessed rapidly increasing interest in the structural and chemical biology of the metal-dependent HDAC isozymes, particularly in view of their value as therapeutic targets for cancer, neurodegenerative disorders, and other diseases. At present, only the structures of HDAC5, HDAC9, and HDAC11 have yet to be determined. Chemical and biological studies reveal that HDACs catalyze the deacetylation of histone proteins and more, both in the nucleus and the cytoplasm of eukaryotic cells. Moreover, HDAC11 is a deacylase and not a deacetylase, with specificity for substrates bearing fatty acid acyl-lysine residues [•11, •12]. The preferred substrate of HDAC10 in vitro is not acetyllysine, but instead N8-acetylspermidine [••13]. Thus, metal-dependent HDACs exhibit an impressively broad range of specificity for small molecule and protein substrates in different cellular compartments.
Crystal structures of HDAC-inhibitor complexes reveal the molecular features underlying substrate specificity and inhibitor selectivity. Isozyme selectivity is desirable to minimize side effects resulting from the inhibition of off-target HDAC isozymes. For example, recent structural studies of HDAC6 show that the most selective inhibitors contain capping groups that make favorable binding interactions in an active site pocket formed by the L1 loop, as well as variable Zn2+ coordination denticity [•63, •69]. Moreover, selective inhibitors can be used for imaging HDAC in the human brain using positron emission tomography, which exemplifies the forefront of HDAC studies [•73]. Clearly, HDAC structural and chemical biology will continue to provide exciting new opportunities at the interface of chemistry, biology, and medicine for many years to come.
Highlights.
Zn2+-histone deacetylases (HDACs) belong to the arginase-deacetylase superfamily
HDACs hydrolyze acetyllysine, fatty acid acyllysine, and acetylpolyamine substrates
HDACs are drug targets for the treatment of cancer and neurodegenerative disease
Crystal structures of HDACs guide current drug design efforts
Acknowledgments
We thank the National Institutes of Health for grant GM49758 in support of this research.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Conflict of interest
Nothing declared.
References
Papers of particular interest, published within the period of review, have been highlighted as:
• of special interest
•• of outstanding interest
- 1.Sterner DE, Berger SL: Acetylation of histones and transcription-related factors. Microbiol Mol Biol Rev 2000, 64:435–459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Zhao S, Xu W, Jiang W, Yu W, Lin Y, Zhang T, Yao J, Zhou L, Zeng Y, Li H, et al. : Regulation of cellular metabolism by protein lysine acetylation. Science 2010, 327:1000–1004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Wang Q, Zhang Y, Yang C, Xiong H, Lin Y, Yao J, Li H, Xie L, Zhao W, Yao Y, et al. : Acetylation of metabolic enzymes coordinates carbon source utilization and metabolic flux. Science 2010, 327:1004–1007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Choudhary C, Weinert BT, Nishida Y, Verdin E, Mann M: The growing landscape of lysine acetylation links metabolism and cell signaling. Nat Rev Mol Cell Biol 2014, 15:536–550. [DOI] [PubMed] [Google Scholar]
- 5.Hornbeck PV, Zhang B, Murray B, Kornhauser JM, Latham V, Skrzypek E: PhosphoSitePlus, 2014: mutations, PTMs and recalibrations. Nucleic Acids Res 2015, 43:D512–520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Kouzarides T: Acetylation: a regulatory modification to rival phosphorylation? EMBO J 2000, 19:1176–1179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.McCullough CE, Marmorstein R: Molecular basis for histone acetyltransferase regulation by binding partners, associated domains, and autoacetylation. ACS Chem Biol 2016, 11:632–642. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Marmorstein R, Zhou MM: Writers and readers of histone acetylation: structure, mechanism, and inhibition. Cold Spring Harb Perspect Biol 2014, 6:a018762. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.de Ruijter AJ, van Gennip AH, Caron HN, Kemp S, van Kuilenburg AB: Histone deacetylases (HDACs): characterization of the classical HDAC family. Biochem J 2003, 370:737–749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Allfrey VG, Faulkner R, Mirsky AE: Acetylation and methylation of histones and their possible role in the regulation of RNA synthesis. Proc Natl Acad Sci USA 1964, 51:786–794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- •11.Kutil Z, Novakova Z, Meleshin M, Mikesova J, Schutkowski M, Barinka C: Histone deacetylase 11 is a fatty-acid deacylase. ACS Chem Biol 2018, 13:685–693.This paper reports that HDAC11 exhibits optimal catalytic activity as a lysine fatty acid deacylase.
- •12.Moreno-Yruela C, Galleano I, Madsen AS, Olsen CA: Histone deacetylase 11 is an ε-N-myristoyllysine hydrolase. Cell Chem Biol 2018, 25:849–856.This paper reports that HDAC11 exhibits optimal catalytic activity as a lysine fatty acid deacylase.
- ••13.Hai Y, Shinsky SA, Porter NJ, Christianson DW: Histone deacetylase 10 structure and molecular function as a polyamine deacetylase. Nat Commun 2017, 8:15368.This paper reports the first crystal structure of HDAC10. Additionally, this paper reports enzymological studies showing that HDAC10 is a polyamine deacetylase highly specific for substrate N8-acetylspermidine.
- 14.Falkenberg KJ, Johnstone RW: Histone deacetylases and their inhibitors in cancer, neurological diseases and immune disorders. Nat Rev Drug Disc 2014, 13:673–691. [DOI] [PubMed] [Google Scholar]
- 15.West AC, Johnstone RW: New and emerging HDAC inhibitors for cancer treatment. J Clin Invest 2014, 124:30–39. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Ma N, Luo Y, Wang Y, Liao C, Ye WC, Jiang S: Selective histone deacetylase inhibitors with anticancer activity. Curr Top Med Chem 2016, 16:415–426. [DOI] [PubMed] [Google Scholar]
- 17.Kanyo ZF, Scolnick LR, Ash DE, Christianson DW: Structure of a unique binuclear manganese cluster in arginase. Nature 1996, 383:554–557. [DOI] [PubMed] [Google Scholar]
- 18.Finnin MS, Donigian JR, Cohen A, Richon VM, Rifkind RA, Marks PA, Breslow R, Pavletich NP: Structures of a histone deacetylase homologue bound to the TSA and SAHA inhibitors. Nature 1999, 401:188–193. [DOI] [PubMed] [Google Scholar]
- 19.Ash DE, Cox JD, Christianson DW: Arginase: a binuclear manganese metalloenzyme. Metal Ions Biol. Syst 2000, 37:407–428. [PubMed] [Google Scholar]
- 20.Lombardi PM, Cole KE, Dowling DP, Christianson DW: Structure, mechanism, and inhibition of histone deacetylases and related metalloenzymes. Curr Opin Struct Biol 2011, 21:735–743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Ouzounis CA, Kyrpides NC: On the evolution of arginases and related enzymes. J Mol Evol 1994, 39:101–104. [DOI] [PubMed] [Google Scholar]
- 22.Hai Y, Kerkhoven EJ, Barrett MP, Christianson DW: Crystal structure of an arginase-like protein from Trypanosoma brucei that evolved without a binuclear manganese cluster. Biochemistry 2015, 54:458–471. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Liepe DD, Landsman D: Histone deacetylases, acetoin utilization proteins and acetylpolyamine amidohydrolases are members of an ancient protein superfamily. Nucleic Acids Res 1997, 25:3693–3697. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Taunton J, Hassig CA, Schreiber SL: A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3p. Science 1996, 272:408–411. [DOI] [PubMed] [Google Scholar]
- 25.Gregoretti IV, Lee YM, Goodson HV: Molecular evolution of the histone deacetylase family: functional implications of phylogenetic analysis. J Mol Biol 2004, 338:17–31. [DOI] [PubMed] [Google Scholar]
- 26.Yuan H, Marmorstein R: Structural basis for sirtuin activity and inhibition. J Biol Chem 2012, 287:42428–42435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Chen B, Zang W, Wang J, Huang Y, He Y, Yan L, Liu J, Zheng W: The chemical biology of sirtuins. Chem Soc Rev 2015, 44:5246–5264. [DOI] [PubMed] [Google Scholar]
- 28.Millard CJ, Watson PJ, Celardo I, Gordiyenko Y, Cowley SM, Robinson CV, Fairall L, Schwabe JW: Class I HDACs share a common mechanism of regulation by inositol phosphates. Mol Cell 2013, 51:57–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
- ••29.Watson PJ, Millard CJ, Riley AM, Robertson NS, Wright LC, Godage HY, Cowley SM, Jamieson AG, Potter BVL, Schwabe JWR: Insights into the activation mechanism of class I HDAC complexes by inositol phosphates. Nat Commun 2016, 7:11262.This paper describes the structural requirements for binding and activation of the class I HDACs 1, 2, and 3 by inositol phosphates, which bind in a pocket at the protein-protein interface of the HDAC–corepressor complex.
- 30.Bressi JC, Jennings AJ, Skene R, Wu Y, Melkus R, De Jong R, O’Connell S, Grimshaw CE, Navre M, Gangloff AR: Exploration of the HDAC2 foot pocket: synthesis and SAR of substituted N-(2-aminophenyl)benzamides. Bioorg Med Chem Lett 2010, 20:3142–3145. [DOI] [PubMed] [Google Scholar]
- ••31.Watson PJ, Fairall L, Santos GM, Schwabe JWR: Structure of HDAC3 bound to corepressor and inositol tetraphosphate. Nature 2012, 481:335–340.This paper was the first to show that an inositol phosphate binds at the protein-protein interface and regulates the assembly of a class I HDAC with its cognate co-repressor.
- 32.Bottomley MJ, Lo Surdo P, Di Giovine P, Cirillo A, Scarpelli R, Ferrigno F, Jones P, Neddermann P, De Francesco R, Steinkühler C, et al. : Structural and functional analysis of the human HDAC4 catalytic domain reveals a regulatory structural zinc-binding domain. J Biol Chem 2008, 283:26694–26704. [DOI] [PMC free article] [PubMed] [Google Scholar]
- ••33.Hai Y, Christianson DW: Histone deacetylase 6 structure and molecular basis of catalysis and inhibition. Nat Chem Biol 2016, 12:741–747.This paper reports the crystal structures of high-activity catalytic domains 2 of human and zebrafish HDAC6, as well as the low-activity catalytic domain 1 of zebrafish HDAC6. Additionally, crystal structures of several enzyme-inhibitor complexes are reported, including hydroxamate-based drugs used in cancer chemotherapy (Vorinostat (SAHA), Belinostat, and Panobinostat) as well as the cyclic tetrapeptide inhibitor HC Toxin.
- ••34.Miyake Y, Keusch JJ, Wang L, Saito M, Hess D, Wang X, Melancon BJ, Helquist P, Gut H, Matthias P: Structural insights into HDAC6 tubulin deacetylation and its selective inhibition. Nat Chem Biol 2016, 12:748–754.This paper reports the crystal structure of a di-domain construct containing catalytic domains 1 and 2 of zebrafish HDAC6, revealing the conformation of the interdomain linker. Additionally, crystal structures of individual catalytic domains complexed with hydroxamate-based inhibitors Trichostatin A and Nexturastat A are reported.
- 35.Schuetz A, Min J, Allali-Hassani A, Schapira M, Shuen M, Loppnau P, Mazitschek R, Kwiatkowski NP, Lewis TA, Maglathin RL, et al. : Human HDAC7 harbors a class IIa histone deacetylase-specific zinc binding motif and cryptic deacetylase activity. J Biol Chem 2008, 283:11355–11363. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Somoza JR, Skene RJ, Katz BA, Mol C, Ho JD, Jennings AJ, Luong C, Arvai A, Buggy JJ, Chi E, et al. : Structural snapshots of human HDAC8 provide insights into the class I histone deacetylases. Structure 2004, 12:1325–1334. [DOI] [PubMed] [Google Scholar]
- 37.Vannini A, Volpari C, Filocamo G, Casavola EC, Brunetti M, Renzoni D, Chakravarty P, Paolini C, De Francesco R, Gallinari P, et al. : Crystal structure of a eukaryotic zinc-dependent histone deacetylase, human HDAC8, complexed with a hydroxamic acid inhibitor. Proc Natl Acad Sci USA 2004, 101:15064–15069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Villagra A, Cheng F, Wang H-W, Suarez I, Glozak M, Maurin M, Nguyen D, Wright KL, Atadja PW, Bhalla K, et al. : The histone deacetylase HDAC11 regulates the expression of interleukin 10 and immune tolerance. Nat Immunol 2009, 10:92–100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Huang J, Wang L, Dahiya S, Beier UH, Han R, Samanta A, Bergman J, Sotomayor EM, Seto E, Kozikowski AP, Hancock WW: Histone/protein deacetylase 11 targeting promotes Fox3p+ Treg function. Sci Rep 2017, 7:8626. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Aramsangtienchai P, Spiegelman NA, He B, Miller SP, Dai L, Zhao Y, Lin H: HDAC8 catalyzes the hydrolysis of long chain fatty acyl lysine. ACS Chem Biol 2016, 11:2685–2692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Gantt SL, Gattis SG, Fierke CA: Catalytic activity and inhibition of human histone deacetylase 8 is dependent on the identity of the active site metal ion. Biochemistry 2006, 45:6170–6178. [DOI] [PubMed] [Google Scholar]
- 42.Vannini A, Volpari C, Galinari P, Jones P, Mattu M, Carfí A, De Francesco R, Steinkühler C, Di Marco S: Substrate binding to histone deacetylases as shown by the crystal structure of the HDAC8-substrate complex. EMBO Rep 2007, 8:879–884. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Dowling DP, Gantt SL, Gattis SG, Fierke CA, Christianson DW: Structural studies of human histone deacetylase 8 and its site-specific variants complexed with substrate and inhibitors. Biochemistry 2008. 47:13554–13563. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Lombardi PM, Angell HD, Whittington DA, Flynn EF, Rajashankar KR, Christianson DW: Structure of a prokaryotic polyamine deacetylase reveals evolutionary functional relationships with eukaryotic histone deacetylases. Biochemistry 2011, 50:1808–1817. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Arrar M, Turnham R, Pierce L, de Oliviera CAF, McCammon JA: Structural insight into the separate roles of inositol tetraphosphate and deacetylase-activating domain in activation of histone deacetylase 3. Prot Sci 2013, 22:83–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Decroos C, Christianson NH, Gullet LE, Bowman CM, Christianson KE, Deardorff MA, Christianson DW: Biochemical and structural characterization of HDAC8 mutants associated with Cornelia de Lange Syndrome spectrum disorders. Biochemistry 2015, 54:6501–6513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Porter NJ, Christianson NH, Decroos C, Christianson DW: Structural and functional influence of the glycine-rich loop G302GGGY on the catalytic tyrosine of histone deacetylase 8. Biochemistry 2016, 55:6718–6729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Gantt SL, Joseph CG, Fierke CA: Activation and inhibition of histone deacetylase 8 by monovalent cations. J Biol Chem 2010, 285:6036–6043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Gantt SL, Decroos C, Lee MS, Gullet LE, Bowman CM, Christianson DW, Fierke CA: General base-general acid catalysis in human histone deacetylase 8. Biochemistry 2016, 55:820–832. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Haider S, Joseph CG, Neidle S, Fierke CA, Fuchter MJ: On the function of the internal cavity of histone deacetylase protein 8: R37 is a crucial residue for catalysis. Bioorg Med Chem Lett 2011, 21:2129–2132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- ••51.Porter NJ, Christianson DW: Binding of the microbial cyclic tetrapeptide trapoxin A to the class I histone deacetylase HDAC8. ACS Chem Biol 2017, 12:2281–2286.This paper reports the crystal structure of the HDAC8-Trapoxin A complex and reveals the structural basis for irreversible inhibition. Surprisingly, the epoxide moiety of the α,β-epoxyketone side chain of Trapoxin A does not covalently modify the enzyme; instead, the ketone carbonyl of the α,β-epoxyketone undergoes a hydration reaction to form a gem-diolate, thereby forming a steric latch that traps the inhibitor in the enzyme active site.
- 52.Lahm A, Paolini C, Pallaoro M, Nardi MC, Jones P, Neddermann P, Sambucini S, Bottomley MJ, Lo Surdo P, Carfi A, et al. : Unraveling the hidden catalytic activity of vertebrate class IIa histone deacetylases. Proc Natl Acad Sci USA 2007, 104:17335–17340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Matthews BW: Structural basis of the action of thermolysin and related zinc peptidases. Acc Chem Res 1988, 21:333–340. [Google Scholar]
- 54.Christianson DW, Lipscomb WN: Carboxypeptidase A. Acc Chem Res 1989, 22:62–69. [Google Scholar]
- 55.Olson DE, Udeshi ND, Wolfson NA, Pitcairn CA, Sullivan ED, Jaffe JD, Svinkina T, Natoli T, Lu X, Paulk J, McCarren P, Wagner FF, Barker D, Howe E, Lazzaro F, Gale JP, Zhang YL, Subramanian A, Fierke CA, Carr SA, Holson EB. An unbiased approach to identify endogenous substrates of “histone” deacetylase 8. ACS Chem Biol 2014, 9:2210–2216. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Lopez JE, Haynes SE, Majmudar JD, Martin BR, Fierke CA. HDAC8 substrates identified by genetically encoded active site photocrosslinking. J Am Chem Soc 2017, 139:16222–16227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Schölz C, Weinert BT, Wagner SA, Beli P, Miyake Y, Qi J, Jensen LJ, Streicher W, McCarthy AR, Westwood NJ, Lain S, Cox J, Matthias P, Mann M, Bradner JE, Choudhary C. Acetylation site specificities of lysine deacetylase inhibitors in human cells. Nat Biotechnol 2015, 33:415–423. [DOI] [PubMed] [Google Scholar]
- 58.Wegener D, Wirschung F, Riester D, Schweinhorst A. A fluorogenic histone deacetylase assay well suited for high-throughput activity screening. Chem Biol 2003, 10:61–68. [DOI] [PubMed] [Google Scholar]
- 59.Gurard-Levin ZA, Mrksich M: The activity of HDAC8 depends on local and distal sequences of its peptide substrates. Biochemistry 2008, 47:6242–6250. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Shinsky SA, Christianson DW: Polyamine deacetylase structure and catalysis: prokaryotic acetylpolyamine amidohydrolase and eukaryotic HDAC10. Biochemistry 2018, 57:3105–3114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Blankenship J: Deacetylation of N8-acetylspermidine by subcellular fractions of rat tissue. Arch Biochem Biophys 1978, 189:20–27. [DOI] [PubMed] [Google Scholar]
- 62.Libby PR: Properties of an acetylspermidine deacetylase from rat liver. Arch Biochem Biophys 1978, 188:360–363. [DOI] [PubMed] [Google Scholar]
- •63.Porter NJ, Mahendran A, Breslow R, Christianson DW: Unusual zinc binding mode of HDAC6-selective hydroxamate inhibitors. Proc Natl Acad Sci USA 2017, 114:13459–13464.This paper highlights a distinctive monodentate hydroxamate-Zn2+ binding mode observed for sterically bulky inhibitors that also exhibit selectivity for the inhibition of HDAC6 over other HDAC isozymes.
- 64.Decroos C, Clausen DJ, Haines BE, Wiest O, Williams RM, Christianson DW: Variable active site loop conformations accommodate the binding of macrocyclic Largazole analogues to HDAC8. Biochemistry 2015, 54:2126–2135. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Porter NJ, Wagner FF, Christianson DW: Entropy as a driver of selectivity for inhibitor binding to histone deacetylase 6. Biochemistry 2018, 57:3916–3924. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Shen S, Kozikowski AP: Why hydroxamates may not be the best histone deacetylase inhibitors – what some may have forgotten or would rather forget? ChemMedChem 2016, 11:15–21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Cole KE, Dowling DP, Boone MA, Phillips AJ, Christianson DW: Structural basis of the antiproliferative activity of Largazole, a depsipeptide inhibitor of the histone deacetylases. J Am Chem Soc 2011, 133:12474–12477. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Mackwitz MKW, Hamacher A, Osko JD, Held J, Schöler A, Christianson DW, Kassack MU, Hansen FK: Multicomponent synthesis and binding mode of imidazo[1,2-a]pyridine-capped selective HDAC6 inhibitors. Org Lett 2018, 20:3255–3258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- •69.Porter NJ, Osko JD, Diedrich D, Kurz T, Hooker JM, Hansen FK, Christianson DW: HDAC6-selective phenylhydroxamate inhibitors and the influence of capping groups on hydroxamate-zinc denticity. J Med Chem 2018. 61:8054–8060.This paper reports crystal structures of HDAC6 complexed with selective hydroxamate-based peptoid inhibitors as well as the brain-penetrant PET imaging probe Bavarostat.
- 70.Porter NJ, Christianson DW: Molecular basis for the selective inhibition of histone deacetylase 6 by a mercaptoacetamide inhibitor. ACS Med Chem Lett 2018. 9:1301–1305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Stolfa DA, Marek M, Lancelot J, Hauser A-T, Walter A, Leproult E, Melesina J, Rumpf T, Wurtz J-M, Cavarelli J, Sippl W, Pierce RJ, Romier C, Jung M: Molecular basis for the antiparasitic activity of a mercaptoacetamide derivative that inhibits histone deacetylase 8 (HDAC8) from the human pathogen Schistosoma mansoni. J Mol Biol 2014, 426:3442–3453. [DOI] [PubMed] [Google Scholar]
- •72.Cohen SM: A bioinorganic approach to fragment-based drug discovery targeting metalloenzymes. Acc Chem Res 2017, 50:2007–2016.This is an outstanding review showing how a diverse library of metal-binding pharmacophores can be used to discover new leads for metalloenzyme inhibition.
- •73.Strebl MG, Campbell AJ, Zhao W-N, Schroeder FA, Riley MM, Chindavong PS, Morin TM, Haggarty SJ, Wagner FF, Ritter T, Hooker JM: HDAC6 brain mapping with [18F]Bavarostat enabled by a Ru-mediated deoxyfluorination. ACS Cent Sci 2017, 3:1006–1014.This paper describes the selective, brain-penetrant HDAC6 inhibitor Bavarostat, additionally showing that it can be labeled with 18F for use as a positron emission tomography probe for imaging HDAC6 in the brain.





