Subscriber access provided by UNIV OF NEWCASTLE
Article
Crystal structure of a histone deacetylase homolog from Pseudomonas aeruginosa Andreas Kraemer, Thomas Wagner, Özkan Yildiz, and Franz-Josef Meyer-Almes Biochemistry, Just Accepted Manuscript • DOI: 10.1021/acs.biochem.6b00613 • Publication Date (Web): 14 Nov 2016 Downloaded from http://pubs.acs.org on November 21, 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.
Biochemistry 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 37
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 56 57 58 59 60
Biochemistry
1
Crystal structure of a histone deacetylase homolog
2
from Pseudomonas aeruginosa
3
Andreas Krämerǂ, Thomas Wagnerǂ, Özkan Yildiz£, Franz-Josef Meyer-Almesǂ*
4 5
ǂ
University of Applied Science, Department of Chemical Engineering and Biotechnology, 64295
6 7 8 9
Darmstadt, Germany £
Max Planck Institute of Biophysics, Department of Structural Biology, 60438 Frankfurt am Main, Germany
10 11 12 13 14 15 16
ACS Paragon Plus Environment
1
Biochemistry
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 56 57 58 59 60
Page 2 of 37
17
KEYWORDS
18
Histone deacetylase, HDAC, Pseudomonas aeruginosa, Acetate, Amidohydrolases, X-Ray
19
Crystallography, Histone Deacetylase Inhibitors, Lysine Deacetylation, Histone deacetylase like
20
amidohydrolase,
21
amdidohydrolase, APAH
HDAH,
Histone
deacetylase
like
protein,
HDLP,
acetylpolyamine-
22 23
ABBREVIATIONS
24
HDAC, histone deacetylase; HDAH, histone deacetylase-like amidohydrolase; HDLP, histone
25
deacetylase-like protein; KDAC, lysine deacetylase; APAH, acetylpolyamine amidohydrolase;
26
SAHA,
27
nonanamide; TFA, trifluoroacetate; Ac, acetate; AMC, 7-Amino-4-methylcoumarin; boc, tert-
28
butyloxycarbonyl; HEPES, 2-[4-(2-hydroxyethyl)piperazin-1-yl]ethanesulfonic acid; TRIS, 2-
29
Amino-2-hydroxymethyl-propane-1,3-diol; RMSD., root mean square deviation; PFSAHA,
30
2,2,3,3,4,4,5,5,6,6,7,7‐dodecafluoro‐N‐hydroxy‐N'‐phenyloctanediamide
N-Hydroxy-N′-phenyloctandiamid;
SATFMK,
9,9,9-trifluoro-8-oxo-N-phenyl-
31 32 33 34 35 36 37 38
ACS Paragon Plus Environment
2
Page 3 of 37
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 56 57 58 59 60
Biochemistry
39 40
ABSTRACT
41
Despite the recently growing interest in the acetylation of lysine residues by prokaryotic
42
enzymes, the underlying biological function is still not well understood. Deacetylation is
43
accomplished by proteins which belong to the Histone deacetylase (HDAC) superfamily. In this
44
report we present the first crystal structure of PA3774, a histone deacetylase homolog from the
45
human pathogen Pseudomonas aeruginosa with high homology to class IIb HDACs. We solved
46
the crystal structure of the ligand-free enzyme and protein-ligand complexes with a
47
trifluoromethylketone inhibitor and the reaction product acetate. Moreover, we produced loss of
48
function mutants and determined the structure of the inhibitor free PA3774H143A mutant, the
49
inhibitor-free PA3774Y313F mutant and the PA3774Y313F mutant in complex with the highly
50
selective hydroxamate inhibitor PFSAHA. The overall structure reveals that the exceptionally
51
long L1-loop mediates the formation of a tetramer composed of two “head-to-head” dimers. The
52
distinctive dimer interface significantly confines the entrance area of the active site suggesting a
53
crucial role for substrate recognition and selectivity.
54 55 56 57 58 59 60
ACS Paragon Plus Environment
3
Biochemistry
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 56 57 58 59 60
61
Page 4 of 37
INTRODUCTION
62
Histone deacetylases (HDAC), acetylpolyamine-amidohydrolases (APAH) and acetoin
63
utilization Proteins (AcuC) belong to an ancient protein superfamily known as the histone
64
deacetylase superfamily 1. In the past years especially histone deacetylases raised much attention
65
due to their important roles in the cell cycle and differentiation. For instance, deregulation of
66
these enzymes can contribute to the development of cancer
67
novel therapeutic target for chemo therapy which led to the approval of new drugs for the
68
treatment of cutaneous T-cell lymphoma such as Vorinostat and Romidepsin 4, 5. The HDACs are
69
classified in four groups based on their sequence and domain organization. Class Ι, IIa, IIb and
70
IV are known as zinc dependent “classical” histone deacetylases, whereas class III HDACs or
71
sirtuins constitute an atypical NAD+ dependent class 6. According to sequence analysis class Ι,
72
IIa, IIb, IV share very similar catalytic domain organization and differ mainly in the length of the
73
N- and C-termini. Class Ι HDACs consist of 370-500 residues, class II enzymes have 660-1220
74
residues and class IV which only consist of HDAC11 that shows characteristics of both class Ι
75
and class II 7. Interestingly, HDACs have also been shown to be involved in the deacetylation of
76
acetylated lysine residues in non-histone proteins 8. For this reason they are sometimes called
77
lysine deacetylases (KDACs) with respect to their basic function 9. It has been further suggested
78
that APAHs and AcuCs are the precursor enzymes of HDACs since prokaryotic cells have no
79
histones. Yet, few prokaryotic enzymes with high homology to zinc dependent HDACs have
80
been studied before. Histone deacetylase-like protein (HDLP) from Aquifex aeolicus
81
the highest homology to class Ι HDACs, whereas histone deacetylase-like amidohydrolase
82
(HDAH) from Bordetella/Alcaligenes was assigned to HDAC class IIb
83
prokaryotic studies have shown that APAHs catalyze the deacetylation of acetylated polyamines
2, 3
. In this context, they became a
11
10
exhibits
. In addition,
ACS Paragon Plus Environment
4
Page 5 of 37
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 56 57 58 59 60
Biochemistry
84
12, 13
, whereas the reactions catalyzed by AcuC proteins are not known 1. Moreover, to this date
85
there are several structures available for human members of the HDAC family including
86
members of class I: HDAC1
87
and HDAC7 21 and recently class IIb: HDAC6 22, 23.
14
, HDAC2
15, 16
, HDAC3
17
, HDAC8
18, 19
; class IIa: HDAC4
20
88
P. aeruginosa is a gram-negative bacterium with one of the largest sequenced bacterial
89
genome until now consisting of 6.3 million base pairs. It shows a vast variety of metabolic
90
pathways and is able to adapt to a wide range of different environments
91
aeruginosa has multiple defense strategies against many structurally diverse antibiotics and is
92
indeed one of the most relevant human pathogens associated with a wide spectrum of infections,
93
especially in immunosuppressed patients. It is particularly associated with cystic fibrosis,
94
infection of the urinary-tract and bacteremia in burn victims 25.
24
. Moreover, P.
95
In the genome of P. aeruginosa three enzymes with sequence similarity to HDACs were
96
identified: PA0321, PA1409 and PA3774 as referenced in the Pseudomonas genome database 26.
97
Although all three were predicted to be acetylpolyamine-amidohydrolases (APAHs)
98
work demonstrated that only PA0321 and PA1409 are capable of deacetylating acetylated
99
polyamines, whereas PA3774 shows no significant activity 27. On the contrary, PA3774 exhibits turnover
rates
artificial
HDAC
substrates
containing
acetylated
recent
100
high
101
trifluoracetylated lysine residues, thus proposing a role as a protein deacetylase with an hitherto
102
unknown substrate
103
strain PA01. We present a total of six crystal structures, including the ligand-free enzyme, the
104
enzyme in complex with acetate, a complex with a trifluoromethylketone inhibitor, and
105
additional three loss-of-function mutant structures: PA3774H143A and PA3774Y313F without
106
ligands and the complex between PA3774Y313F and the hydroxamate inhibitor PFSAHA. The
27
against
26
and
. We here present the crystal structure of PA3774 from the P. aeruginosa
ACS Paragon Plus Environment
5
Biochemistry
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 56 57 58 59 60
Page 6 of 37
107
observed quaternary structure suggests that tetramer assembly is essential for substrate
108
recognition and selectivity. The structural information on the active site extends our
109
understanding of class II HDACs and provides a solid base for the rational design of specific and
110
potent inhibitors for this class. Furthermore, our mutational studies demonstrate the important
111
role of catalytic core residues and invoke the hypothesis of different deacetylation mechanisms
112
in class Ι and class IIb HDACs.
113 114
MATERIAL AND METHODS
115
Reagents. Standard chemicals and solvents for activity test and crystallization were purchased
116
at the highest available purity level from Sigma-Aldrich (USA), Merck (Germany), Roth
117
(Germany). Boc-Lys(Ac)-AMC and Boc-Lys(TFA)-AMC from Bachem (Switzerland),
118
Paratone-N from Hampton.
119
Overexpression, purification and mutagenesis. The enzyme PA3774 was overexpressed and
120
purified as published elsewhere 28 (For detailed chromatograms and SDS PAGEs see supporting
121
online material Fig. S1). The PA3774H143A, PA3774H144A, PA3774Y313H and PA3774Y313F
122
mutants were prepared using the QuickChange site-directed mutagenesis Kit (Stratagene), the
123
pET21a(+)/PA3774-CPD plasmid as template and the oligonucleotide primers listed in the
124
supplementary data (Fig. S2).
125
Following PCR mutagenesis and DpnI digestion, the resulting plasmids were used to transform
126
E. Coli XL1-Blue competent cells. After transformation the cells were plated on LB-agar plates
127
containing 100 µg/mL ampicillin. Plasmid DNA isolated from single colonies of transformants
128
was sequenced to verify the mutations by the sequencing service of LMU Munich. Plasmids
129
containing the desired mutations were then used to transform E. coli BL21(DE3) cells.
ACS Paragon Plus Environment
6
Page 7 of 37
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 56 57 58 59 60
Biochemistry
130
Overexpression and purification of the mutated enzymes was identical as described for the native
131
enzyme.
132
Crystallization and data collection. Crystals of native and mutated PA3774 were grown by
133
hanging drop vapor diffusion against 500 µl reservoir (Fig. S3). The reservoir solution contained
134
0.5 M K2HPO4, 0.5 M Na2HPO4, 0.1 M (NH4)2SO4 at a final pH of 7.5. The reservoir solution
135
and protein (6 mg/ml dissolved in 20 mM HEPES pH 8, 50 mM NaCl, 2 mM TCEP) were mixed
136
in a 1:1 ratio to a final volume of 4 µl. Crystals grew in two to four days at 20°C. The obtained
137
crystals were soaked with the inhibitor added in 10 fold molar excess directly to the drop and
138
incubated for at least 48 h. To obtain the PA3774-acetate complex the crystals were soaked with
139
50 mM ammonium-acetate. The crystals were transferred into Paratone-N and flash-frozen in
140
liquid nitrogen. Diffraction data were collected on beamline PXII at the Swiss Light Source
141
(SLS, Villigen, Switzerland).
142
Structure determination and refinement. Diffraction data were indexed and integrated using 29
and scaled with AIMLESS 30. The structure was solved by molecular replacement
143
iMOSFLM
144
with MOLREP 31 using one monomer of the HDAH structure from Bordetella/Alcaligenes strain
145
FB188 (pdb-id 1ZZ0) as a model. Model building was carried out in COOT 32 and REFMAC5 33
146
was used for refinement. All used programs belong to the CCP4 suite 34. Water molecules were
147
added automatically with COOT and afterwards checked manually for reasonably hydrogen
148
bonding. All Figures were created with PyMOL.
149 150
Enzyme activity assay. Enzyme activity was determined following the standard fluorogenic 35
151
enzyme activity assay in
. Briefly, the cleavage of trifluoroacetate or acetate by the enzyme
152
allows trypsin to release the fluorophore from the substrate, thereby resulting in a shift in
ACS Paragon Plus Environment
7
Biochemistry
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 56 57 58 59 60
Page 8 of 37
153
fluorescence (Ex.: 340 nm Em.: 460 nM). Tests were run at room temperature with 100 nM
154
enzyme and 0.5 mg/mL trypsin solved in assay buffer (20 mM Tris-HCl pH 8, 50 mM NaCl,
155
0.001 % Pluronic). Boc-Lys(Ac)-AMC and Boc-Lys(TFA)-AMC at a concentration of 20 µM
156
were used as fluorogenic substrates. Because trypsin does not digest the enzyme during the test
157
procedure the activity assay can be run in a continuous manner and the linear slope of the
158
fluorescence signal is used to determine the enzyme activity. The experiment was carried out in a
159
PheraStar Fluorescence Spectrometer (BMG Labtech) and data were analyzed with the program
160
Graph Pad Prism.
161 162
RESULTS
163
Overall structure. PA3774 and all variants crystallized in the tetragonal space group P41212
164
with two molecules in the asymmetric unit (Table 1) forming a “head-to-head” homo dimer with
165
a protein-protein interface of about 2090 Å2. The electron density could be interpreted for
166
residues 2-374 of one monomer and 2-379 of the other out of a total number of 384 possible
167
residues. The enzyme core adopts an open α/β fold which is characteristic for this class of
168
enzymes
169
and two smaller antiparallel β-turns (Fig. 1).
18, 36, 37
. It consists of eight central stranded parallel β-sheets surrounded by 14 helices
170
Oligomerization. The unit cell of the tetragonal crystals consists of 8 “head-to-head” dimers
171
described for the asymmetric unit. In here, there are two “head-to-head” dimers in very close
172
distance to each other forming a tetramer. Multi angle laser light scattering (MALLS)
173
experiments following analytical size exclusion chromatography suggest that the tetramer is the
174
relevant quaternary structure in solution (see Fig. S4 and S5). The tetrameric structure is also in
175
line with PISA analysis
38
(for detailed analysis see Fig. S6). However, since the interface area
ACS Paragon Plus Environment
8
Page 9 of 37
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 56 57 58 59 60
Biochemistry
176
between the two “head-to-head” dimers is with 2090 Å2 much larger compared to each of the
177
other interfaces between subunits which are about 685 and 407 Å2, the structure is better
178
described as a dimer of dimers. The tetramer has therefore the dihedral point group D2 (also
179
noted as 222). A closer look at the crystal structure reveals that the elongated L1-Loop inside the
180
residues 15-42 plays a pivotal role in the stabilization of the highly symmetric PA3774 tetramer.
181
Every L1-Loop in each subunit points to the symmetry center of the tetramer where all subunits
182
are in contact with each other (Fig. 1D and Fig. S6). A similar assembly can be found in the
183
crystal structures of the related enzyme HDAH from Bordetella/Alcaligenes (pdb-id 1ZZ0) 39, 40
184
(Fig. 2). In this enzyme a similar elongated L1-loop is found. The importance of this long loop
185
region for tetramerization becomes evident by the monomeric structures of human HDAC8 18, 41
186
which has only 4 residues in the corresponding loop (Fig. 2 and 3). Moreover, the oligomeric
187
state and especially the head-to-head dimer interface have a direct influence on the active site. A
188
large area of the entrance cavity in each monomer gets partially blocked by another monomer
189
resulting in a substantially narrowed entrance surface (Fig. 4) suggesting a crucial role for
190
substrate recognition and selectivity.
191
Catalytic center. The catalytic center shows the typical motif observed for many enzymes of
192
the histone deacetylase superfamily. The catalytically zinc ion is trigonal bipyramidal-
193
coordinated by the side chains of D181, H183, D269 and a corresponding ligand. Close to the
194
zinc ion there are three conserved catalytic residues H143, H144 and Y313 which make
195
hydrogen bonds with bound ligands. Single mutations of these residues abolish completely the
196
activity (Table 2). We further localized two octahedrally coordinated potassium ions in PA3774.
197
These monovalent cation sites are also found in other HDACs and HDAC-like proteins
198
potassium ion is coordinated by the main-chain carbonyl of D179, D181, H183, L203 and by the
42
. One
ACS Paragon Plus Environment
9
Biochemistry
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 56 57 58 59 60
Page 10 of 37
199
hydroxyl and carboxyl side chain of S203 and D179, respectively. This potassium ion seems to
200
keep the zinc-coordinating residues in position. Especially the hydrogen bridges between the
201
amino acid pairs D179-H143 and H144-N183 are important for the mechanism (Fig. 5A) and
202
found in HDAH, APAH and in most class II HDACs
203
asparagine pair is replaced by histidine-aspartate40. The other potassium ion is coordinated by the
204
main chain carbonyl of V198, Y192, R195, F227 and two water molecules. In contrast to the
205
potassium ions the occupancy for the zinc ion is lower (60-70%), probably due to the absence of
206
zinc ions the crystallization buffer.
207
PA3774-acetate complex. The complex with the reaction product acetate was achieved by
208
soaking the crystals with ammonium acetate in reservoir solution (final concentration 50 mM).
209
The 1.70 Å structure allows a precise interpretation of nearly all side chains and backbone
210
interactions. In this complex the acetate anion coordinates the zinc ion with both oxygen atoms
211
(Fig. 5A), resulting in a trigonal bipyramidal coordination of the zinc ion. The observed binding
212
is almost identical to the acetate bound HDAH structure from Bordetella/Alcaligenes (pdb-id
213
1ZZ0) 40.
214
PA3774–SATFMK complex. The trifluoromethylketone inhibitor SATFMK was soaked in a
215
tenfold molar excess to the protein and the structure was solved at 1.71 Å. The inhibitor has a
216
remarkably high affinity towards the enzyme which is represented by an IC50 value of 9.7 nM 28
217
(Fig. S13). The observed electron density indicates that the ketone inhibitor binds in its gem-diol
218
form in a bidentate fashion (Fig. 5B and Fig. S7). The overall binding geometry of the zinc ion is
219
described best by a trigonal bipyramid. The two oxygen zinc distances are almost identical, with
220
2.08 Å and 2.06 Å respectively. The binding of gem-diol mimics the tetrahedral transition state
221
analogue during the deacetylation. One oxygen of the gem-diol accepts hydrogen bonds from
13, 20, 40
. In class Ι HDACs the histidine-
ACS Paragon Plus Environment
10
Page 11 of 37
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 56 57 58 59 60
Biochemistry
222
H144 and H143, whereas the other one from Y313. The high affinity of trifluoromethylketones
223
can be ascribed further to favorable interaction of this group with the buried binding pocket. The
224
fluorine atoms form orthogonal multipolar interactions with the side chain hydroxyl group of
225
Y313, the backbone amide group of G311, the thiol group of C154 as well as van-der-Waals
226
contacts with P141. The binding pocket around the catalytic center is highly conserved in class
227
IIa HDACs and class IIb HDACs. A similar binding was observed for the same inhibitor bound
228
to HDAH (pdb-id 2GH6) 39 and another trifluoromethylketone inhibitor bound to HDAC4 (pdb-
229
id 2VGJ) 20. In general, derivatives of trifluoromethylketones are likely promising candidates for
230
developing highly potent and selective inhibitors for this class of enzymes 43.
231
Ligand-free PA3774H143A. Based on the observed structure, several PA3774 mutants were
232
created (Table 2). The structure of the PA3774H143A mutant was solved at 1.99 Å and is
233
essentially identical with the wild type enzyme. It is worth noticing that the backbone structure
234
of the enzyme remains unchanged by the mutation and therefore the inactivation of the enzyme
235
solely results from the loss of the catalytic effect of the histidine sidechain (Fig. S9).
236
Ligand-free PA3774Y313F and complex with PFSAHA. The structure of ligand-free
237
PA3774Y313F was determined at a resolution of 2.05 Å. The structure was also solved with the
238
inhibitor PFSAHA at a resolution of 2.48 Å (Fig. 6 and S8). This inhibitor is a potent
239
perfluorinated derivative of the commercial drug SAHA (Vorinostat)
240
towards PA3774 with an IC50 value of 62 nM, whereas the affinity against human HDACs is far
241
less pronounced with IC50 values in the µM range (HDAC6: 11 µM, HDAC1: 29 µM,
242
HDAC8: 17 µM, and HDAC7: 11 µM)
243
PFSAHA, the different selectivity arises exclusively from the bulkier and polar nature of the
244
fluorine atoms. In the structure of PA3774Y313F-PFSAHA complex the hydroxamate group
44
28
and is highly selective
. Given the high similarity between SAHA and
ACS Paragon Plus Environment
11
Biochemistry
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 56 57 58 59 60
Page 12 of 37
40
245
chelates the zinc ion in the typical manner as observed for HDAH (pdb-id: 1ZZ1)
246
(pdb-id: 1C3S)
247
hydroxylamine oxygen makes a hydrogen bond with H143. The catalytically active Y313 usually
248
donates a hydrogen bond to the carbonyl oxygen of the zinc chelating group adopting an
249
“inward” orientation. Yet, in the PA3774 mutant structure (Fig. 6), this Y313 is replaced by
250
phenylalanine which takes an alternative “outward” conformation pointing away from the active
251
site. Notably, in the same mutant without inhibitor, both conformations – “inward” and
252
“outward” – were found indicating the exceptional malleability of the phenyl ring at this
253
location. The reason why the “outward” conformation was exclusively observed in the inhibitor
254
bound structure might be due to the bulkier nature of the fluorine atoms and a likely repulsion of
255
the aromatic ring of F313 by partially negatively charged fluorine atoms. Moreover, the
256
“outward” conformation is in agreement with the structures of HDAC4
257
both belong to class IIa HDACs
258
from the active site. Furthermore, the “outward” conformation was found in native APAH from
259
M. ramosa (pdb-id 3Q9F)
260
warhead and the perfluorinated linker of PFSAHA is reduced (Fig. S8). However, the X-ray data
261
unequivocally prove the presence of the inhibitor within the binding pocket. The additional
262
electron density inside the binding pocket of the protein-ligand complex is not present in the free
263
mutant structure. The refined complex structure reveals a good agreement with the electron
264
density in respect to the resolution of 2.48 Å. The observed reduced electron is probably caused
265
by an incomplete occupancy of only about 70% bound PFSAHA as well as electron-withdrawing
266
effects of the adjacent fluorine atoms and the hydroxamate group on the other side.
36
or HDLP
. The hydroxylamine nitrogen forms a hydrogen bond with His144, while the
13
21
20
and HDAC7 which
. There, the corresponding histidine residue also points away
. The electron density between the Zn2+ chelating hydroxamate
ACS Paragon Plus Environment
12
Page 13 of 37
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 56 57 58 59 60
267
Biochemistry
Comparison with related structures.
268
The most related protein to PA3774 is HDAH from Bordetella/Alcaligenes with a sequence
269
identity of 49% (Fig. 2 and 3). The crystal structure of HDAH has been solved before
270
shows not only a structurally highly conserved core, as expected, but also a highly similar overall
271
structure with a RMSD of 0.75 Å over 269 Cα atoms. Although the overall structure of HDAH
272
has been reported as a monomer
273
similar quaternary structure as PA3774. Distinct structural differences between the two closely
274
related proteins appear in the entrance of the active site. In HDAH a “closed state” in complex
275
with acetate (pdb-id 1ZZ0) and “open state” (pdb-id 1ZZ1) in complex with SAHA was
276
observed
277
pocket is accessible regardless whether a big inhibitor or a small molecule like acetate is bound
278
to the active site. Another striking difference is the fact that PA3774 is much more active on the
279
artificial Boc-Lys(TFA)-AMC substrate compared to Boc-Lys(Ac)-AMC 27, whereas the activity
280
of HDAH on both substrates in basically identical 46.
281
PA3774 exhibits 27% sequence identity and a RMSD of 1.24 Å over 201 Cα atoms with the
282
APAH from M. ramosa, another enzyme of the HDAC superfamily. The major difference
283
between PA3774 and the APAH from M. ramosa is a loop insert (residues 76-113 in APAH)
284
which is 31 amino acids longer than the corresponding loop region in PA3774. This insert is
285
responsible for a completely different quaternary structure in APAH compared to PA3774 (Fig.
286
2C). Furthermore, this loop is not found in other HDACs and seems to be a unique feature of
287
acetylpolyamine deacetylases. This is particularly interesting because PA3774 has originally
288
been annotated as a putative APAH but shows no activity against any tested acetylpolyamine 27.
289
On the contrary, P. aeruginosa contains two further putative acetylpolamine deacetylases:
40
11, 45
11
and
, a closer look at the crystal structure of HDAH reveals a
. In contrast, such a “closed state” was not observed for PA3774 and the binding
ACS Paragon Plus Environment
13
Biochemistry
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 56 57 58 59 60
26
Page 14 of 37
290
PA0321 and PA1409
291
There is no structure available for PA1409 or PA0321 but sequence analyses indicate that both
292
enzymes share a similar loop insert like the APAH from M. ramosa (indicated by the red box in
293
Fig. 3). The dimer assembly of APAH might be very different in the quaternary structure but
294
both, APAH and PA3774, have in common that the respective adjacent subunit limits the access
295
to the active site, albeit in a different manner. Presumably, this is one of the key reasons for
296
known differences in substrate selectivity and recognition. Further sequence features unique to
297
APAHs are highlighted by the red boxes in Fig. 3. For instance, two remarkable differences are
298
located inside the binding pocket. Throughout all HDACs there is a highly conserved
299
phenylalanine (F153 in PA3774 nomenclature) but in the APAH from M. ramosa, as well as in
300
both APAHs (PA1409 and PA0321) from P. aeruginosa, this residue is replaced by a tyrosine.
301
This tyrosine forms a hydrogen bond with a glutamate (E17-Y168 in APAH) in the structure
302
from M. ramosa
303
varies in all other enzymes of this family. The second difference is L276 which is replaced by an
304
isoleucine in APAHs. The exact function of these distinct differences between APAHs and all
305
other enzymes of this family is not clear, since they do not undergo direct interactions with any
306
substrate in the APAH structure. However, given their high conservation inside the active site,
307
these distinctions seem to be unique features to amidohydrolases and may thus be useful for the
308
prediction and computational annotation of yet unknown APAHs.
309
The closest sequence identity (34%) among human HDACs has HDAC10 and the second
310
domain of HDAC6 which both belong to the class IIb HDACs. Interestingly, both class IIb
311
HDACs have two deacetylase domains. HDAC6 has two functional deacetylase domains,
312
whereas HDAC10 contains only one functional domain and a second, putatively non-functional
13
. These enzymes are indeed capable of metabolizing acetylpolyamines.
. The glutamate residue is also conserved in APAHs from P. aeruginosa but
ACS Paragon Plus Environment
14
Page 15 of 37
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 56 57 58 59 60
Biochemistry
47
313
domain. Furthermore, both are found in the cytoplasm
. Very recently the structure of both
314
catalytical domains 1 and 2 of HDAC6 has been published, by two different groups
315
expected both structures show similar domain organizations compared to PA3774.
316
Despite the fact that PA3774 exhibits only around 20% sequence identity with class Ι HDACs,
317
their α/β fold domain as well as the monovalent cation sites are highly conserved. This is
318
represented in the small RMSD of their backbone superimposition. Differences mainly appear in
319
loop regions. With class IIa HDACs the sequence identity is about 26% and the RMSD values
320
are even lower compared to class Ι. HDACs. Detailed structure comparisons and RMSD values
321
of all HDAC and PA3774 can be found in the supplement Fig. S11.
22, 23
. As
322 323
Mutational studies. Based on the observed crystal structures we generated several mutants. The
324
amino acids H143 and H144 which form hydrogen bonds with ligands were replaced by
325
nonreactive alanines. The relative activities regarding the common substrates Boc-Lys(Ac)-AMC
326
and Boc-Lys(TFA)-AMC are listed in table 2. Neither of the single histidine mutants showed
327
detectable activity with the acetylated substrate and a weak activity (about 1%) with the
328
trifluoroacetylated
329
trifluoroacetylated substrates and a weak one against acetylated substrates
330
difference between class Ι and class IIb, in regard to class IIa HDACs, is the replacement of the
331
mechanistically important tyrosine residue near the catalytic zinc ion to a histidine. Bottemley et
332
al. demonstrated that a mutation of this histidine to tyrosine leads to significantly higher activity
333
towards acetylated lysines
334
comparison, two different mutants were produced, one where Y313 is replaced by phenylalanine
335
and another where it is replaced by histidine. Both show basically no activity against the
substrate.
20
Class
IIa
HDACs
only exhibit
high
activity 20
against
. The main
, but a reduced activity using trifluoroacetylated substrates. For
ACS Paragon Plus Environment
15
Biochemistry
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 56 57 58 59 60
Page 16 of 37
336
acetylated substrate which is in line with previous studies 13, 48. However, in contrast to class IIa
337
the activity of the Y313F mutant against the trifluoroacetylated substrate was not affected and
338
only slightly reduced in the Y313H mutant. This implies that the chemical activation by the
339
trifluoroacetyl group is strong enough to completely overcome the catalytic effect of the tyrosine.
340 341 342 343
DISCUSSION
344
highly conserved protein core structures is to understand how subtle structural differences e.g. at
345
the rim of the catalytic site pocket determine selectivity in molecular recognition of substrates or
346
inhibitors. HDAH from Bordetella/Alcaligene and PA3774 from P. aeruginosa are similar with
347
respect to overall sequence identity (49%) but the imposing L1-loops of both enzymes share only
348
very low sequence identity (21%). Accordingly, one could expect that the structure of the inner
349
core of the monomeric enzymes would be very similar as is the case for all Zn2+ containing
350
enzymes from HDAC family. However, it would not be possible to predict the structure and
351
orientation of the L1-loop in PA3774 with reliable precision. Although the sequence identity is
352
very low, the structural accordance of the L1-loops in HDAH and PA3774 is surprisingly high
353
(Fig. S10) and underlines the importance of its structure for assembly of subunits. While the
354
quaternary structure of PA3774 and HDAH is maintained by the L1-loop, the related APAH
355
from M. ramosa dimerizes through a different loop
356
polyamines rather than acetylated lysine residues. The different substrate specificity is supposed
357
to be largely dependent on distinct dimer arrangements defining access to the active site pocket.
358
Similarly, the physiologically occurring multi-enzyme complexes of human HDACs in different
359
compositions are believed to show different selectivity profiles than isolated enzymes 7, 49.
Impact of quaternary structure. One of the big issues with members of the HDAC family with
13
. The substrates of APAH are acetylated
ACS Paragon Plus Environment
16
Page 17 of 37
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 56 57 58 59 60
Biochemistry
360
Inference of the mechanism. Two types of deacetylation mechanisms have been previously
361
discussed for zinc dependent HDACs (Fig. S12). The first one is a general acid-base catalytic
362
pair mechanism of the histidine pair H143 (inner histidine), as well as the H144 (outer histidine)
363
and was originally proposed by Finnin and co-workers
364
mechanism based on DFT QM/MM studies of HDLP
365
ab initio QM/MM MD simulations on HDAC8 51 which suggest that the outer histidine acts as a
366
general base and acid while the inner histidine is stabilizing the negative charge of the oxyanion
367
intermediate. Most recently, Gantt and colleagues published a highly comprehensive study about
368
this subject 52. Their results favor the proton shuttle mechanism and are based on the fact that a
369
mutation of outer histidine decreases activity of HDAC8 more strongly than a mutation of the
370
inner histidine. In contrast, in our enzymological study on PA3774 both histidines are equally
371
important. Both single mutations of these histidines abolish activity completely against
372
acetylated lysine substrate proving the particular importance of both histidines for the enzyme
373
mechanism. Moreover, with the trifluoroacetylated substrate the effect appears to be even
374
inverted compared with class Ι HDAC8 (Table 2). There the “outer histidine mutant” is being
375
more active than the “inner histidine mutant”. This is also in line with the recently published
376
study on HDAC6, in which the corresponding “outer histidine mutant” is also more active than
377
the “inner histidine mutant”
378
based on HDAC8 and HDLP which are both as class Ι HDACs, whereas PA3774 and HDAC6
379
are classified as class IIb HDACs, the picture of a general mechanism for all HDAC classes is
380
probably oversimplified and has to be refined by further experiments and calculations.
22
50
36
. The second one is a proton shuttle
and furthermore on Born-Oppenheimer
. Since the current discussion of a general HDAC mechanism is
ACS Paragon Plus Environment
17
Biochemistry
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 56 57 58 59 60
Page 18 of 37
381
One key difference between these classes is the replacement of the outer aspartate-histidine dyad
382
to asparagine-histidine dyad in class IIb HDACs. These and other structural variations may give
383
rise to the contrary observed effect of the inner and outer histidine on enzyme activity.
384 385
In vivo function of PA3774. Against the background that PA3774 is not an APAH, the in vivo
386
function is still unclear. Since HDAH from Bordetella/Alcaligenes is highly homologous to
387
PA3774, it is obvious to speculate that these enzymes have similar functions and substrates.
388
HDAH is most active towards small peptides which contain at least one basic neighboring amino
389
acid, which is in line with the acidic electrostatic surface potential at the entrance tunnel of
390
HDAH. However, the electrostatic potential of PA3774 in this area is more basic to neutral and
391
therefore the substrate might be different. The high turnover rates of PA3774 against the
392
artificial HDAC substrates lead only to the suggestion that the natural substrate is a yet
393
unidentified lysine-acetylated protein. Recent studies revealed that many proteins in P.
394
aeruginosa are acetylated including influence factors and the DNA binding protein HU 53. It has
395
been suggested that these HU proteins have similar functions in bacteria as histones in
396
eukaryotes 54. The genome database of the P. aeruginosa strain PAO1 contains at least one HU
397
called hupB (gene number PA1805)
398
might be a substrate for PA3774. Further work to identify the substrates of PA3774 within the
399
acetylome of P. aeruginosa is ongoing.
26
. It is therefore tempting to speculate that this protein
400 401
ACS Paragon Plus Environment
18
Page 19 of 37
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 56 57 58 59 60
Biochemistry
402
CONCLUSION
403
This study reveals the first structure of the HDAC like amidohydrolase PA3774 from the
404
opportunistic pathogen P. aeruginosa. The distinct tetrameric quaternary structure comprising
405
dimers of “head-to-head” dimers significantly influences the accessibility of the active site and
406
represents a crucial determinant of molecular recognition and substrate selectivity. Similarly, the
407
closely related APAH from M. ramosa
408
completely different substrate selectivity. This leads us to conclude that a full comprehension of
409
the substrate selectivity of human HDACs and their bacterial homologs is only possible by a
410
detailed understanding of the respective oligomeric states and multi enzyme complexes.
411
Consequently, the composition and structure of homo- or heteromultimeric protein complexes
412
must be considered when designing selective active substances against HDACs and HDAC like
413
proteins.
13
forms a dimer through another loop resulting in
414 415 416
AUTHOR INFORMATION
417
Corresponding Author
418
Prof.
419
E-Mail:
[email protected] 420
Author Contributions
421
AK produced and crystallized the protein, solved the structures, analyzed the data. TW produced
422
and tested the mutants. AK and OY collected the diffraction data, solved and analyzed the
Dr.
Franz-Josef
Meyer-Almes,
Haardring
100,
64295
Darmstadt,
Germany
ACS Paragon Plus Environment
19
Biochemistry
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 56 57 58 59 60
Page 20 of 37
423
structures. FJMA conceived the project. The manuscript was written by AK and FJMA. All
424
authors have given approval to the final version of the manuscript.
425
Funding Sources
426
This work has been supported by the Deutsche Forschungsgesellschaft (grant GZ: ME 3122/2-1).
427 428
ACKNOWLEDGEMENT
429
We want to thank the beamline scientists on the PXII - X10SA at the Swiss Light Source for
430
support during data collection and Elmar Jaenicke (University Mainz) for performing and
431
analyzing the SEC/MALLS experiments. Michael Schröder is gratefully acknowledged for his
432
generous support at the lab. We also thank Javier Carrera-Casanova and Katharina van Pee for
433
their technical support during the crystallization experiments and Gregor Rolshausen
434
(Senckenberg Institute Frankfurt) for additional proof reading.
435 436
Supporting Information Available:
437
Fig. S1: Chromatograms and PAGEs of PA3774 purification steps
438 439
Fig. S2: The modified pET21a(+) vector construct, gene sequence, amino acid sequence and mutagenesis primer
440
Fig. S3: Crystals of PA3774
441
Fig. S4: Chromatogram of the analytical SEC
442
Fig. S5: MALLS experiment
443
Fig. S6: Interfaces in the quaternary structure
444
Fig. S7: Ligand electron density map of SATFMK bound to the PA3774
445
Fig. S8: Ligand electron densities maps of PFSAHA bound to the PA3774Y313F Mutant
446
Fig. S9: Close up of the PA3774H143A mutant structure
ACS Paragon Plus Environment
20
Page 21 of 37
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 56 57 58 59 60
Biochemistry
447
Fig. S10: L1-Loop comparison between PA3774 and HDAH
448
Fig. S11: Comparison of PA3774 with human HDACs and HDLP
449
Fig. S12: Comparison of the two discussed mechanisms for HDACs enzymes
450
Fig. S13: IC50 value determination and structure of SATFMK
451
Fig. S14: IC50 value determination and structure of PFSAHA
452
Fig. S15: Structural formula of Boc-Lys(TFA)-AMC and Boc-Lys(Ac)-AMCThis material is
453
available free of charge via the Internet at http://pubs.acs.org
454
455
REFERENCES
456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474 475 476 477 478 479 480 481 482
[1] Leipe, D. D., and Landsman, D. (1997) Histone deacetylases, acetoin utilization proteins and acetylpolyamine amidohydrolases are members of an ancient protein superfamily, Nucleic Acids Res. 25, 3693‐3697. [2] Kouzarides, T. (1999) Histone acetylases and deacetylases in cell proliferation, Current opinion in genetics & development 9, 40-48. [3] Archer, S. Y., and Hodin, R. A. (1999) Histone acetylation and cancer, Current opinion in genetics & development 9, 171-174. [4] Mwakwari, S. C., Patil, V., Guerrant, W., and Oyelere, A. K. (2010) Macrocyclic histone deacetylase inhibitors, Current topics in medicinal chemistry 10, 1423-1440. [5] Monneret, C. (2007) Histone deacetylase inhibitors for epigenetic therapy of cancer, Anticancer drugs 18, 363-370. [6] Gregoretti, I. V., Lee, Y. M., and Goodson, H. V. (2004) Molecular evolution of the histone deacetylase family: functional implications of phylogenetic analysis, Journal of molecular biology 338, 17-31. [7] 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. [8] Glozak, M. A., Sengupta, N., Zhang, X., and Seto, E. (2005) Acetylation and deacetylation of non-histone proteins, Gene 363, 15-23. [9] Choudhary, C., Kumar, C., Gnad, F., Nielsen, M. L., Rehman, M., Walther, T. C., Olsen, J. V., and Mann, M. (2009) Lysine acetylation targets protein complexes and co-regulates major cellular functions, Science 325, 834-840. [10] Vannini, A., Volpari, C., Filocamo, G., Casavola, E. C., Brunetti, M., Renzoni, D., Chakravarty, P., Paolini, C., De Francesco, R., Gallinari, P., Steinkuhler, 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, 15064-15069.
ACS Paragon Plus Environment
21
Biochemistry
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 56 57 58 59 60
483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525 526 527
Page 22 of 37
[11] Nielsen, T. K., Hildmann, C., Dickmanns, A., Schwienhorst, A., and Ficner, R. (2005) Crystal structure of a bacterial class 2 histone deacetylase homologue, J. Mol. Biol. 354, 107-120. [12] Sakurada, K., Ohta, T., Fujishiro, K., Hasegawa, M., and Aisaka, K. (1996) Acetylpolyamine amidohydrolase from Mycoplana ramosa: gene cloning and characterization of the metal-substituted enzyme, J. Bacteriol. 178, 5781‐5786. [13] Lombardi, P. M., Angell, H. D., Whittington, D. A., Flynn, E. F., Rajashankar, K. R., and Christianson, D. W. (2011) Structure of prokaryotic polyamine deacetylase reveals evolutionary functional relationships with eukaryotic histone deacetylases, Biochemistry 50, 1808-1817. [14] Millard, C. J., Watson, P. J., Celardo, I., Gordiyenko, Y., Cowley, S. M., Robinson, C. V., Fairall, L., and Schwabe, J. W. (2013) Class I HDACs share a common mechanism of regulation by inositol phosphates, Mol. Cell 51, 57-67. [15] Lauffer, B. E., Mintzer, R., Fong, R., Mukund, S., Tam, C., Zilberleyb, I., Flicke, B., Ritscher, A., Fedorowicz, G., Vallero, R., Ortwine, D. F., Gunzner, J., Modrusan, Z., Neumann, L., Koth, C. M., Lupardus, P. J., Kaminker, J. S., Heise, C. E., and Steiner, P. (2013) Histone deacetylase (HDAC) inhibitor kinetic rate constants correlate with cellular histone acetylation but not transcription and cell viability, J. Biol. Chem. 288, 26926-26943. [16] Bressi, J. C., Jennings, A. J., Skene, R., Wu, Y., Melkus, R., De Jong, R., O'Connell, S., Grimshaw, C. E., Navre, M., and Gangloff, A. R. (2010) Exploration of the HDAC2 foot pocket: Synthesis and SAR of substituted N-(2-aminophenyl)benzamides, Bioorg. Med. Chem. Lett. 20, 3142-3145. [17] Watson, P. J., Fairall, L., Santos, G. M., and Schwabe, J. W. (2012) Structure of HDAC3 bound to co-repressor and inositol tetraphosphate, Nature 481, 335-340. [18] Somoza, J. R., Skene, R. J., Katz, B. A., Mol, C., Ho, J. D., Jennings, A. J., Luong, C., Arvai, A., Buggy, J. J., Chi, E., Tang, J., Sang, B.-C., Verner, E., Wynands, R., Leahy, E. M., Dougan, D. R., Snell, G., Navre, M., Knuth, M. W., Swanson, R. V., McRee, D. E., and Tari, L. W. (2004) Structural Snapshots of Human HDAC8 Provide Insights into the Class I Histone Deacetylases, Structure 12, 1325-1334. [19] 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, 5048-5056. [20] Bottomley, M. J., Lo Surdo, P., Di Giovine, P., Cirillo, A., Scarpelli, R., Ferrigno, F., Jones, P., Neddermann, P., De Francesco, R., Steinkuhler, C., Gallinari, P., and Carfi, A. (2008) Structural and functional analysis of the human HDAC4 catalytic domain reveals a regulatory structural zinc-binding domain, The Journal of biological chemistry 283, 26694-26704. [21] Schuetz, A., Min, J., Allali-Hassani, A., Schapira, M., Shuen, M., Loppnau, P., Mazitschek, R., Kwiatkowski, N. P., Lewis, T. A., Maglathin, R. L., McLean, T. H., Bochkarev, A., Plotnikov, A. N., Vedadi, M., and Arrowsmith, C. H. (2008) Human HDAC7 harbors a class IIa histone deacetylase-specific zinc binding motif and cryptic deacetylase activity, The Journal of biological chemistry 283, 11355-11363. [22] Hai, Y., and Christianson, D. W. (2016) Histone deacetylase 6 structure and molecular basis of catalysis and inhibition, Nat. Chem. Biol. 12, 741-747.
ACS Paragon Plus Environment
22
Page 23 of 37
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 56 57 58 59 60
528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573
Biochemistry
[23] Miyake, Y., Keusch, J. J., Wang, L., Saito, M., Hess, D., Wang, X., Melancon, B. J., Helquist, P., Gut, H., and Matthias, P. (2016) Structural insights into HDAC6 tubulin deacetylation and its selective inhibition, Nat. Chem. Biol. 12, 748-754. [24] Stover, C. K., Pham, X. Q., Erwin, A. L., Mizoguchi, S. D., Warrener, P., Hickey, M. J., Brinkman, F. S., Hufnagle, W. O., Kowalik, D. J., Lagrou, M., Garber, R. L., Goltry, L., Tolentino, E., Westbrock-Wadman, S., Yuan, Y., Brody, L. L., Coulter, S. N., Folger, K. R., Kas, A., Larbig, K., Lim, R., Smith, K., Spencer, D., Wong, G. K., Wu, Z., Paulsen, I. T., Reizer, J., Saier, M. H., Hancock, R. E., Lory, S., and Olson, M. V. (2000) Complete genome sequence of Pseudomonas aeruginosa PAO1, an opportunistic pathogen, Nature 406, 959-964. [25] Bodey, G. P., Bolivar, R., Fainstein, V., and Jadeja, L. (1983) Infections caused by Pseudomonas aeruginosa, Reviews of infectious diseases 5, 279-313. [26] Winsor, G. L., Lam, D. K. W., Fleming, L., Lo, R., Whiteside, M. D., Yu, N. Y., Hancock, R. E. W., and Brinkman, F. S. L. (2011) Pseudomonas Genome Database: improved comparative analysis and population genomics capability for Pseudomonas genomes, Nucleic acids research 39, D596-D600. [27] Kramer, A., Herzer, J., Overhage, J., and Meyer-Almes, F.-J. (2016) Substrate specificity and function of acetylpolyamine amidohydrolases from Pseudomonas aeruginosa, BMC Biochemistry 17, 4. [28] Meyners, C., Wawrzinek, R., Kramer, A., Hinz, S., Wessig, P., and Meyer-Almes, F. J. (2014) A fluorescence lifetime-based binding assay for acetylpolyamine amidohydrolases from Pseudomonas aeruginosa using a [1,3]dioxolo[4,5-f][1,3]benzodioxole (DBD) ligand probe, Anal. Bioanal. Chem. 406, 4889-4897. [29] Powell, H. R., Johnson, O., and Leslie, A. G. W. (2013) Autoindexing diffraction images with iMosflm, Acta Crystallographica Section D 69, 1195-1203. [30] Evans, P. R., and Murshudov, G. N. (2013) How good are my data and what is the resolution?, Acta Crystallographica Section D 69, 1204-1214. [31] Vagin, A., and Teplyakov, A. (2010) Molecular replacement with MOLREP, Acta crystallographica. Section D, Biological crystallography 66, 22-25. [32] Emsley, P., Lohkamp, B., Scott, W. G., and Cowtan, K. (2010) Features and development of Coot, Acta crystallographica. Section D, Biological crystallography 66, 486-501. [33] Murshudov, G. N., Skubak, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R. A., Winn, M. D., Long, F., and Vagin, A. A. (2011) REFMAC5 for the refinement of macromolecular crystal structures, Acta crystallographica. Section D, Biological crystallography 67, 355-367. [34] Potterton, E., Briggs, P., Turkenburg, M., and Dodson, E. (2003) A graphical user interface to the CCP4 program suite, Acta Crystallographica Section D 59, 1131-1137. [35] Wegener, D., Wirsching, F., Riester, D., and Schwienhorst, A. (2003) A fluorogenic histone deacetylase assay well suited for high-throughput activity screening, Chemistry & biology 10, 61-68. [36] Finnin, M. S., Donigian, J. R., Cohen, A., Richon, V. M., Rifkind, R. A., Marks, P. A., Breslow, R., and Pavletich, N. P. (1999) Structures of a histone deacetylase homologue bound to the TSA and SAHA inhibitors, Nature 401, 188-193. [37] Lombardi, P. M., Cole, K. E., Dowling, D. P., and Christianson, D. W. (2011) Structure, mechanism, and inhibition of histone deacetylases and related metalloenzymes, Curr. Opin. Struct. Biol. 21, 735-743.
ACS Paragon Plus Environment
23
Biochemistry
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 56 57 58 59 60
574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619
Page 24 of 37
[38] Krissinel, E., and Henrick, K. (2007) Inference of Macromolecular Assemblies from Crystalline State, Journal of molecular biology 372, 774-797. [39] Nielsen, T. K., Hildmann, C., Riester, D., Wegener, D., Schwienhorst, A., and Ficner, R. (2007) Complex structure of a bacterial class 2 histone deacetylase homologue with a trifluoromethylketone inhibitor, Acta crystallographica. Section F, Structural biology and crystallization communications 63, 270-273. [40] Nielsen, T. K., Hildmann, C., Dickmanns, A., Schwienhorst, A., and Ficner, R. (2005) Crystal structure of a bacterial class 2 histone deacetylase homologue, Journal of molecular biology 354, 107-120. [41] Dowling, D. P., Gantt, S. L., Gattis, S. G., Fierke, C. A., and Christianson, D. W. (2008) Structural studies of human histone deacetylase 8 and its site-specific variants complexed with substrate and inhibitors, Biochemistry 47, 13554-13563. [42] Gantt, S. L., Joseph, C. G., and Fierke, C. A. (2010) Activation and inhibition of histone deacetylase 8 by monovalent cations, The Journal of biological chemistry 285, 60366043. [43] Micelli, C., and Rastelli, G. (2015) Histone deacetylases: structural determinants of inhibitor selectivity, Drug discovery today 20, 718-735. [44] Henkes, L. M., Haus, P., Jager, F., Ludwig, J., and Meyer-Almes, F. J. (2012) Synthesis and biochemical analysis of 2,2,3,3,4,4,5,5,6,6,7,7-dodecafluoro-N-hydroxy-octanediamides as inhibitors of human histone deacetylases, Bioorg. Med. Chem. 20, 985-995. [45] Nielsen, T. K., Hildmann, C., Riester, D., Wegener, D., Schwienhorst, A., and Ficner, R. (2007) Complex structure of a bacterial class 2 histone deacetylase homologue with a trifluoromethylketone inhibitor, Acta Crystallogr Sect F: Struct Biol Cryst Commun 63. [46] Riester, D., Wegener, D., Hildmann, C., and Schwienhorst, A. (2004) Members of the histone deacetylase superfamily differ in substrate specificity towards small synthetic substrates, Biochem. Biophys. Res. Commun. 324, 1116-1123. [47] Haggarty, S. J., Koeller, K. M., Wong, J. C., Grozinger, C. M., and Schreiber, S. L. (2003) Domain-selective small-molecule inhibitor of histone deacetylase 6 (HDAC6)-mediated tubulin deacetylation, Proceedings of the National Academy of Sciences of the United States of America 100, 4389-4394. [48] Moreth, K., Riester, D., Hildmann, C., Hempel, R., Wegener, D., Schober, A., and Schwienhorst, A. (2007) An active site tyrosine residue is essential for amidohydrolase but not for esterase activity of a class 2 histone deacetylase-like bacterial enzyme, The Biochemical journal 401, 659-665. [49] Micelli, C., and Rastelli, G. (2015) Histone deacetylases: structural determinants of inhibitor selectivity, Drug Discov. Today 20, 718-735. [50] Corminboeuf, C., Hu, P., Tuckerman, M. E., and Zhang, Y. (2006) Unexpected deacetylation mechanism suggested by a density functional theory QM/MM study of histone-deacetylase-like protein, Journal of the American Chemical Society 128, 45304531. [51] Wu, R., Wang, S., Zhou, N., Cao, Z., and Zhang, Y. (2010) A proton-shuttle reaction mechanism for histone deacetylase 8 and the catalytic role of metal ions, Journal of the American Chemical Society 132, 9471-9479. [52] Gantt, S. M. F., 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.
ACS Paragon Plus Environment
24
Page 25 of 37
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 56 57 58 59 60
620 621 622 623 624 625 626 627 628 629
Biochemistry
[53] Ouidir, T., Cosette, P., Jouenne, T., and Hardouin, J. (2015) Proteomic profiling of lysine acetylation in pseudomonas aeruginosa reveals the diversity of acetylated proteins, Proteomics. [54] Drlica, K., and Rouviere-Yaniv, J. (1987) Histonelike proteins of bacteria, Microbiological reviews 51, 301-319. [55] Magis, C., Taly, J. F., Bussotti, G., Chang, J. M., Di Tommaso, P., Erb, I., EspinosaCarrasco, J., and Notredame, C. (2014) T-Coffee: Tree-based consistency objective function for alignment evaluation, Methods Mol Biol 1079, 117-129. [56] Eddy, S. R. (2004) Where did the BLOSUM62 alignment score matrix come from?, Nature biotechnology 22, 1035-1036.
630 631
ACS Paragon Plus Environment
25
Biochemistry
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 56 57 58 59 60
Page 26 of 37
Table1: Data collection and refinement statistics PA3774 SATFMK
PA3774 Acetat
PA3774 H143A
PA3774 Y313F
unliganded
unliganded
unliganded
PA3774 Y313F PFSAHA
5G0Y
5G10
5G0X
5G13
5G12
5G11
P41212 81.7, 81.7, 205.2
P41212 81.7, 81.7 , 205.9
P41212
P41212
P41212
a, b, c (Å )
P41212 81.5, 81.5, 203.2
81.4, 81.4 , 204.1
81.6, 81.6, 205.2
81.8 81.8 204.9
α, β, γ (deg)
90, 90, 90
90, 90, 90
90, 90, 90
90, 90, 90
90, 90, 90
90, 90, 90
I/sd(I)
5.3 (2.2)
8.2 (2.1)
5.9 (2.4)
11.2 (4.8)
6.2 (2.7)
11.4 (6.9)
Wavelength (Å) Resolution range (Å) overall observations unique reflections Completeness (%)a
1.00001
0.97902
1.00002
0.97903
1.00001
1.00002
75.68 - 2.29
75.94 - 1.71
75.70 – 1.70
75.80 - 1.99
75.79 - 2.02
75.96 - 2.48
175199
578699
885648
750443
457059
288702
31717
75585
77011
48272
46280
25472
99.9 (100.0)
100.0 (100.0)
100.0 (99.9)
99.5 (94.1)
99.4 (96.8)
100.0 (100.0)
PA3774 Dataset PDB CODE Data Processing Space group
a
a
Multiplicity
5.5 (5.5)
7.7 (8.0)
11.5 (12.2)
15.5 (11.7)
9.9 (9.1)
11.3 (11.6)
0.117 (0.502)
0.062 (0.478)
0.088 (0.339)
0.044 (0.108)
0.092 (0.467)
0.050 (0.097)
Rcrystc
0.1870
0.1579
0.1343
0.1726
0.179
0.1678
Rfreed
0.2266
0.1937
0.1840
0.2172
0.224
0.2240
5953
6359
6307
5671
6122
6011
749
746
746
746
743
744
Rpimb Refinement
total atomse e
protein residues ligand moleculese
2
2
6
6
6
6
6
6
217
625
535
454
380
276
0.009
0.018
0.013
0.009
0.010
0.011
1.36
1.46
1.51
1.32
1.35
1.58
metal ionse watere Rmsd from ideal Bond lengths (Å) Bond angles (deg)
2
Ramachandran plot (%) most favoured
94.36
95.42
95.28
95.01
95.48
95.00
allowed
4.43
3.64
3.77
4.04
3.56
3.92
outliers a
1.21
0.94
0.94
0.94
0.96
1.08
b
Number in parentheses refer to the outer shell. Given the high multiplicity for the outer shells of these data sets, Rpim is a more appropriate measure of the data quality than Rmerge. c Rcryst = ∑||Fo| - |Fc|| / ∑ |Fo|, where Fo and Fc are the structure factor amplitudes from the data and the model, respectively. Intensity calculated from replicate. d Rfree is Rcryst with 5% of test set structure factors data. e Per asymmetric unit. 632 633
ACS Paragon Plus Environment
26
Page 27 of 37
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 56 57 58 59 60
Biochemistry
634 Table 2: Relative activities of PA3774 mutants in % Mutant
635
Boc-Lys(TFA)-AMC Boc-Lys(Ac)-AMC
Wild type
100 ± 2
100 ± 4
H143A
1.0 ± 0.1
≤ 0.05*
H144A
1.4 ± 0.1
≤ 0.05*
Y313F
98 ± 3
≤ 0.05*
Y313H
85 ± 3
≤ 0.05*
*lower limit of detection
636 637 638 639 640 641 642 643 644 645 646
ACS Paragon Plus Environment
27
Biochemistry
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 56 57 58 59 60
Page 28 of 37
647 648
Figure Legends
649
Fig. 1: Overall structure of PA3774 (A) Representation of one monomer in rainbow color
650
scheme, whereas red represents the C-terminus and blue the N-terminus. The enzyme adopts an
651
open α/β fold domain. Potassium ions are colored purple and the zinc ion grey. The marked L1-
652
loop region is the key factor for tetramer formation as shown in C and D. (B) A detailed
653
topology diagram with the same color coding as in A. Helices are numbered 1-14 and ß-sheets
654
are numbered a-l. The two ß-turns and the parallel 8 stranded ß-sheets are highlighted in blue
655
boxes. (C) Top view of the “head-to-head” dimer present in the asymmetric unit. One monomer
656
is colored in grey and the other in orange. The pivotal L1-loops which make extensive contacts
657
with the adjacent monomer are colored blue. (D) Tetrameric assembly inside the unit cell. The
658
color scheme of one dimer is identical to (C). The symmetry related “head-to-head” dimer is
659
colored dark grey and bright orange. The contacts between two head-to-head dimers eventually
660
forming the final tetramer are mainly mediated by L1-Loop interactions.
661 662
Fig. 2: Comparison with related structures: (A) The superimposition of PA3774 (grey) and
663
HDAH (green) shows similar tetrameric assembly. The inset shows the comparison of the L1-
664
loop which is responsible for tetramerization in more detailed. In both enzymes the L1-loop is
665
elongated compared with other enzymes of the HDAC family. The sequence identity is only 21%
666
but the orientation of the loops is very similar and consequentially the resulting assembly of
667
subunits. (B) The superimposition of PA3774 (grey) with HDAC8 (teal) shows the close
668
structural relationship between these enzymes. The main difference compared with the
669
monomeric HDAC8 is the elongated L1-Loop (highlighted in dark blue). (C) Comparison of
ACS Paragon Plus Environment
28
Page 29 of 37
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 56 57 58 59 60
Biochemistry
670
PA3774 “head-to-head” dimer (grey/dark-grey) and APAH dimer from M. ramosa
671
(orange/wheat): The grey and orange monomers are superimposed. A different loop insert in
672
APAH leads to entirely different assembly as indicated. Pdb-id’s used in this figure: 1ZZ0
673
(HDAH), 1T69 (HDAC8), 4ZUM (APAH)
674 675
Fig. 3: Sequence alignment of PA3774 with HDAH (the closest related structure), PA1409 and
676
PA0321 (APAHs from P. aeruginosa), APAH from M. ramosa, HDAC8 (a class Ι HDAC), the
677
second domain of HDAC6 (a class 2IIb HDAC) and the catalytic core domain of HDAC4 (a
678
class IIa HDAC). Orange dots indicate the zinc coordinating residues, green triangles indicate
679
the residues which are involved in the deacetylation process, the black box indicates the loop
680
region which is responsible for tetramer formation in PA3774 and HDAH and the red box
681
indicates the loop region which is responsible for dimer formation in APAH. The red framed
682
amino acids indicate sequence motifs which appear to be unique to APAHs. Alignment was
683
performed with T-Coffee 55, coloring indicates similarity according to BLOSUM62 56.
684 685
Fig. 4: Binding pocket of PA3774: (A) Cross section of the protein with focus on the binding
686
pocket. The figure demonstrates the crucial role of the “head-to-head” dimer assembly for
687
substrate binding. One monomer is colored in blue, the other one in orange and the inhibitor
688
SATFMK in red (stick representation). A wide area of the V-shaped pocket (orange) gets
689
covered by the adjacent monomer (blue) resulting in a substantially narrowed entrance surface.
690
(B) Anatomy of the binding pocket in stick representation. The specific amino acids which are
691
forming the binding pocket are indicated in orange. The amino acids of the adjacent subunit
ACS Paragon Plus Environment
29
Biochemistry
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 56 57 58 59 60
Page 30 of 37
692
responsible for restricting the entrance area are indicated in blue. The catalytic center consist of
693
basic and polar amino acids, the binding pocket is formed mostly by nonpolar amino acids.
694 695 696
Fig. 5: Stereo view of bound ligands to the wild type enzyme. Hydrogen bonds and metal
697
coordination are shown as green and black dashed lines, respectively. Zinc and potassium ions
698
are shown as grey and purple spheres, respectively. (A) Binding of the reaction product acetate
699
to the catalytic center. The electron density map shown is a Fo-Fc map before ligand
700
incorporation, contoured at 3 σ. (B) Binding of the inhibitor SATFMK. The ketone inhibitor is
701
binding in its gem-diol form and mimics the tetrahedral transition state. (C) Structural formula of
702
the trifluoromethylketone inhibitor SATFMK.
703 704
Fig. 6: (A) Stereoview of the ligand PFSAHA bound to the Y313F mutant. The ligand is shown
705
in orange, the residues of the binding pocket in grey and the teal colored residue belong to the
706
adjacent monomer. Hydrogen bonds and metal coordination are shown as green and black
707
dashed lines, respectively. The zinc is shown as a grey sphere. The main difference between the
708
structures is the mutated residue F313 which adopts an “inward” or “outward” conformation in
709
the inhibitor-free structure (superimposed yellow residue), whereas in the inhibitor complex only
710
the outward conformation is observed. (B) Structural formula of PFSAHA.
ACS Paragon Plus Environment
30
Page 31 of 37
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 56 57 58 59 60
Biochemistry
Figure 1: Overall structure of PA3774 177x149mm (300 x 300 DPI)
ACS Paragon Plus Environment
Biochemistry
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 56 57 58 59 60
Figure 2: Comparison with related structures 120x110mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 32 of 37
Page 33 of 37
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 56 57 58 59 60
Biochemistry
Fig. 3: Sequence alignment 172x102mm (300 x 300 DPI)
ACS Paragon Plus Environment
Biochemistry
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 56 57 58 59 60
Fig. 4: Binding pocket of PA3774 69x131mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 34 of 37
Page 35 of 37
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 56 57 58 59 60
Biochemistry
Fig. 5: Stereo view of bound ligands to the wild type enzyme 141x106mm (300 x 300 DPI)
ACS Paragon Plus Environment
Biochemistry
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 56 57 58 59 60
Figure 6: Stereo view of the ligand PFSAHA bound to the Y313F mutant 143x58mm (300 x 300 DPI)
ACS Paragon Plus Environment
Page 36 of 37
Page 37 of 37
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 56 57 58 59 60
Biochemistry
For Table of Contents Use Only
ACS Paragon Plus Environment