You are on page 1of 23

Subscriber access provided by University of Texas Libraries

Letter
Binding of the Microbial Cyclic Tetrapeptide Trapoxin
A to the Class I Histone Deacetylase HDAC8
Nicholas J. Porter, and David W. Christianson
ACS Chem. Biol., Just Accepted Manuscript • DOI: 10.1021/acschembio.7b00330 • Publication Date (Web): 28 Aug 2017
Downloaded from http://pubs.acs.org on August 29, 2017

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 22 ACS Chemical Biology

1
2
3
4
5
6
7
8
9
10 Binding of the Microbial Cyclic Tetrapeptide Trapoxin A to the Class I
11
12 Histone Deacetylase HDAC8
13
14
15
16
17
18
19
20
21 Nicholas J. Porter and David W. Christianson*
22
23
24
25
26
27
28
29
30
31
32
33
Roy and Diana Vagelos Laboratories, Department of Chemistry, University of Pennsylvania,
34
35 Philadelphia, PA 19104-6323, United States
36
37
38
39
40
41
42
43
44
45
46 *author to whom correspondence should be sent: E-mail chris@sas.upenn.edu
47
48
49
50
51
52
53
54
55
56
57
58
59
60 1
ACS Paragon Plus Environment
ACS Chemical Biology Page 2 of 22

1
2
3
4
5
6 Abstract
7
8 Trapoxin A is a microbial cyclic tetrapeptide that is an essentially irreversible inhibitor of
9
10 class I histone deacetylases (HDACs). The inhibitory warhead is the α,β-epoxyketone side-
11
12
chain of (2S,9S)-2-amino-8-oxo-9,10-epoxydecanoic acid (L-Aoe), which mimics the side-chain
13
14
15 of the HDAC substrate acetyl-L-lysine. We now report the crystal structure of the HDAC8–
16
17 trapoxin A complex at 1.24 Å resolution, revealing that the ketone moiety of L-Aoe undergoes
18
19 nucleophilic attack to form a zinc-bound tetrahedral gem-diolate that mimics the tetrahedral
20
21 intermediate and its flanking transition states in catalysis. Mass spectrometry, activity
22
23 measurements, and isothermal titration calorimetry confirm that trapoxin A binds tightly (Kd = 3 ±
24
25
1 nM) and does not covalently modify the enzyme, so the epoxide moiety of L-Aoe remains
26
27
28 intact. Comparison of the HDAC8–trapoxin A complex with the HDAC6-HC toxin complex
29
30 provides new insight regarding the inhibitory potency of L-Aoe-containing natural products
31
32 against class I and class II HDACs.
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 2
ACS Paragon Plus Environment
Page 3 of 22 ACS Chemical Biology

1
2
3 Ever since the discovery of histone acetylation in transcriptional regulation more than 50
4
5
6 years ago,1 thousands of histone and non-histone proteins have been identified as lysine
7
8 acetylation targets in the mammalian acetylome.2-4 The reversible acetylation of L-lysine side
9
10 chains is a critical molecular strategy for the regulation of protein function in vivo and rivals
11
12 phosphorylation in the regulation of diverse biological processes, including the cell cycle and
13
14 central carbon metabolism.5,6 The acetylation of a specific L-lysine residue on a target protein is
15
16 catalyzed by a histone acetyl transferase (HAT) using acetyl CoA as a co-substrate;7,8 the
17
18
19 hydrolysis of acetyl-L-lysine to form free L-lysine and acetate is catalyzed by a histone
20
21 deacetylase (HDAC).9,10
22
23 Upregulated HDAC activity is associated with tumorigenesis, neurodegenerative
24
25 diseases, and immune disorders, so this enzyme family represents an attractive target for
26
27 therapeutic intervention.11,12 Phylogenetic analysis indicates four classes of deacetylases: class
28
29 I HDACs (1, 2, 3, and 8), class IIa (4, 5, 7, and 9) and class IIb (6 and 10) HDACs, class III
30
31
32 HDACs (sirtuins 1-7), and the single class IV enzyme HDAC11.13 Class I, II, and IV HDACs
33
34 require Zn2+ or Fe2+ for optimal catalytic activity.14 The first crystal structure of a metal-
35
36 dependent deacetylase15 revealed an α/β fold identical to that first observed in arginase.16,17
37
38 This fold is distinct from that adopted by sirtuins, which employ a different catalytic mechanism
39
40 that requires the cofactor NAD+.18 Among the metal-dependent HDACs, HDAC8 was the first to
41
42
43
yield a crystal structure.19,20 Important catalytic residues in the HDAC8 active site include
44
45 tandem histidines (electrostatic catalyst H142 and general base-general acid H143)21,22 and a
46
47 conformationally-flexible23 tyrosine residue that assists the Zn2+ ion in the polarization of the
48
49 substrate carbonyl group.24,25
50
51 The tightest-binding HDAC inhibitors are those capable of coordinating to the active site
52
53 Zn2+ ion and also forming hydrogen bonds with catalytically-important active site residues.
54
55
56 Currently, four HDAC inhibitors are approved for clinical use in cancer chemotherapy, one of
57
58 which is the cyclic depsipeptide romidepsin.26,27 Macrocyclic HDAC inhibitors such as
59
60 3
ACS Paragon Plus Environment
ACS Chemical Biology Page 4 of 22

1
2
3 romidepsin or the related depsipeptide largazole28,29 comprise rigid scaffolds to which a Zn2+-
4
5
6 coordinating functional group is attached through a linker that is the approximate length of a
7
8 lysine side chain.
9
10 Trapoxin A, first isolated from the microbial parasite Helicoma ambiens, is a macrocyclic
11
12 tetrapeptide HDAC inhibitor with the amino acid sequence cyclo-[L-Phe–L-Phe–D-hPro–L-Aoe]
13
14 (hPro = homoproline, also known as pipecolic acid; L-Aoe = (2S,9S)-2-amino-8-oxo-9,10-
15
16
epoxydecanoic acid).30 The α,β-epoxyketone moiety of the L-Aoe side chain serves as the Zn2+-
17
18
19 coordinating group and its ketone carbonyl group is isosteric with the scissile carbonyl of the
20
21 HDAC substrate acetyl-L-lysine (Figure 1). Trapoxin A is proposed to act as an irreversible
22
23 inhibitor of class I HDACs31,32 and was used for the first isolation of HDAC1 from the nuclear
24
25 extracts of human Jurkat T cells.33 The epoxide moiety of the L-Aoe side chain was thought to
26
27 react with the nucleophilic side chain of an active site residue; however, a covalent enzyme-
28
29
30
inhibitor complex was not observable by SDS-PAGE/autoradiography.32 Curiously, although
31
32 trapoxin A was found to irreversibly inhibit HDAC1, a class I enzyme, it was found to reversibly
33
34 inhibit the class II enzyme HDAC6.34 The molecular basis of these activity differences has
35
36 remained unclear in the absence of structural data.
37
38 Here, we report the first X-ray crystal structure of trapoxin A complexed with a class I
39
40 histone deacetylase, HDAC8, at 1.24 Å resolution (Figure 2a). Experimental procedures,
41
42
43
including the preparation of a new HDAC8 construct for crystallographic studies, are outlined in
44
45 the Supporting Information (data collection and refinement statistics are recorded in Supporting
46
47 Information Table 1). The structure of the HDAC8–trapoxin A complex reveals that the
48
49 conformation of the tetrapeptide backbone of trapoxin A is identical to that observed in the
50
51 crystal structure of the uncomplexed inhibitor,30 with root-mean-square (rms) deviations of 0.21–
52
53 0.24 Å for 16 main chain atoms of the cyclic peptide backbone; the L-Aoe side chains adopt
54
55
56 different conformations (Supporting Information Figure S1). All peptide linkages of trapoxin A
57
58 adopt the trans configuration except for the Phe-hPro linkage, which forms a cis-peptide. All
59
60 4
ACS Paragon Plus Environment
Page 5 of 22 ACS Chemical Biology

1
2
3 three backbone NH groups donate hydrogen bonds to the side-chain carboxylate group of
4
5
6 D101. The hydrogen bonds between D101 and the backbone NH groups of L-Aoe and the
7
8 adjacent L-Phe residue are similar to those observed between D101 and linear tetrapeptide
9
10 substrates bound to inactivated HDAC8 (Supporting Information Figure S2).24,25 Apart from
11
12 interactions described below for the L-Aoe side chain, no other direct enzyme-inhibitor hydrogen
13
14 bonds are observed. Although there are no intramolecular hydrogen bonds in trapoxin A, the
15
16 side chains of its two L-Phe residues make a favorable quadrupole-quadrupole interaction with
17
18
19 edge-to-face geometry.
20
21 Conformational changes are required in the enzyme active site to accommodate the
22
23 steric bulk of the rigid peptide macrocycle in comparison with substrate binding, primarily in the
24
25 L2 loop (residues G86-I108, of which D87-I94 are disordered). Relative to the structure of
26
27 H143A HDAC8 complexed with a tetrapeptide substrate,25 the greatest change is observed for
28
29
Y100, which undergoes a 116˚ change in side chain torsion angle χ1 (Supporting Information
30
31
32 Figure S2). Similar conformational flexibility of Y100 accommodates the binding of the
33
34 macrocyclic depsipeptide inhibitor largazole.29 In the HDAC8–trapoxin A complex, the
35
36 conformational change of Y100 is triggered by one of the inhibitor L-Phe residues, with which
37
38 one of two observed conformers makes a favorable edge-to-face interaction.
39
40 Surprisingly, the ketone carbonyl of the L-Aoe side chain undergoes nucleophilic attack
41
42
43
by water upon binding to HDAC8, such that the inhibitory species is a gem-diolate (or perhaps a
44
45 gem-diol) stabilized by Zn2+ coordination and three hydrogen bonds. Thus, trapoxin A mimics
46
47 the binding of the tetrahedral intermediate and its flanking transition states in catalysis; the
48
49 origins of high affinity are undoubtedly rooted in the fact that trapoxin A binds as an analogue of
50
51 the postulated transition state. The Zn2+–O1 and Zn2+–O2 distances for the gem-diolate are 2.5
52
53 Å and 1.9 Å, respectively. The O1 hydroxyl group also forms hydrogen bonds with H142 and
54
55
56 H143 (O–O separations = 2.7 Å each), and the O2 oxyanion accepts a hydrogen bond from
57
58 Y306 (O–O separation = 2.6 Å). While it is unusual to see an unactivated ketone binding as a
59
60 5
ACS Paragon Plus Environment
ACS Chemical Biology Page 6 of 22

1
2
3 tetrahedral gem-diolate, which exists to less than 0.2% in aqueous solution (based on the
4
5
6 hydration of the unactivated ketone carbonyl of acetone in aqueous solution35), it is notable that
7
8 the L-Aoe side chain of the cyclic tetrapeptide inhibitor HC toxin (Figure 1) similarly undergoes
9
10 nucleophilic attack in the recently-determined structure of its complex with catalytic domain 2 of
11
12 HDAC6.36 This behavior is also reminiscent of the binding of unactivated aldehyde and ketone
13
14 substrate analogues to carboxypeptidase A, which similarly bind as tetrahedral gem-diolate
15
16 transition state analogues.37,38
17
18
19 In the HDAC8–trapoxin A complex, the epoxide ring of the L-Aoe side chain is clearly
20
21 intact and makes no hydrogen bond interactions with any enzyme residues or water molecules
22
23 (Figure 2b). The closest side chains to the epoxide moiety are those of W141, H142, C153, and
24
25 Y306 with interatomic separations of 3.3–3.7 Å. The epoxide moiety is believed to be required
26
27 for essentially irreversible inhibitory activity against class I HDACs, based on the lack of
28
29 irreversible inhibitory activity for the cyclic tetrapeptide inhibitor apicidin (Figure 1), which lacks
30
31
32 an epoxide moiety.30 Thus, it is curious that the epoxide moiety of trapoxin A does not react with
33
34 the enzyme. However, the crystal structure reveals that although the epoxide moiety binds in
35
36 the vicinity of catalytic general base H143 and highly conserved C153, neither of these potential
37
38 nucleophiles is positioned or oriented for nucleophilic attack at the epoxide (Figure 2b).
39
40
41 We confirmed the irreversibility of trapoxin A inhibition by assaying HDAC8 activity
42
43 following multiple rounds of dialysis of the enzyme-inhibitor complex. Regeneration of activity
44
45 was not observed for HDAC8 preincubated with a 10-fold molar excess of trapoxin A, but was
46
47 observed under the same conditions using apicidin (Figure 3). This result confirms that the
48
49
epoxide moiety is required for essentially irreversible inhibition. However, this result does not
50
51
52 prove that a covalent enzyme-inhibitor complex is formed.
53
54 To study the covalent modification of HDAC8 by trapoxin A in solution, we employed
55
56 liquid chromatography-tandem mass spectrometry (LC-MS/MS) on HDAC8 preincubated for 18
57
58
59
60 6
ACS Paragon Plus Environment
Page 7 of 22 ACS Chemical Biology

1
2
3 hours with 10 molar equivalents of trapoxin A and subsequently digested with trypsin. Prior to
4
5
6 digestion, mass peaks corresponding to HDAC8 covalently modified with one or two trapoxin A
7
8 molecules were observed for this sample by MALDI mass spectrometry (Supporting Information
9
10 Figure S3). Following trypsin digestion, LC-MS/MS analysis, and sequence analysis for residue
11
12 modifications, a mass shift corresponding to the molecular weight of trapoxin A was sporadically
13
14 observed for various cysteine and histidine residues located on the surface of HDAC8
15
16 (Supporting Information Figure S4 and Table S2). Only nonspecific labeling of the enzyme was
17
18
19 observed in the presence of excess inhibitor, with no particular preference for covalent
20
21 modification in the active site. Additionally, incubation of trapoxin A for 1 hour in the presence
22
23 and absence of HDAC8 followed by LC-MS analysis indicated the presence of only intact
24
25 trapoxin A (i.e., with an intact epoxide ring; Supporting Information Figure S5). These results
26
27 strongly suggest that trapoxin A is simply an exceptionally tight-binding, noncovalent transition
28
29 state analogue inhibitor of HDAC8. As noted by Schramm and colleagues,39 transition state
30
31
32 analogues with essentially irreversible binding behavior are feasible for enzymes that exhibit
33
34 catalytic rate enhancements of 1010 or greater, which is likely the case for an amide hydrolase
35
36 such as HDAC8 based on uncatalyzed amide bond hydrolysis half-lives measured in centuries
37
38 by Radzicka and Wolfenden.40
39
40 Isothermal titration calorimetry (ITC) measurements yield HDAC8 dissociation constants
41
42
Kd = 3 ± 1 nM for trapoxin A and Kd = 250 ± 70 nM for apicidin (Supporting Information Figure
43
44
45 S6), indicating 83-fold enhanced binding affinity for trapoxin A relative to apicidin. Given the
46
47 structural similarity between the two cyclic tetrapeptides, and the fact that the macrocycle
48
49 makes very few intermolecular interactions in the HDAC8–trapoxin A complex, it is clear that the
50
51 epoxide moiety of trapoxin A is indeed responsible for tight binding to HDAC8.
52
53 ITC measurements of HDAC8 active site variants indicate that H142 is critical for
54
55
trapoxin A binding just as it is important for catalysis, since the H142A mutation results in
56
57
58 significantly weaker affinity with Kd = 17 ± 5 µM – H142A HDAC8 exhibits a 233-fold reduced
59
60 7
ACS Paragon Plus Environment
ACS Chemical Biology Page 8 of 22

1
2
3 kcat/KM value relative to wild-type HDAC8.22 This residue functions in catalysis with a positively
4
5
6 charged imidazolium group that serves as an electrostatic catalyst.22 Electron density is
7
8 observed for the Nε-H proton of H142 in the 1.24 Å electron density map (Supporting
9
10 Information Figure S7), so the H142 imidazolium group must similarly stabilize the zinc-bound
11
12 gem-diolate.
13
14 Although the L-Aoe side chains of trapoxin A and HC toxin bind to HDAC8 and HDAC6,
15
16 respectively, as the gem-diolate, there are notable differences between the structures of each
17
18
19 enzyme-inhibitor complex that may explain the tighter binding of these inhibitors to class I
20
21 HDACs.34 In the class IIb HDAC6-HC toxin complex (Ki = 350 nM),36 the C–O bond of the
22
23 epoxide adopts an energetically unfavorable eclipsed conformation with respect to the C–O
24
25 bond of the hydroxyl group of the zinc-bound gem-diolate (Figure 4a); in the class I HDAC8–
26
27 trapoxin A complex (Kd = 3 nM), the C–O bond of the epoxide adopts an energetically favorable
28
29 staggered conformation (Figure 4b). The energetically unfavorable eclipsed conformation in the
30
31
32 HDAC6-HC toxin complex appears to be caused by the bulky P571 residue in the L3 loop,
33
34 conserved in all class II HDACs – if the epoxide adopted an energetically favorable staggered
35
36 conformation in the HDAC6 active site, the epoxide methylene group would clash with P571. In
37
38 class I HDACs, P571 is not conserved and the active site is more open. Thus, a more favorable
39
40 binding conformation is accessible to the epoxyketone moiety only in the active site of a class I
41
42
HDAC.
43
44
45 Although the α,β-epoxyketone epoxide moiety of trapoxin A remains intact in the crystal
46
47 structure of its complex with HDAC8, it is notable that this novel functionality is chemically
48
49 reactive in the binding of inhibitors to other enzymes. For example, the proteasome inhibitor
50
51 carfilzomib contains an α,β-epoxyketone that forms a covalent adduct to block proteasome
52
53
54
function.41 The crystal structure of a similar natural product, epoxomicin, complexed with the
55
56 yeast 20S proteasome indicates a multistep cyclization sequence leading to the formation of a
57
58
59
60 8
ACS Paragon Plus Environment
Page 9 of 22 ACS Chemical Biology

1
2
3
morpholino ring between the former α,β-epoxyketone of the inhibitor and the reactive N-terminal
4
5
6 threonine residue of the proteasome subunit.42 A two-step mechanism for inhibitor binding is
7
8 initiated by nucleophilic attack of the threonine hydroxyl group at the epoxyketone carbonyl
9
10 followed by a 6 Exo-Tet ring closure reaction between the α-amino group of threonine and the
11
12 epoxide to generate the morpholino product.43 The chemistry of inhibitor binding in this system
13
14
15 indicates that the carbonyl group appears to be more reactive than the epoxide of the α,β-
16
17 epoxyketone moiety. This is consistent with reactivity trends observed in organic synthesis for
18
19 various α,β-epoxycarbonyl derivatives, where the carbonyl group preferentially undergoes
20
21 nucleophilic addition while leaving the epoxide moiety intact.44-46 With respect to trapoxin A, it is
22
23
notable that an additional peak is observed in mass spectra consistent with gem-diol formation
24
25
26 even in the absence of enzyme (Supporting Information Figure S5).
27
28 Finally, it is interesting to compare the structures of zinc coordination polyhedra in
29
30 different HDAC8-inhibitor complexes (Supporting Information Figure S8). Only one oxygen of
31
32 the gem-diolate of trapoxin A is sufficiently close for inner-sphere coordination, so the overall
33
34 zinc coordination geometry is best described as 4-coordinate distorted tetrahedral. In this
35
36
regard, zinc coordination geometry approaches that observed in the HDAC8–largazole complex,
37
38
39 which exhibits nearly perfect tetrahedral coordination geometry through the binding of the
40
41 largazole thiolate group.29 In contrast with these examples, the hydroxamate group of
42
43 trichostatin A coordinates to zinc in bidentate fashion, so that the overall coordination geometry
44
45 is 5-coordinate square pyramidal.20
46
47 In conclusion, the current work demonstrates that trapoxin A is an essentially irreversible
48
49
50 noncovalent inhibitor of HDAC8: the α,β-epoxyketone side chain of the inhibitor undergoes
51
52 nucleophilic attack by zinc-bound water to bind as a tetrahedral gem-diolate transition state
53
54 analogue. Along with a favorable staggered conformation of the intact epoxide moiety relative to
55
56
57
58
59
60 9
ACS Paragon Plus Environment
ACS Chemical Biology Page 10 of 22

1
2
3 the zinc-bound gem-diolate, these structural features contribute to an exceptionally tight
4
5
6 enzyme-inhibitor complex effectively locked into the enzyme active site.
7
8
9
10 ASSOCIATED CONTENT
11
12 Supporting Information
13
14 The Supporting Information is available free of charge on the ACS Publications website at DOI:
15
16 Detailed methods, figures, and tables
17
18
19 Accession Code
20
21 The atomic coordinates and crystallographic structure factors of the HDAC8-trapoxin A complex
22
23 have been deposited in the Protein Data Bank (www.rcsb.org) with accession code 5VI6.
24
25
26
27 AUTHOR INFORMATION
28
29 Corresponding Author
30
31
32 *Tel.: 215-898-5714. E-mail: chris@sas.upenn.edu
33
34 ORCHID
35
36 David W. Christianson: 0000-0002-0194-5212
37
38 Funding
39
40 We thank the National Institutes of Health for grant GM49758 in support of this research.
41
42 Notes
43
44
45 The authors declare no competing financial interest.
46
47
48
49 ACKNOWLEDGMENT
50
51 N.J.P. thanks the Department of Chemistry at the University of Pennsylvania for the award of a
52
53 graduate research fellowship. Additionally, we thank C. Smith at beamline 9-2 of the Stanford
54
55
Synchrotron Radiation Lightsource (SSRL) for assistance with data collection. The SSRL, SLAC
56
57
58 National Accelerator Laboratory, is supported by the DOE Office of Science, Office of Basic
59
60 10
ACS Paragon Plus Environment
Page 11 of 22 ACS Chemical Biology

1
2
3 Energy Sciences, under Contract No. DE-AC02-76SF00515. The SSRL Structural Molecular
4
5
6 Biology Program is supported by the DOE Office of Biological and Environmental Research, and
7
8 by the NIH, National Institute of General Medical Sciences (including P41GM103393).
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
56
57
58
59
60 11
ACS Paragon Plus Environment
ACS Chemical Biology Page 12 of 22

1
2
3 REFERENCES
4
5
6 1. Allfrey, V. G., Faulkner, R., and Mirsky, A. E. (1964) Acetylation and methylation of histones
7
8 and their possible role in the regulation of RNA synthesis. Proc. Natl. Acad. Sci. U. S. A. 51,
9
10 786–794.
11
12
13 2. Choudhary, C., Kumar, C., Gnad, F., Nielsen, M. L., Rehman, M., Walther, T. C., Olsen, J. V.,
14
15 and Mann, M. (2009) Lysine acetylation targets protein complexes and co-regulates major
16
17
cellular functions. Science 325, 834–840.
18
19
20 3. Zhao, S., Xu, W., Jiang, W., Yu, W., Lin, Y., Zhang, T., Yao, J., Zhou, L., Zeng, Y., Li, H., Li,
21
22 Y., Shi, J., An, W., Hancock, S. M., He, F., Qin, L., Chin, J., Yang, P., Chen, X., Lei, Q., Xiong,
23
24
Y., and Guan, K.-L. (2010) Regulation of cellular metabolism by protein lysine acetylation.
25
26
27 Science 327, 1000–1004.
28
29 4. Wang, Q., Zhang, Y., Yang, C., Xiong, H., Lin, Y., Yao, J., Li, H., Xie, L., Zhao, W., Yao, Y.,
30
31
Ning, Z.-B., Zeng, R., Xiong, Y., Guan, K.-L., Zhao, S., and Zhao, G.-P. (2010) Acetylation of
32
33
34 metabolic enzymes coordinates carbon source utilization and metabolic flux. Science 327,
35
36 1004–1007.
37
38
5. Kouzarides, T. (2000) Acetylation: a regulatory modification to rival phosphorylation? EMBO
39
40
41 J. 19, 1176–1179.
42
43 6. Norvell, A. and McMahon, S. B. (2010) Rise of the rival. Science 327, 964–965.
44
45
46 7. Sterner, D. E. and Berger, S. L. (2000) Acetylation of histones and transcription-related
47
48 factors. Microbiol. Mol. Biol. Rev. 64, 435–459.
49
50
51 8. McCullough, C. E. and Marmorstein, R. (2016) Molecular basis for histone acetyltransferase
52
53 regulation by binding partners, associated domains, and autoacetylation. ACS Chem. Biol. 11,
54
55 632–642.
56
57
58
59
60 12
ACS Paragon Plus Environment
Page 13 of 22 ACS Chemical Biology

1
2
3 9. de Ruijter, A. J., van Gennip, A. H., Caron, H. N., Kemp, S., and van Kuilenburg, A. B. (2003)
4
5
6 Histone deacetylases (HDACs): characterization of the classical HDAC family. Biochem. J. 370,
7
8 737–749.
9
10 10. Lombardi, P. M., Cole, K. E., Dowling, D. P., and Christianson, D. W. (2011) Structure,
11
12
13 mechanism, and inhibition of histone deacetylases and related metalloenzymes. Curr. Opin.
14
15 Struct. Biol. 21, 735–743.
16
17
11. Dokmanovic, M., Clarke, C., and Marks, P. A. (2007) Histone deacetylase inhibitors:
18
19
20 overview and perspectives. Mol. Cancer Res. 5, 981–989.
21
22 12. Falkenberg, K. J. and Johnstone, R. W. (2014) Histone deacetylases and their inhibitors in
23
24
cancer, neurological diseases and immune disorders. Nat. Rev. Drug Discov. 13, 673–691.
25
26
27 13. Gregoretti, I. V., Lee, Y. M., and Goodson, H. V. (2004) Molecular evolution of the histone
28
29 deacetylase family: functional implications of phylogenetic analysis. J. Mol. Biol. 338, 17–31.
30
31
32 14. Gantt, S. L., Gattis, S. G., and Fierke, C. A. (2006) Catalytic activity and inhibition of human
33
34 histone deacetylase 8 is dependent on the identity of the active site metal ion. Biochemistry 45,
35
36 6170–6178.
37
38
39 15. Finnin, M. S., Donigian, J. R., Cohen, A., Richon, V. M., Rifkind, R. A., Marks, P. A.,
40
41 Breslow, R., and Pavletich, N. P. (1999) Structure of a histone deacetylase homologue bound to
42
43 the TSA and SAHA inhibitors. Nature 401, 188–193.
44
45
46 16. Kanyo, Z. F., Scolnick, L. R., Ash, D. E., and Christianson, D. W. (1996) Structure of a
47
48 unique binuclear manganese cluster in arginase. Nature 383, 554–557.
49
50
51 17. Ash, D. E., Cox, J. D., and Christianson, D. W. (2000) Arginase: a binuclear manganese
52
53 metalloenzyme. Metal Ions Biol. Syst. 37, 407–428.
54
55
56
57
58
59
60 13
ACS Paragon Plus Environment
ACS Chemical Biology Page 14 of 22

1
2
3 18. Bheda, P., Jing, H., Wolberger, C., and Lin, H. (2016) The substrate specificity of sirtuins.
4
5
6 Annu. Rev. Biochem. 85, 405–429.
7
8 19. Vannini, A., Volpari, C., Filocamo, G., Casavola, E. C., Brunetti, M., Renzoni, D.,
9
10 Chakravarty, P., Paolini, C., De Francesco, R., Gallinari, P., Steinkühler, C., and Di Marco, S.
11
12
13 (2004) Crystal structure of a eukaryotic zinc-dependent histone deacetylase, human HDAC8,
14
15 complexed with a hydroxamic acid inhibitor. Proc. Natl. Acad. Sci. U. S. A. 101, 15064–15069.
16
17
20. Somoza, J. R., Skene, R. J., Katz, B. A., Mol, C., Ho, J. D., Jennings, A. J., Luong, C., Arvai,
18
19
20 A., Buggy, J. J., Chi, E., Tang, J., Sang, B.-C., Verner, E., Wynands, R., Leahy, E. M., Dougan,
21
22 D. R., Snell, G., Navre, M., Knuth, M. W., Swanson, R. V., McRee, D. E., and Tari, L. W. (2004)
23
24 Structural snapshots of human HDAC8 provide insights into the class I histone deacetylases.
25
26 Structure 12, 1325–1334.
27
28
29 21. Gantt, S. L., Joseph, C. G., and Fierke, C. A. (2010) Activation and inhibition of histone
30
31 deacetylase 8 by monovalent cations. J. Biol. Chem. 285, 6036–6043.
32
33
34 22. Gantt, S. L., Decroos, C., Lee, M. S., Gullett, L. E., Bowman, C. M., Christianson, D. W., and
35
36 Fierke, C. A. (2016) General base-general acid catalysis in human histone deacetylase 8.
37
38 Biochemistry 55, 820–832.
39
40
41 23. Porter, N. J., Christianson, N. H., Decroos, C., and Christianson, D. W. (2016) Structural
42
43 and functional influence of the glycine-rich loop G302GGGY on the catalytic tyrosine of histone
44
45 deacetylase 8. Biochemistry 55, 6718–6729.
46
47
48 24. Vannini, A., Volpari, C., Galinari, P., Jones, P., Mattu, M., Carfí, A., De Francesco, R.,
49
50 Steinkühler, C., and Di Marco, S. (2007) Substrate binding to histone deacetylases as shown by
51
52 the crystal structure of the HDAC8-substrate complex. EMBO Rep. 8, 879–884.
53
54
55
56
57
58
59
60 14
ACS Paragon Plus Environment
Page 15 of 22 ACS Chemical Biology

1
2
3 25. Dowling, D. P., Gantt, S. L., Gattis, S. G., Fierke, C. A., and Christianson, D. W. (2008)
4
5
6 Structural studies of human histone deacetylase 8 and its site-specific variants complexed with
7
8 substrate and inhibitors. Biochemistry 47, 13554–13563.
9
10 26. Furumai, R., Matsuyama, A., Kobashi, N., Lee, K.-H., Nishiyama, M., Nakajima, H., Tanaka,
11
12
13 A., Komatsu, Y., Nishino, N., Yoshida, M., and Horinouchi, S. (2002) FK228 (depsipeptide) as a
14
15 natural prodrug that inhibits class I histone deacetylases. Cancer Res. 62, 4916–4921.
16
17
27. Guan, P. and Fang, H. (2010) Clinical development of histone deacetylase inhibitor
18
19
20 romidepsin. Drug Discovery Ther. 4, 388–391.
21
22 28. Ying, Y., Taori, K., Kim, H., Hong, J., and Luesch, H. (2008) Total synthesis and molecular
23
24
target of largazole, a histone deacetylase inhibitor. J. Am. Chem. Soc. 130, 8455–8459.
25
26
27 29. Cole, K. E., Dowling, D. P., Boone, M. A., Phillips, A. J., and Christianson, D. W. (2011)
28
29 Structural basis of the antiproliferative activity of largazole, a depsipeptide inhibitor of the
30
31
histone deacetylases. J. Am. Chem. Soc. 133, 12474–12477.
32
33
34 30. Itazaki, H., Nagashima, K., Sugita, K., Yoshida, H., Kawamura, Y., Yasuda, Y., Matsumoto,
35
36 K., Ishi, K., Uotani, N., Nakai, H., Terui, A., Yoshimatsu, S., Ikenishi, Y., and Nakagawa, Y.
37
38
(1990) Isolation and structural elucidation of new cyclotetrapeptides, trapoxins A and B, having
39
40
41 detransformation activities as antitumor agents. J. Antibiot. 43, 1524–1532.
42
43 31. Kijima, M., Yoshida, M., Sugita, K., Horinouchi, S., and Beppu, T. (1993) Trapoxin, an
44
45
antitumor cyclic tetrapeptide, is an irreversible inhibitor of mammalian histone deacetylase. J.
46
47
48 Biol. Chem. 268, 22429–22435.
49
50 32. Taunton, J., Collins, J. L., and Schreiber, S. L. (1996) Synthesis of natural and modified
51
52
53
trapoxins, useful reagents for exploring histone deacetylase function. J. Am. Chem. Soc. 118,
54
55 10412–10422.
56
57
58
59
60 15
ACS Paragon Plus Environment
ACS Chemical Biology Page 16 of 22

1
2
3 33. Taunton, J., Hassig, C. A., and Schreiber, S. L. (1996) A mammalian histone deacetylase
4
5
6 related to the yeast transcriptional regulator Rpd3p. Science 272, 408–411.
7
8 34. Fumurai, R., Komatsu, Y., Nishino, N., Khochbin, S., Yoshida, M., and Horinuichi, S. (2001)
9
10 Potent histone deacetylase inhibitors built from trichostatin A and cyclic tetrapeptide antibiotics
11
12
13 including trapoxin. Proc. Nat. Acad. Sci. U. S. A. 98, 87–92.
14
15 35. Lewis, C. A. and Wolfenden, R. (1997) Antiproteolytic aldehydes and ketones: substituent
16
17
and secondary deuterium isotope effects on equilibrium addition of water and other
18
19
20 nucleophiles. Biochemistry 16, 4886–4890.
21
22 36. Hai, Y. and Christianson, D.W. (2016) Histone deacetylase 6 structure and molecular basis
23
24
of catalysis and inhibition. Nat. Chem. Biol. 12, 741–747.
25
26
27 37. Christianson, D. W. and Lipscomb, W. N. (1985) Binding of a possible transition state
28
29 analogue to the active site of carboxypeptidase A. Proc. Natl. Acad. Sci. U. S. A. 82, 6840–
30
31
6844.
32
33
34 38. Christianson, D. W., David, P. R., and Lipscomb, W. N. (1987) Mechanism of
35
36 carboxypeptidase A: hydration of a ketonic substrate analogue. Proc. Natl. Acad. Sci. U. S. A.
37
38
84, 1512–1515.
39
40
41 39. Singh, V., Evans, G. B., Lenz, D. H., Mason, J. M., Clinch, K., Mee, S., Painter, G. F., Tyler,
42
43 P. C., Furneaux, R. H., Lee, J. E., Howell, P. L., and Schramm, V. L. (2005) Femtomolar
44
45
transition state analogue inhibitors of 5'-methylthioadenosine/S-adenosylhomocysteine
46
47
48 nucleosidase from Escherichia coli. J. Biol. Chem. 280, 18265–18273.
49
50 40. Radzicka, A. and Wolfenden, R. (1996) Rates of uncatalyzed peptide bond hydrolysis in
51
52
53
neutral solution and the transition state affinities of proteases. J. Am. Chem. Soc. 118, 6105–
54
55 6109.
56
57 41. Spaltenstein, A., Leban, J. J., Huang, J. J., Reinhardt, K. R., Viveros, O. H., Sigafoos, J.,
58
59
60 16
ACS Paragon Plus Environment
Page 17 of 22 ACS Chemical Biology

1
2
3
and Crouch, R. (1996) Design and synthesis of novel protease inhibitors. Tripeptide α',β'-
4
5
6 epoxyketones as nanomolar inactivators of the proteasome. Tet. Lett. 37, 1343–1346.
7
8 42. Johnson, H. W. B., Anderl, J. L., Bradley, E. K., Bui, J., Jones, J., Arastu-Kapur, S., Kelly, L.
9
10
11
M., Lowe, E., Moebius, D. C., Muchamuel, T., Kirk, C., Wang, Z., and McMinn, D. (2017)
12
13 Discovery of highly selective inhibitors of the immunoproteasome low molecular mass
14
15 polypeptide 2 (LMP2) subunit. ACS Med. Chem. Lett. 8, 413–417.
16
17
18 43. Groll, M., Kim, K. B., Kairies, N., Huber, R., and Crews, C. M. (2000) Crystal structure of
19
20 epoxomicin:20S proteasome reveals a molecular basis for selectivity of α',β'-epoxyketone
21
22 proteasome inhibitors. J. Am. Chem. Soc. 122, 1237–1238.
23
24
25 44. Urabe, H., Matsuka, T., and Sato, F. (1992) (2S,3S)-2,3-epoxy-3-trimethylsilylpropanal as a
26
27 new conjunctive reagent. Tet. Lett. 33, 4179–4182.
28
29
30 45. Howe, G. P., Wang, S., and Procter, G. (1987) Stereoselective additions to α,β-epoxy-
31
32 aldehydes; the formation of "non-chelation controlled" products. Tet. Lett. 28, 2629–2632.
33
34
35 46. Pégorier, L., Petit, Y., and Larchevêque, M. (1994) A synthetic route to anti aminoalkyl
36
37 epoxides by stereocontrolled reductive amination of ketoepoxides. J. Chem. Soc., Chem.
38
39 Commun. 5, 633–634.
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 17
ACS Paragon Plus Environment
ACS Chemical Biology Page 18 of 22

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 Figure 1. The canonical HDAC substrate peptidyl acetyl-L-lysine and structurally-related cyclic
43
44 tetrapeptide inhibitors. The ketone carbonyl group of each inhibitor is isosteric with the scissile
45
46 carbonyl of acetyl-L-lysine.
47
48
49
50
51
52
53
54
55
56
57
58
59
60 18
ACS Paragon Plus Environment
Page 19 of 22 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
35
36
37 Figure 2. 1.24 Å-resolution structure of the HDAC8–trapoxin A complex. (a) Stereoview of
38
39 trapoxin A (C = orange, N = blue, O = red) bound in the active site of HDAC8 (C = light blue, N =
40
41 blue, O = red). The simulated annealing omit map of trapoxin A (green) is contoured at 3.0σ.
42
43 Metal coordination and hydrogen bond interactions are indicated by solid and dashed black
44
45
46
lines, respectively. (b) Close-up stereoview showing the zinc-bound gem-diolate. Atomic color
47
48 coding is identical to that in (a).
49
50
51
52
53
54
55
56
57
58
59
60 19
ACS Paragon Plus Environment
ACS Chemical Biology Page 20 of 22

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
Figure 3. HDAC activity assays. Recovered HDAC8 activity from samples incubated with
39
40 DMSO, apicidin, and trapoxin A (TpxA) after no (gray), one (orange) or two (blue) rounds of
41
42 dialysis against 104-fold excess fresh buffer.
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 20
ACS Paragon Plus Environment
Page 21 of 22 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 Figure 4. Epoxyketone binding modes. (a) HC toxin (C = yellow, N = blue, O = red) bound to
32
33 HDAC6 (C = gray, N = blue, O = red). (b) Trapoxin A (C = orange, N = blue, O = red) bound to
34
35 HDAC8 (C = light blue, N = blue, O = red). Residues affecting epoxide conformations are
36
37
labeled. Zinc ions are shown as gray spheres.
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60 21
ACS Paragon Plus Environment
ACS Chemical Biology Page 22 of 22

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 78x40mm (300 x 300 DPI)
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 Paragon Plus Environment

You might also like