Subscriber access provided by La Trobe University Library
Article
Construction of thermophilic xylanase and its structural analysis Masahiro Watanabe, Harumi Fukada, and Kazuhiko Ishikawa Biochemistry, Just Accepted Manuscript • DOI: 10.1021/acs.biochem.6b00414 • Publication Date (Web): 13 Jul 2016 Downloaded from http://pubs.acs.org on July 29, 2016
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
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 47
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
Construction of thermophilic xylanase and its structural analysis
2 Masahiro Watanabe1,2, Harumi Fukada3 and Kazuhiko Ishikawa1,4
3 4 5
1
6
Technology (AIST), 3-11-32 Kagamiyama, Higashi-Hiroshima, Hiroshima 739-0046, Japan
7
2
8
Science and Technology (AIST), 3-11-32 Kagamiyama, Higashi-Hiroshima, Hiroshima
9
739-0046, Japan
Biomass Refinery Research Center, National Institute of Advanced Industrial Science and
Research Institute for Sustainable Chemistry, National Institute of Advanced Industrial
10
3
11
Gakuen-cho, Nakaku, Sakai, Osaka 599-8531, Japan
12
4
13
Technology (AIST), 1-8-31 Midorigaoka, Ikeda, Osaka, 563-8577, Japan
Graduate School of Life and Environmental Sciences, Osaka Prefecture University, 1-1
Biomedical Research Institute, National Institute of Advanced Industrial Science and
14 15
*Running title: Construction of thermophilic xylanase
16 17
To whom correspondence should be addressed: 3-11-32 Kagamiyama, Higashi-Hiroshima,
18
Hiroshima 739-0046, Japan. Tel.: +81-82-420-8258, Fax: +81-82-420-7820, E-mail:
19
[email protected] 20 21
Keywords: biomass; cellulase; hemi-cellulase; protein engineering; thermostability; X-ray
22
crystallography; differential scanning calorimetry; fungus
23
Background: Thermophilic xylanase holds a great deal of potential in many industrial
24
applications for biomass reduction.
1 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
25
Results: By modifying the N-terminal region of a xylanase, thermophilic mutants with a high
26
melting temperature were constructed, and the structure of the mutant was analyzed.
27
Conclusion: By increasing C-C contacts, stabilizing loop structure, and introducing disulfide
28
bond in the N-terminal region, the optimum temperature and thermostability of the xylanase
29
was improved.
30
Significance: The thermostability of a xylanase was improved by over 20°C by modifying the
31
N-terminal region.
32 33
ABSTRACT
34
The Glycoside hydrolase family 11 xylanase has been utilized in a wide variety of industrial
35
applications, from food processing to kraft pulp bleaching. Thermostability enhances the
36
economic value of industrial enzymes by making them more robust. Recently we solved the
37
crystal structure of an endo-ß-1,4-xylanase (GH11) from mesophilic Talaromyces
38
cellulolyticus, named XylC. Ligand-free XylC exists to two conformations (open/closed
39
forms). We found that the “closed” structure possessed an unstable region within the
40
N-terminal region far from the active site. In this study, we designed the thermostable xylanase
41
by the structure-based site-directed mutagenesis on the N-terminal region. In total nine
42
mutations (S35C, N44H, Y61M, T62C, N63L, D65P, N66G, T101P, and S102N) and an
43
introduced disulfide bond of the enzyme were contributed to the improvement in
44
thermostability. By combining the mutations, we succeeded in constructing a mutant of which
45
the melting temperature was partially additively increased by over 20°C (measured by a 2 ACS Paragon Plus Environment
Page 2 of 47
Page 3 of 47
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
46
differential scanning calorimetry) and the activity was additively enhanced at elevated
47
temperatures, without loss of the original specific activity. The crystal structure of the most
48
thermostable mutant was determined at 2.0-Å resolution to elucidate the structural basis of
49
thermostability. From the crystal structure of the mutant, it was revealed that the formation of a
50
disulfide bond induces new C-C contacts and a conformational change in the N-terminus. The
51
resulting induced conformational change in the N-terminus is key for stabilizing this region and
52
for constructing thermostable mutants without compromising the activity.
53 54
INTRODUCTION
55
Cellulosic materials constitute most of the biomass on earth and can be converted into
56
biofuels or bio-based materials. Such conversions utilize fermentable sugars obtained from
57
cellulosic biomass by the activity of cellulase-related enzymes (1-3). Xylan is a major structural
58
component of the plant cell wall and is the second most abundant renewable biomass resource
59
(4, 5). Xylan consists of a ß-1,4-D-xylan backbone with short side chains of O-acetyl,
60
ß-L-arabinofuranosyl, D-beta-glucuronic acid, and phenolic acid (6). Cellulose and xylan are
61
closely linked together in plant cell walls (7). Thus, cellulases and hemicellulases work
62
coordinately in the enzymatic degradation of these polysaccharides. Filamentous fungi produce
63
a wide spectrum of enzymes for degrading cellulose and xylan (8). Xylanases (endo-
64
ß-1,4-xylanases; EC 3.2.1.8) catalyze the hydrolysis of the ß-1,4 bonds of xylan and are thus
65
important enzymes for the degradation of hemicellulosic polysaccharides in lignocellulosic
66
biomass (9, 10). Based on their amino acid sequence similarities, xylanases are classified 3 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
67
primarily into glycoside hydrolase (GH; http://www.cazy.org/Glycoside-Hydrolases.html)
68
families 10 and 11 (11). GH10 xylanases generally have a molecular weight ≥30 kDa with a
69
low pI value, whereas GH11 xylanases are generally smaller (approximately 20 kDa) with a
70
high pI value (12). The crystal structures of xylanases show that GH10 enzymes fold into a
71
(ß/α)8-barrel (13-15), whereas GH11 enzymes have a ß-jelly roll structure (16-18).
72
Talaromyces cellulolyticus (19), which was isolated by Yamanobe et al. (20),
73
produces an abundance of cellulolytic enzymes. Fujii et al. (21) reported that T. cellulolyticus
74
culture supernatant has a higher cellulase specific activity and yields more glucose from
75
lignocellulosic materials than does Trichoderma reesei (T. reesei) culture supernatant. Six of
76
the seven GH11 xylanases produced by T. cellulolyticus were recently cloned and characterized
77
by Watanabe et al. (22). Of the six xylanases characterized, TcXylC had the highest activity
78
against xylan, despite the low expression level of the gene (22). In addition, a system for the
79
over-expression and purification of recombinant TcXylC using E.coli was developed and the
80
crystal structure of TcXylC was solved at 1.98-Å resolution (23). Structural analysis revealed
81
that TcXylC consists of a common ß-jelly roll structure exhibiting a two-state structure (open-
82
and closed-form) in the asymmetric unit. RMSD of the Cα atoms between the open and closed
83
forms was 0.55 Å. However, the only closed form possesses a tunnel structure of the active site
84
formed by Trp52 and Pro159 (23). The tunnel structure retained hydrogen-bonding networks
85
with 10 water molecules and is thought to mimic the structure of the closed form. The
86
N-terminal region including Trp52 in the closed form exhibits high B-factor values (23).
87
Conformational changes in the flexible region are presumed to induce and accelerate the 4 ACS Paragon Plus Environment
Page 4 of 47
Page 5 of 47
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
88
enzymatic reaction. Although the structures and functions of GH11 xylanases from many
89
species have been well characterized (24-26) and compared (14, 16), the structural basis of
90
xylanase thermostability remains unclear, due in part to the challenge of selecting an
91
appropriate residue (or location) for site-directed mutagenesis, as even minimal substitution of
92
residues can lead to misfolding of the target protein (27-29). Some disulfide bond designs
93
tested increased the irreversible thermostability of the B. circulans xylanase, but not all
94
enhanced the activity of the enzyme at elevated temperatures (30).
95
TcXylC exhibits the highest activity of the xylanases produced by T. cellulolyticus
96
and has been expected to be used for industrial application. In this study, therefore, we
97
engineered TcXylC to improve both its thermostability and optimum temperature using
98
site-directed mutagenesis and elucidated the structural basis for thermostabilization.
99 100
EXPERIMENTAL PROCEDURES
101
Cloning of TcXylC and its variants.
102
The genes encoding TcXylC without the putative signal peptide (34 amino acid
103
residues) and its variants were cloned into pET11a (Novagen) at the NdeI and BamHI
104
restriction enzyme sites, such that the NdeI site (CATATG) included the initiator methionine
105
codon (23). All mutant genes were constructed by PCR using a PrimeSTAR Max DNA
106
Polymerase kit (Takara Bio) and the primers listed in Supplemental Table 1.
107 108
Protein expression and purification. 5 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
109
Plasmids encoding TcXylC and its mutants were transformed into E. coli BL21
110
(DE3) cells (Novagen). Cells were cultured at 37°C in Luria-Bertani medium containing 100
111
mg/L of ampicillin-Na. When the culture reached an OD600 of 0.6, expression of recombinant
112
protein was induced by addition of isopropyl ß-D-1-thiogalactopyranoside (IPTG) to a final
113
concentration of 200 µM. The culture was subsequently incubated at 20°C for 16 h, after
114
which the cells were harvested by centrifugation at 5,000 × g for 15 min at 4°C. The resulting
115
cell pellets were resuspended in 20 mM Tris-HCl (pH 8.0) containing 50 mM NaCl. Soluble
116
protein was extracted by three freeze-thaw cycles (freezing at −80°C for 1 h followed by
117
complete thawing at room temperature) and recovered by centrifugation at 35,870 × g for 20
118
min at 4°C. The recombinant proteins were purified using a modified method described
119
previously (23). The supernatants were applied to a Q-HP column (GE Healthcare)
120
equilibrated in 20 mM Tris-HCl (pH 8.0) and eluted with an ascending salt concentration
121
gradient. Wild-type (WT) and mutant TcXylC proteins were eluted from the column at
122
between 200 and 400 mM NaCl. Fractions containing the recombinant protein were collected,
123
dialyzed against 20 mM Tris-HCl (pH 8.0)/2 M NaCl, and applied to a Butyl HP column (GE
124
Healthcare) equilibrated in the same buffer. Proteins were eluted with an ascending salt
125
concentration gradient. The recombinant protein was found to elute at approximately 1 M
126
NaCl. As a final step, the eluted protein was applied directly to a Superdex 200 16/60 gel
127
filtration column (GE Healthcare) equilibrated with 20 mM Tris-HCl (pH 8.0)/50 mM NaCl.
128
The purity of the target protein was assessed using SDS-PAGE, and the xylanase activity was
6 ACS Paragon Plus Environment
Page 6 of 47
Page 7 of 47
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
129
also determined (Supplemental Fig. 1). All TcXylC mutants were purified using a method
130
similar to that used for the WT protein.
131 132
Protein assay.
133
All protein concentrations were determined according to the BCA method (31),
134
with bovine serum albumin as a standard. Protein solution (25 µL) was mixed with 200 µL of
135
BCA reagent (Thermo Scientific Pierce) and then incubated at 37°C for 30 min, after which
136
the concentration was determined by measuring the absorption at 570 nm in comparison with
137
standards of known concentration.
138 139
Enzyme assay.
140
Xylanase activity was measured as reported previously (32) using birch-wood xylan
141
as the substrate under various buffer conditions between pH 4.0 and 8.0 (33). Liberation of
142
reducing sugar was determined using the dinitrosalicylic acid (DNS) method (32) with xylose
143
as a standard. Enzyme reaction was started by mixing 50 µL of an appropriate dilution of the
144
enzyme solution with 0.45 mL of substrate solution (1 %) in 50 mM sodium acetate buffer (pH
145
5.0) and incubated for 10 min at 50°C. The reaction was stopped by addition of 0.75 mL of
146
DNS solution and boiling for 5 min. Reducing sugars were determined by measuring the
147
absorption at 540 nm. One unit of xylanase activity was defined as the quantity of enzyme
148
required to liberate 1 µmol of xylose equivalent per min. For the estimation of kinetic
149
parameters, the initial reaction rates of the enzymes were measured at substrate concentrations 7 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
Page 8 of 47
150
ranging from 0.1 to 5 mg/mL. The kinetic parameters Km and kcat of the enzymes
were
151
calculated by the nonlinear least-squares method (34). Optimum temperature for the xylanase
152
activities of the mutants were examined at temperatures ranging between 45 and 75°C. The
153
activity was assessed by incubating the enzyme (1.0 mg/mL) in 100 µL of 50 mM sodium
154
acetate buffer (pH 5.0) containing 50 mM NaCl for 10 min at the respective temperatures.
155 156
Differential scanning calorimetry (DSC).
157
Thermal stability of the enzymes was examined using differential scanning
158
calorimetry (DSC). The mutants in 50 mM sodium acetate buffer (pH 5.0) were used at a final
159
concentration of 1.0 mg/mL. A nanoDSC instrument (TA Instruments) was used at a scanning
160
speed of 60°C/h. Control runs in the absence of protein were carried out before and after each
161
sample run. DSC scans in the presence of protein were performed two or three times for each
162
protein examined. DSC scans for the proteins was also carried out with xylobiose (competitive
163
inhibitor of xylanase; X2) (36 mM).
164 165
Crystallization.
166
The purified thermostable mutant was dialyzed against 20 mM Tris-HCl (pH 8.0)/50
167
mM NaCl and then concentrated to 13 mg/mL using Vivaspin 20-10K centrifugal filter units
168
(GE Healthcare). Initial crystallization screening of the mutant protein was performed
169
manually using the crystallization screening kits of Crystal Screen HT (Hampton Research),
170
JCSG-plus (Molecular Dimensions) and Wizard Classic 1 and 2, 3 and 4 (Rigaku) 8 ACS Paragon Plus Environment
Page 9 of 47
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
171
crystallization regents according to the sitting-drop vapor diffusion method at 293 K in 96-well
172
plates. Each drop (0.5 µL) was mixed with 0.5 µL of reservoir solution and then equilibrated
173
with 60 µL of reservoir solution. After a week, well-formed crystals were obtained using
174
JCSG-plus condition no. 20 (10 % [w/v] polyethylene glycol 8,000, 0.1 M Tris-HCl [pH 7.0],
175
0.2 M magnesium chloride), and was further optimized to 13 % (w/v) polyethylene glycol
176
8,000, 0.1 M sodium citrate (pH 5.0), and 0.2 M magnesium chloride. Optimized drops were
177
obtained by mixing 1 µL of the protein solution with 1 µL of reservoir solution (2 µL total) and
178
equilibrating with 1 mL of reservoir solution using the hanging-drop vapor diffusion method at
179
293 K in 24-well plates.
180 181
X-ray data collection.
182
Crystals obtained as described above were soaked in cryoprotectant solution
183
consisting of 30 % (v/v) ethylene glycol, 13 % (w/v) polyethylene glycol 8,000, 0.1 M sodium
184
citrate (pH 5.0), and 0.2 M magnesium chloride. The soaked crystals were then collected with a
185
cryo-loop and flash-cooled under a cryo-stream of nitrogen gas at 100 K. X-ray data were
186
collected at the SPring-8 BL44XU (Hyogo, Japan). The dataset was collected at a single
187
wavelength of 0.9 Å using a Rayonix MX300HE detector. The crystal was rotated 200° with an
188
oscillation angle of 0.5° per frame (total of 400 frames). The distance from the crystal to the
189
detector was 300 mm. Data were processed using the HKL2000 software package (35).
190 191
Structure solution and refinement. 9 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
192
Crystals of XylCmt9 were prepared according to the hanging-drop vapor diffusion
193
method, as described above. The initial phase of the XylCmt9 structure was determined by
194
molecular replacement with PHASER (36) using WT structure (PDB # 3WP3) as the search
195
model. Manual adjustment of the models and model refinement were carried out using COOT
196
(37), and CNS (38) and REFMAC5 (39). Water molecules were added to the models using
197
ARP/wARP (40), and the 2Fo-Fc and Fo-Fc maps were then manually inspected. The
198
stereochemical properties of the models were assessed using PROCHECK (41) and the
199
validation tools of COOT. Figures were created using CCP4MG (42).
200 201
RESULTS
202
Structure-based sequence homology of various GH11 xylanases.
203
Seven ORFs exhibiting high homology to GH11 xylanase have been identified in
204
the T. cellulolyticus genome (43). In the seven ORFs, six GH11 xylanases were recently
205
cloned and characterized by Watanabe et al. (22). They reported that the six xylanases
206
showed similar optimum temperature and pH for enzyme activity. In the six xylanases,
207
TcXylC exhibits the highest activity and its crystal structure was determined (23). Therefore,
208
TcXylC was selected as the construction of thermophilic xylanase in this study. TcXylC
209
consists of 14 ß-sheets (designated A2 through A6 and B1 through B9) and one α-helix, as
210
shown in Figure 1. TcXylC exhibits 30~60 % amino acid sequence identity with other
211
xylanases. Alignment analysis indicates that most of the sequences are similar and include
212
two catalytic glutamic acid residues (Fig. 1). Watanabe et al. (22) suggested that the 10 ACS Paragon Plus Environment
Page 10 of 47
Page 11 of 47
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
213
N-terminus (Met1-Arg33) of TcXylC (Fig. 1) serves as a signal sequence of the enzyme.
214
Furthermore, the side chain of Gln34 faces completely outward in the structure of TcXylC.
215
Therefore, we decided to remove the N-terminus of TcXylC up to Glu34 for the preparation
216
of the recombinant enzymes. The recombinant TcXylC without the N-terminus up to Gln34
217
was prepared and produced by the previous methods described in the Experimental
218
Procedures. The activity of enzyme was not influenced by the above modification and the
219
enzyme was designated as WT_TcXylC. It was reported that some of the alkalophilic
220
xylanases were activated by metal ions (44, 45). In the case of TcXylC and their mutants,
221
however, the activity was not affected by the presence of EDTA (10 mM) and Mg2+ (2.0 - 50
222
mM) (data not shown).
223 224
Structure-based targeting of site-directed mutagenesis sites for constructing a
225
thermophilic xylanase.
226
Analysis of the TcXylC crystal structures determined (23) revealed that the two
227
different conformations (open and closed forms) were observed in the asymmetric unit. In the
228
closed form, the N-terminal region exhibits high B-factor values (23). The region with high
229
B-factor values is the flexible and/or unstable region in the protein. For constructing a
230
thermostable mutant, it was key to improve the stability of the above structures. The
231
substrate-binding region is important for enzyme activity. Therefore, we focused on the
232
N-terminal region for constructing thermostable mutants using the structure of the closed
11 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
233
form of TcXylC (PDB; 3WP3) (23). The target mutation sites for the purpose were indicated
234
in Fig. 2.
235
In the N-terminus of WT, Ser35 is located close to Thr62 (neighboring ß-strand)
236
(Fig. 2). The distance between Ser35 and Thr62 is ideal for formation of a disulfide bond via
237
S35C/T62C substitutions (their Cα-Cα and Cß-Cß distances are estimated to be 4.43 Å and
238
4.77 Å, respectively (23)). Therefore, we attempted to introduce a disulfide bond via these
239
substitutions. In the case of SspXyl, hydrophobic interactions between Met67 and Leu69 at
240
the A2 ß-sheet are observed (46). These residues are not conserved at the N-terminus of the
241
ß-sheet (A2) (Fig. 1) in TcXylC. Met67 and Leu69 in SspXyl correspond to Tyr61 and Asn63
242
in TcXylC, respectively (Fig. 1). Therefore, we constructed the mutants in which Tyr61 and
243
Asn63 were replaced with Met and Leu residues (Y61M and N63L), respectively in order to
244
increase the C-C contact in the corresponding region in TcXylC. At the position of Asn44, at
245
least one or two hydrogen bond(s) with the backbone of the adjacent strand (B2) would be
246
expected to form by substitution of Asn44 with a His residue (N44H) (Fig. 2A). Proline
247
residue is known to impart a rigid and stable conformation to ß-turn or loop structures (47-49).
248
Regarding selection of a site for introducing a Pro residue to enhance the stability, we
249
selected Asp65, which is located in the loop region (between A2 and A3) (Fig. 2A). Sequence
250
homology analysis (Fig. 1) indicated that Asp65-Asn66 of TcXylC corresponds to
251
Pro31-Gly32 of TrXyl. We therefore performed D65P and N66G substitutions at the
252
N-terminus of the ß-strand (A2) of TcXylC. From the sequence homology, furthermore, Pro
253
is observed at the position corresponding to Thr101 of TcXylC (located in the loop region 12 ACS Paragon Plus Environment
Page 12 of 47
Page 13 of 47
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
254
between A5 and B5) in some xylanases (Fig. 1). Therefore, we introduced a Pro residue at the
255
position of Thr101 in the loop region (Fig. 2A). Based on sequence homology, we introduced
256
T101P and S102N substitutions in TcXylC (A5) (Fig. 2A). A typical ß-turn was observed at
257
Asp65-Asn66 but not at Thr101-Ser102.
258
The specific activity and temperature dependence of activity were examined for
259
each mutant. For all mutants prepared, the significant change of the specific activities at 45°C
260
were not observed (Table 1). The temperature dependencies of the enzyme activity were
261
examined at different temperatures. Fig. 3 shows the temperature dependent activity of the
262
mutants. For Y61M, N63L, D65P and N66G mutants, significant instability was observed at
263
60°C in comparison with WT (Fig. 3). However, The Y61M/N63L and D65P/N66G double
264
mutations retain their activities at 60°C (Fig. 3). For the other mutants (S35C, N44H, T62C,
265
T101P, and S102N and S35C/T62C), a decrease in the activity at 60°C was not observed (Fig.
266
3). Therefore, the temperature dependent activity (45-75°C) for the mutants excluding Y61M,
267
N63L, D65P and N66G was examined. The results indicate that Y61M/N63L, D65P/M66G,
268
S35C, N44H, T62C, S35C/T62C, T101P, and S102N mutations shifted the temperature
269
optima of the enzymes (Fig. 3). Thermal stability and the melting temperatures (Tm) of WT
270
and mutant enzymes were also examined using DSC (Table 1, Fig. 4). Table 1 shows the
271
values for melting temperatures (Tm), ∆H, and ∆S for heat denaturation process. The result
272
that the Tm values of the WT, S102N, and S35C/T62C TcXylC mutants were estimated to be
273
58.9, 60.0, and 74.5°C, respectively indicating that S102N, D65P/M66G, and S35C/T62C
274
TcXylC mutants enhanced the thermostability of the enzyme (Table 1, Fig. 2B). However, for 13 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
275
the other mutants, improvements in the thermostability corresponding to Tm of the enzyme
276
were not observed (Table 1, Fig. 4). Thermal stability of WT and mutant enzymes with the
277
substrate analogue xylobiose (X2) (36 mM) was also examined using DSC (the detailed data
278
not shown). Most of the values of Tm were increased (Table 1).
279
For the S35C/T62C substitution, the relative activity at the high temperature was
280
markedly increased (Fig. 3). Furthermore, the Tm of the S35C/T62C mutant was over 15°C
281
higher than that of the WT enzyme (Table 1, Fig. 4). However, the S35C and T62C single
282
substitution mutants did not at over 65°C compared with WT (Fig. 3). The above result
283
suggests that one disulfide bond is introduced in the enzyme via the S35C/T62C substitution
284
and contributes to the thermostability greatly.
285
Thus, we decided to introduce a total of nine point mutations that enhanced the
286
activity of the enzyme. The mutant gene, designated XylCmt9 (encoding S35C, N44H, Y61M,
287
T62C, N63L, D65P, N66G, T101P, and S102N substitutions) was constructed as described
288
above in the Experimental Procedures. The prepared XylCmt9 mutant produced the highest Tm
289
(80.5°C) (Table 1, Fig. 4) of all of the mutants examined, without any loss in original specific
290
activity (Fig. 3).
291 292
Structure of thermostable xylanase (XylCmt9).
293
In order to examine the structural basis of the enhanced thermostability of the
294
enzyme, we analyzed the crystal structure of the XylCmt9 variant. Crystals of XylCmt9 were
295
prepared (approximately 300 × 300 × 10 µm in size) and the structure was determined, as 14 ACS Paragon Plus Environment
Page 14 of 47
Page 15 of 47
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
296
described in the Experimental Procedures. The prepared crystals belonged to space group P21,
297
with unit cell a = 75.9 Å, b = 179.8 Å, c = 93.3 Å, and ß = 103.8°. Data collection and
298
refinement statistics are shown in Table 2. In the asymmetric unit, 12 molecules of XylCmt9
299
were identified; the Matthews coefficient (50) (VM) was calculated as 2.50 Å3 Da−1, with a
300
solvent content of 50.9 % (v/v). After refinement, R factors of the model were estimated as
301
Rwork = 18.0 % and Rfree = 22.8 %. A Ramachandran plot (51) for the model showed that 94.8 %
302
of the residues were in the most favored regions, with 4.4 % of the residues in additional
303
allowed regions. The coordinates and structure factors for XylCmt9 were deposited in Protein
304
Data Bank under the accession code 5HXV. The overall structures of 12 molecules of XylCmt9
305
were built from Cys35 to Ser223, and the root mean square deviation (RMSD) was less than
306
0.47 Å over 189 Cα atoms. The RMSD value indicated that the overall structures of the 12
307
molecules of XylCmt9 were almost identical (Fig. 5A). It was determined that the overall
308
structure of XylCmt9 was almost identical to that of the WT except for the N-terminal region.
309
In spite of high amounts of magnesium chloride (200 mM) in the crystallization reagent, the
310
significant electron density map corresponding to Mg2+ was not observed. All nine mutations
311
were identified in the structure. The chain A molecule of the 12 molecules (PDB; 5HXV)
312
exhibited relatively low B-factor values. Superimposition of chain A and closed form of WT
313
showed the identity of the whole structure (Fig. 5B). It was difficult to distinguish the
314
difference between open and closed forms in XylCmt9. Using the chain A molecule, therefore,
315
the difference of the N-terminal region was examined. As expected from the above result, the
316
disulfide bond between Cys35 and Cys62 was observed in XylCmt9 (Fig. 6A). Their Cα-Cα 15 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
317
and Cß-Cß distances were changed and the N-terminal region was slightly pulled into the
318
molecule by the newly introduced disulfide bond. It is difficult to compare the B-factor values
319
between the different crystals (WT and XylCmt9). However, the B-factor values of the region
320
were relatively low in XylCmt9. In addition, the C-C contacts between Ile36 Cδ1 and Leu63 Cδ1
321
(bond length = 3.77 Å) and between Met61 Cε and Leu63 Cδ2 (bond length = 3.65 Å) in A2 were
322
observed as good bond in XylCmt9 (Fig. 6A). Both Met61 and Leu63 were found to be essential
323
for the formation of a hydrophobic network around the N-terminus when the disulfide bond was
324
formed (Fig. 6A). As expected, the imidazole ring of His44 Nδ1 atom and the O atom of the
325
Gly64 main chain (neighboring strand A2) is found to form a new hydrogen bond at 3.23 Å in
326
XylCmt9 as a result of the N44H substitution (Fig. 6B). At a ß-turn (between A2 and A3),
327
Pro65 and Gly66 were observed without the conformational change in the region. At the loop
328
region (between A5 and B5), Pro101 and Asn102 were observed with the formation of new
329
hydrogen bonds between the N and O atoms of Asn102 and Ser215 main chain in the
330
neighboring ß-strand (A4) (Fig. 6C).
331 332
Characterization of the mutants.
333
Xylanase activity was measured by incubating the enzyme at 50°C using 1 % (w/v)
334
birch-wood xylan as the substrate. XylCmt9 showed an optimum pH of 4.0 - 5.0, similar to
335
WT. Compared with WT, a slight increase in the specific activity was observed for XylCmt9
336
(Table 3). The kinetic parameters Km and kcat of the enzymes were calculated by the nonlinear
16 ACS Paragon Plus Environment
Page 16 of 47
Page 17 of 47
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
337
least-squares method (Table 3). The increased value of kcat may be due to a faster rate of
338
product release, and the higher mobility may be associated with higher activity.
339 340
DISCUSSION
341
In this study, we constructed thermostable mutants of TcXylC using protein
342
engineering methods and examined the structural basis for thermostabilization. Crystal
343
structure analysis revealed two conformers of TcXylC structure (open and closed forms) (23).
344
The RMSD between closed and open forms were estimated to be 0.55 Å over 189 Cα (23).
345
The structure of the closed form may mimic the forming enzyme–substrate complex. In the
346
closed form, the N-terminal region, located far from the catalytic site, exhibits high B-factor
347
values and appeared to be a suitable target for the construction of a thermostable xylanase.
348
From BRENDA database (52), one of the xylanases (EC 3.2.1.8) has observed the activation
349
in the presence of Mg2+ (maximum 6 mM) (45). In the case of TcXylC and their mutants,
350
however, the activity was not affected by the presence of EDTA and Mg2+ (data not shown).
351
Furthermore, the significant electron density map corresponding to Mg2+ was not observed in
352
its crystal structure. From these results, it was clarified that TcXylC was not activated by the
353
bound metal ions.
354
Thus, we focused on stabilizing the N-terminal region to enhance the
355
thermostability. The effects of the mutations on the thermostability were evaluated based on
356
the Tm values determined by DSC, and the temperature dependence of the activity for each
357
mutant. Result of Tm values determined by DSC and the enzyme activity measurement 17 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
Page 18 of 47
358
experiments differed. Most of the values of Tm of the mutants were lower than 60°C (WT)
359
(Table 1, Fig. 4). For the mutants excluding Y61M, N63L, D65P, and N66G, however,
360
significant increases in activity were observed at 65°C (Fig. 3). The result suggests that the
361
enzyme is stabilized by the substrate. Therefore, thermal stability of the enzymes with the
362
substrate analogue xylobiose (X2) was measured by DSC. Most of the values of Tm of the
363
mutants were increased by adding X2 (36 mM) (Table 1). The increased values of Tm of the
364
mutants are dependent on the ligand binding (53). However, disagreement between the values
365
of Tm by DSC and activity measurement experiments was also observed. Similar
366
disagreement was observed in the case of xylanase mutants from B. circulans (30). Therefore,
367
it is speculated that the protein unfolding process estimated by DSC is not necessarily same as
368
the thermostabilization of the enzyme.
369
We gave priority to the kinetic results obtained from the enzyme activity temperature
370
dependence experiments. In the mesophilic xylanase SspXyl, a significant C-C contact is
371
observed between Met67 and Leu69 in the structure of N-terminal region (46). We have no
372
structural information of Y61M/N63L. From the analysis of the crystal structure of XylCmt9
373
(Fig. 6A), however, the expected C-C contact between Met61 Cε and Leu63 Cδ2 seems to
374
contribute the thermostability in Y61M/N63L. Five ß-turns (Asn44-Asn45-Gly46-Tyr47,
375
between
376
Ser132-Ser133-Gly134-Leu135, between B6 and B9; Ser143-Asn144-Gly145-Gly146,
377
between B9 and B8; and Ser160-Ile161-Glu162-Gly163, between B8 and B9) were observed in
378
the structure of WT (Fig. 1). The region Asp65-Asn66-Gly67-Glu68 forms a typical ß-turn I.
B1
and
B2;
Asp65-Asn66-Gly67-Glu68,
18 ACS Paragon Plus Environment
between
A2
ans
A3;
Page 19 of 47
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
379
The single substitution mutations D65P and N66G reduced the Tm. But in the double mutant
380
(D65P/N66G), the Tm increased relative to WT (Table 1). The activity of the double mutant
381
(D65P/N66G) was also higher than that of WT at 65°C (Fig. 3). In the region, Asn66 was
382
observed in an allowed but not preferred region in the Ramachandran plot. Therefore, it was
383
assumed that a stable conformation in the turn structure was formed due to the combination of
384
the introductions (Pro65 and Gly66), thus enhancing the enzyme’s thermostability. In addition,
385
the result that the value of ∆H by DSC was increased by the mutation (Table 1) suggests that the
386
improved thermostability is also derived by the C-C contact of the introduced Pro. From the
387
theoretical analysis of the ß-turn I structure, however, the Pro introduction seemed to be
388
preferable at the position of Asn66 for its stability (47). The detailed analysis and experiment
389
for this position are in progress. The loop region containing Thr101 and Ser102 does not form a
390
typical ß-turn. From the crystal structural analysis for XylCmt9, it is elucidated that the
391
substitution of Thr101 with a Pro residue contributes to the C-C contact to the side chain of
392
Phe99 and the formation of the main chain in the neighboring ß-strand (B4) (Fig. 6C).
393
Therefore, T101P or S102, or both mutations may have induced the conformational change in
394
the turn region (between A5 and B5), located in an area accessible to the surface, making it
395
possible to couple two hydrogen bonds. These structural features are in agreement with the
396
results of DSC experiments and temperature dependence experiments, in which thermostability
397
increased slightly (Table 1). Interestingly, the five ß-turns observed in WT contain no Pro
398
residues. Therefore, we have a chance to improve its thermostability more by introducing Pro at
399
the sites. 19 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
400
By introducing a disulfide bond between Ser35 and Thr62 in TcXylC (via
401
S35C/T62C substitutions), we constructed a mutant with greater thermostability (Fig. 3, 4). The
402
effect of introducing a disulfide bond in the flexible N-terminal region of AoXyn11A was
403
discussed in a recent report (27). Based on molecular dynamics simulation, it has been
404
estimated that one disulfide bond stabilizes the structure and increases the thermostability of
405
5°C. In the present study, however, we were able to increase the optimum temperature by more
406
than 10°C relative to the WT enzyme by introducing one disulfide bond. We have no
407
information for the structure of the S35C/T62C mutant. However, the positions of the two
408
cysteines (Cys35-Cys62) selected were more suitable sites for stabilizing the enzyme. The
409
value of ∆H by DSC was not improved by the disulfide bond (Table 1). Therefore, the effect of
410
the thermostability seems to be controlled by the entropic effect of the disulfide bond
411
introduced.
412
The crystal structural analysis for XylCmt9 showed not only the formation of a
413
disulfide bond between Cys35 and Cys62, but also that a newly induced C-C contact contribute
414
to the thermostability of the enzyme (Fig. 6A). The main chain Cα of Ile36 moved significantly
415
towards the interior of the protein (1.31 Å) and the side chain flipped toward Leu63 resulting
416
conformational changes allowed a new C-C contact to occur between the Cδ1 of Ile36 and the
417
Cδ1 of Leu63 (Fig. 6A). Furthermore, it was also possible that a new contact formed between
418
Met61 Cε and Leu63 Cδ2 (Fig. 6A). Thus, it was elucidated that the formation of a new
419
hydrophobic contact was triggered by the introduction of the disulfide bond. In the N44H
420
mutant, a new hydrogen bond between the side chain of His44 and the Nδ1 and O atoms of the 20 ACS Paragon Plus Environment
Page 20 of 47
Page 21 of 47
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
421
Gly64 main chain was observed (Fig. 6B). Table 1 shows that the values of ∆H by DSC for
422
Y61M/N63L and D65P/N66G mutants were increased (+22 and +63 kJ/mol) compared with
423
WT. On the other hand, the values of ∆H by DSC for T101P, S102N and N44H mutants were
424
decreased (-45, -37 and -85 kJ/mol). Therefore, the increased value (+87 kJ/mol) for ∆H and
425
value (+0.5 kJ/K mol) for ∆S for XylCmt9 also supports that the formation of new hydrogen
426
bonding and C-C contacts enhanced by the introduced disulfide bond contribute to
427
thermostability. The positive effects of the mutations seem to be partially additive (54). The
428
study mentioned when the increases in activity by a double mutant are greater than the
429
individual increases observed for each single mutation attributes to a cooperative interaction
430
between the two residues mutated (54). Similar phenomena were observed in our study.
431
In this study, we not only enhanced the thermostability of TcXylC to 80.5°C by
432
site-directed mutagenesis, also elucidated the structural basis of thermostability. Currently, we
433
are evaluating the application of thermostable XylCmt9 to the degradation of actual biomass
434
sources, such as bagasse and rice husks.
435 436
AUTHOR INFORMATION
437
Corresponding Author
438
Tel.: +81-82-420-8258, Fax: +81-82-420-7820, E-mail:
[email protected] 439 440
Funding
441
This work was supported by the Japan Science and Technology Agency (JST)-Advanced Low
442
Carbon Technology Research and Development Program (ALCA) (DGA20264004). 21 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
443 444
ACKNOWLEDGEMENTS
445
X-ray diffraction data were obtained using the beam line BL44XU in SPring-8, Hyogo, Japan,
446
with approval of the Institute for Protein Research, Osaka University, Osaka, Japan (proposal
447
No. 2014A6903 and 2014B6903). This work was supported by Toray Industries Inc. New
448
Frontiers Research Laboratories, (Kanagawa, Japan). We would like to thank Dr. Han-woo
449
Kim and Ms. Yuka Maeno a former member at National Institute of Advanced Industrial
450
Science and Technology (AIST) for the preliminary experiment of constructing thermostable
451
xylanases. K. Ishikawa thanks Dr. Wing Sung of National Research Council Canada and Dr.
452
Jeffrey Tolan of Iogen Co. (Ottawa, ON Canada) for giving him a chance to participate in the
453
pulp breaching project using xylanase supported by G7 summit fellowship program in Canada.
454 455
ABBREVIATIONS USED
456
GH, glycoside hydrolase; TcXylC, xylanase 11C from T. cellulolyticus; DNS, dinitrosalicylic
457
acid; DSC, differential scanning calorimetry; RMSD, root mean square deviation.
458 459
REFERENCES
460
1.
461 462
Kumar, R., Singh, S., and Singh, O. V. (2008) Bioconversion of lignocellulosic biomass: biochemical and molecular perspectives. J. Ind. Microbiol. Biotechnol. 35, 377-391.
2.
Ragauskas, A. J., Williams, C. K., Davison, B. H., Britovsek, G., Cairney, J., Eckert, C. A.,
463
Frederick, W. J. Jr., Hallett, J. P., Leak, D. J., Liotta, C. L., Mielenz, J. R., Murphy, R.,
464
Templer, R., and Tschaplinski, T. (2006) The path forward for biofuels and biomaterials.
465
Science 311, 484-489.
466 467
3.
Sanchez, C. (2009) Lignocellulosic residues: biodegradation and bioconversion by fungi. Biotechnol. Adv. 27, 185-194. 22 ACS Paragon Plus Environment
Page 22 of 47
Page 23 of 47
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
468
Biochemistry
4.
469 470
latest developments in xylan research. Biotechnol. Adv. 30, 1627-1640. 5.
471 472
Deutschmann, R., and Dekker, R. F. (2012) From plant biomass to bio-based chemicals:
York, W. S., and O’Neill, M. A. (2008) Biochemical control of xylan biosynthesis – which end is up?. Curr. Opin. Plant. Biol. 11, 258-265.
6.
Coughlan, M. P., and Hazlewood, G. P. (1993) Beta-1,4-D-xylan-degrading
473
enzyme-systems - biochemistry, molecular-biology and applications. Biotechnol. Appl.
474
Biochem. 17, 259-289.
475
7.
Carpita, N. C., and Gibeaut, D. M. (1993) Structural models of primary-cell walls in
476
flowering plants - consistency of molecular-structure with the physical-properties of the
477
walls during growth. Plant J. 3, 1-30.
478
8.
479 480 481 482 483 484 485
van den Brink, J., and de Vries, R. P. (2011) Fungal enzyme sets for plant polysaccharide degradation. Appl. Microbiol. Biotechnol. 91, 1477-1492.
9.
Collins, T., Gerday, C., and Feller, G. (2005) Xylanases, xylanase families and extremophilic xylanases. FEMS Microbiol. Rev. 29, 3-23.
10. Prade, R. A. (1996) Xylanases: from biology to biotechnology. Biotechnol. Genet. Eng. Rev. 13, 101-131. 11. Coutinho, P. M., and Henrissat, B. (1999) Carbohydrate-active enzymes: an integrated database approach. Recent Adv. Carbohydr. Bioeng. 246, 3-12.
486
12. Beaugrand, J., Chambat, G., Wong, V. W. K., Goubet, F., Remond, C., Paës, G.,
487
Benamrouche, S., Debeire, P., O'Donohue, M., and Chabbert, B. (2004) Impact and
488
efficiency of GH10 and GH11 thermostable endoxylanases on wheat bran and
489
alkali-extractable arabinoxylans. Carbohydr. Res. 339, 2529-2540.
490
13. Dimarogona, M., Topakas, E., Christakopoulos, P., and Chrysina, E. D. (2012) The
491
structure of a GH10 xylanase from Fusarium oxysporum reveals the presence of an
23 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
Page 24 of 47
492
extended loop on top of the catalytic cleft. Acta Crystallogr., Sect. D: Biol. Crystallogr.
493
68, 735-742.
494
14. Harris, G. W., Jenkins, J. A., Connerton, I., Cummings, N., Loleggio, L., Scott, M.,
495
Hazlewood, G. P., Laurie, J. I., Gilbert, H. J., and Pickersgill, R. W. (1994) Structure of the
496
catalytic core of the family F-xylanase from Pseudomonas fluorescens and identification
497
of the xylopentaose-binding sites. Structure 2, 1107-1116.
498
15. Lo Leggio, L., Kalogiannis, S., Bhat, M. K., and Pickersgill, R. W. (1999) High resolution
499
structure and sequence of T. aurantiacus xylanase I: implications for the evolution of
500
thermostability in family 10 xylanases and enzymes with (beta)alpha-barrel architecture.
501
Proteins 36, 295-306.
502
16. Paës,
G.,
Berrin,
J.
G.,
and
Beaugrand,
J.
(2012)
GH11
xylanases:
503
structure/function/properties relationships and applications. Biotechnol. Adv. 30, 564-592.
504
17. Sidhu, G., Withers, S. G., Nguyen, N. T., McIntosh, L. P., Ziser, L., and Brayer, G. D.
505
(1999) Sugar ring distortion in the glycosyl-enzyme intermediate of a family G/11
506
xylanase. Biochemistry (Mosc) 38, 5346-5354.
507
18. Wan, Q., Zhang, Q., Hamilton-Brehm, S., Weiss, K., Mustyakimov, M., Coates, L.,
508
Langan, P., Graham, D., and Kovalevsky, A. (2014) X-ray crystallographic studies of
509
family 11 xylanase Michaelis and product complexes: implications for the catalytic
510
mechanism. Acta Crystallogr., Sect. D: Biol. Crystallogr. 70, 11-23.
511
19. Fujii, T., Hoshino, T., Inoue, H., and Yano, S. (2014) Taxonomic revision of the
512
cellulose-degrading fungus Acremonium cellulolyticus nomen nudum to Talaromyces
513
based on phylogenetic analysis. FEMS Microbiol. Lett. 351, 32-41.
514
20. Yamanobe, T., Mitsuishi, Y., and Takasaki, Y. (1987) Isolation of a cellulolytic enzyme
515
producing microorganism, culture conditions and some properties of the enzymes. Agric.
516
Biol. Chem. 51, 65-74. 24 ACS Paragon Plus Environment
Page 25 of 47
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
517
21. Fujii, T., Fang, X., Inoue, H., Murakami, K., and Sawayama, S. (2009) Enzymatic
518
hydrolyzing performance of Acremonium cellulolyticus and Trichoderma reesei against
519
three lignocellulosic materials. Biotechnol. Biofuels 2, 24.
520
22. Watanabe, M., Inoue, H., Inoue, B., Yoshimi, M., Fujii, T., and Ishikawa, K. (2014)
521
Xylanase (GH11) from Acremonium cellulolyticus: homologous expression and
522
characterization. AMB Express 4, 27.
523
23. Kataoka, M., Akita, F., Maeno, Y., Inoue, B., Inoue, H., and Ishikawa, K. (2014) Crystal
524
Structure of Talaromyces cellulolyticus (formerly known as Acremonium cellulolyticus)
525
GH Family 11 xylanase. Appl. Biochem. Biotechnol. 174, 1599-1612.
526
24. Törrönen, A., Harkki, A., and Rouvinen, J. (1994) Three-dimensional structure of
527
endo-1,4-ß-xylanase II from Trichoderma reesei: two conformational states in the active
528
site. EMBO J. 13, 2493-2501.
529
25. Vardakou, M., Dumon, C., Murray, J. W., Christakopoulos, P., Weiner, D. P., Juge, N.,
530
Lewis, R. J., Gilbert, H. J., and Flint, J. E. (2008) Understanding the structural basis for
531
substrate and inhibitor recognition in eukaryotic GH11 xylanases. J. Mol. Biol. 375,
532
1293-1305.
533
26. Wakarchuk, W. W., Campbell, R. L., Sung, W. L., Davoodi, J., and Yaguchi, M. (1994)
534
Mutational and crystallographic analyses of the active site residues of the Bacillus
535
circulans xylanase. Protein Sci. 3, 467-475.
536
27. Chen, Z., Zhang, H., Wang, J., Tang, C., Wu, J., and Wu, M. (2013) Engineering the
537
thermostability of a xylanase from Aspergillus oryzae by an enhancement of the
538
interactions between the N-terminus extension and the ß-sheet A2 of the enzyme.
539
Biotechnol. Lett. 35, 2073-2079.
540
28. Cheng, Y. S., Chen, C. C., Huang, C. H., Ko, T. P., Luo, W., Huang, J. W., Liu, J. R., and
541
Guo, R. T. (2014) Structural analysis of a glycoside hydrolase family 11 xylanase from 25 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
542
Neocallimastix patriciarum: insights into the molecular basis of a thermophilic enzyme. J.
543
Biol. Chem. 289, 11020-11028.
544
29. Georis, J., de Lemos Esteves, F., Lamotte-Brasseur, J., Bougnet, V., Devreese, B.,
545
Giannotta, F., Granier, B., and Frère, J. M. (2000) An additional aromatic interaction
546
improves the thermostability and thermophilicity of a mesophilic family 11 xylanase:
547
structural basis and molecular study. Protein Sci. 9, 466-475.
548
30. Wakarchuk, W. W., Sung, W. L., Campbell, R. L., Cunningham, A., Watson, D. C., and
549
Yaguchi, M. (1994) Thermostabilization of Bacillus circulans xylanase by the
550
introduction of disulfide bonds. Protein Eng. 7, 1379-1386.
551
31. Smith, P. K., Krohn, R. I., Hermanson, G. T., Mallia, A. K., Gartner, F. H., Provenzano, M.
552
D., Fujimoto, E. K., Goeke, N. M., Olson, B. J., and Klenk, D. C. (1985) Measurement of
553
protein using bicinchoninic acid. Anal. Biochem. 150, 76-85.
554 555 556 557
32. Bailey, M. J., Biely, P., and Poutanen, K. (1992) Interlaboratory testing of methods for assay of xylanase activity. J. Biotechnol. 23, 257-270. 33. McIlvaine, T. C. (1921) A buffer solution for colorimetric comparison. J. Biol. Chem. 49, 183-186.
558
34. Sakoda, M., and Hiromi, K. (1976) Determination of the best-fit values of kinetic
559
parameters of the Michaelis-Menten equation by the method of least squares with the
560
Taylor expansion. J. biochem. 80, 547-555.
561 562
35. Otwinowski, Z., and Minor, W. (1997) Processing of X-ray diffraction data collected in oscillation mode. Methods Enzymol. 276, 307-326.
563
36. McCoy, A. J., Grosse-Kunstieve, R. W., Adams, P. D., Winn, M. D., Storoni, L. C., and
564
Read, R. J. (2007) Phaser crystallographic software. J. Appl. Crystallogr. 40, 658-674
565
37. Emsley, P., and Cowtan, K. (2004) Coot: model-building tools for molecular graphics.
566
Acta Crystallogr., Sect. D: Biol. Crystallogr. 60, 2126-2132. 26 ACS Paragon Plus Environment
Page 26 of 47
Page 27 of 47
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
567
38. Brünger, A. T., Adams, P. D., Clore, G. M., DeLano, W. L., Gros, P., Grosse-Kunstleve, R.
568
W., Jiang, J. S., Kuszewski, J., Nilges, M., Pannu, N. S., Read, R. J., Rice, L. M.,
569
Simonson, T., and Warren, G. L. (1998) Crystallography & NMR System: a new software
570
suite for macromolecular structure determination. Acta Crystallogr., Sect. D: Biol.
571
Crystallogr. 54, 905-921.
572
39. Murshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R.
573
A., Winn, M. D., Long, F., and Vagin, A. A. (2011) REFMAC5 for the refinement of
574
macromolecular crystal structures. Acta Crystallogr., Sect. D: Biol. Crystallogr. 67,
575
355-367.
576
40. Gerrit, G. L., Serge, X. C., Victor, S. L., and Anastassis, P. (2008) Automated
577
macromolecular model building for X-ray crystallography using ARP/wARP version 7.
578
Nat Protoc. 3, 1171-1179.
579
41. Laskowski, R. A., MacArthur, M. W., Moss, D. S., and Thornton, J. M. (1993)
580
PROCHECK: a program to check the stereochemical quality of protein strucutres. J. Appl.
581
Crystallogr. 26, 283-291.
582
42. Potterton, L., McNicholas, S, Krissinel, E., Gruber, J., Cowtan, K., Emsley, P., Murshudov,
583
G. N., Cohen, S., Perrakis, A., and Noble, M. (2004) Developments in the CCP4 molecular
584
graphics project. Acta Crystallogr., D: Biol. Crystallogr. 60, 2288-2294.
585
43. Fujii, T., Koike, H., Sawayama, S., Yano, S., and Inoue, H. (2015) Draft Genome
586
Sequence of Talaromyces cellulolyticus Strain Y-94, a Source of Lignocellulosic
587
Biomass-Degrading
588
10.1128/genomeA.00014-15.
Enzymes.
Genome
Announc.
3,
pii:
e00014-15.
doi:
589
44. Ratanakhanokchai, K., Kyu, K. L., and Tanticharoen, M. (1999) Purification and
590
properties of a xylan-binding endoxylanase from Alkaliphilic bacillus sp. strain K-1. Appl.
591
Environ. Microbiol. 65, 694-697. 27 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
592
45. Manikandan, K., Bhardwaj, A., Gupta, N., Lokanath, N. K., Ghosh, A., Reddy, V. S., and
593
Ramakumar, S. (2006) Crystal structures of native and xylosaccharide-bound alkali
594
thermostable xylanase from an alkalophilic Bacillus sp. NG-27: structural insights into
595
alkalophilicity and implications for adaptation to polyextreme conditions. Protein Sci. 8,
596
1951-1960.
597
46. Wouters, J., Georis, J., Engher, D., Vandenhaute, J., Dusart, J., Frere, J. M., Depiereux, E.,
598
and Charlier, P. (2001) Crystallographic analysis of family 11 endo-beta-1,4-xylanase
599
Xyl1 from Streptomyces sp. S38. Acta Crystallogr., Sect. D: Biol. Crystallogr. 57,
600
1813-1819.
601
47. Matthews, B. W., Nicholson, H., and Becktel, W. J. (1987) Enhanced protein
602
thermostability from site-directed mutations that decrease the entropy of unfolding. Proc.
603
Natl. Acad. Sci. U. S. A. 84, 6663-6667.
604
48. Trevino, S. R., Schaefer, S., Scholtz, J. M., and Pace, C. N. (2007) Increasing protein
605
conformational stability by optimizing beta-turn sequence. J. Mol. Biol. 373, 211-218.
606
49. Chou, P. Y., and Fasman, G. D. (1974) Conformational parameters for amino acids in
607
helical, beta-sheet, and random coil regions calculated from proteins. Biochemistry 13,
608
211-222.
609
50. Matthews, B. W. (1968) Solvent content of protein crystals. J. Mol. Biol. 33, 491-497.
610
51. Ramachandran, G. N., and Sasiskharan, V. (1968) Conformation of polypeptides and
611 612 613
proteins. Adv. Protein Chem. 23, 283-438. 52. Schomburg, I., Chang, A., and Schomburg, D. (2002) BRENDA, enzyme data and metabolic information. Nucl. Acids Res. 30, 47-49.
614
53. Schellman, J. A. (1975) Macromolecular binding. Biopolymers 14, 999-1018.
615
54. Mildvan, A. S., Weber, D. J., and Kuliopulos, A. (1992) Quantitative interpretations of
616
double mutations of enzymes. Arch. Biochem. Biophys. 294, 327-340. 28 ACS Paragon Plus Environment
Page 28 of 47
Page 29 of 47
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
617
55. Thompson, J. D., Higgins, D. G., and Gibson, T. J. (1994) CLUSTAL W: improving the
618
sensitivity of progressive multiple sequence alignment through sequence weighting,
619
position-specific gap penalties and weight matrix choice. Nucl. Acids Res. 22, 4673-4680.
620
56. Gouet, P., Robert, X., and Courcelle, E. (2003) ESPript/ENDscript: extracting and
621
rendering sequence and 3D information from atomic structures of proteins. Nucl. Acids
622
Res. 31, 3320-3323.
623
29 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
624
FIGURE LEGENDS
625
Figure 1. Structure-based sequence alignments of various GH11 xylanases. Strictly conserved
626
residues are boxed in red. Similar residues are shown in black bold characters and boxed in
627
yellow. The catalytic residues are indicated with blue boxes. One α-helix and 14 ß-strands
628
(A2-6, B1-9) in TcXylC are representative as a black coil and black arrows, respectively. The
629
species producing the various xylanases are abbreviated as follows: TcXylC: GH11 from
630
Talaromyces cellulolyticus (GenBank Accession Number BAO51921), TrXyl: GH11 from
631
Trichoderma reesei (GenBank Accession Number ACB38137), ScXyl: GH11 from
632
Schizophyllum commune (GenBank Accession Number AAB29056), SoXyl: GH11 from
633
Streptomyces olivaceoviridis (GenBank Accession Number AHK22787), SspXyl: GH11 from
634
Streptomyces sp. S38 (GenBank Accession Number CAA67143), BcXyl: GH11 from Bacillus
635
circulans (GenBank Accession Number P09850), AnXyl: GH11 from Aspergillus niger
636
(GenBank Accession Number AAS46913), and NpXyl: GH11 from Neocallimastix
637
patriciarum (GenBank Accession Number 3WP4_A). The figure was constructed using
638
ClustalW (55) and ESPript (56).
639 640
Figure 2. Mutations mapped on the closed form of WT_TcXylC. A is colored with increased
641
the Tm by 0-5 degrees (yellow; S102), by 5-10 degrees (cyan; D65/N66), by 10-15 degrees (red;
642
S35/T62). B is colored with increased activity by single mutation sites (yellow; S35, N44, T62,
643
T101 and S102), by double mutation sites (cyan; Y61/N63 and D65/N66). Each residue was
644
drawn on the WT_TcXylC structure (colored white).
645 646
Figure 3. Relative activities of WT and mutants. Optimum temperature of all proteins (1.0
647
mg/mL) was measured at temperatures ranging from 45 to 60-75°C using the DNS method in
648
50 mM sodium acetate (pH 5.0). Samples were pre-heated for 10 min at the indicated 30 ACS Paragon Plus Environment
Page 30 of 47
Page 31 of 47
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
649
temperature before mixing with the substrate. Each sample is shown as follows (WT: white
650
squares with dashed line; S35C: open diamonds; N44H: white squares; Y61M: black bars;
651
T62C: open triangles; N63L: black crosses; D65P: black squares with dashed line; N66G: black
652
triangles with dashed line; T101P: black crosses; S102N: black squares; S35C/T62C: black
653
triangles; Y61M/N63L: black diamonds; D65P/N66G: black circles; XylCmt9: open circles).
654
In both figures, the percent relative activity is based on the specific activity (U/mg) of XylCmt9
655
at 75°C. The data plotted are averages of three experiments.
656 657
Figure 4. DSC results. The DSC curves for nine point mutations and a double mutation are
658
shown (S35C: brown; N44H: red; Y61M: green; T62C: blue; N63L: pink; D65P: purple;
659
N66G: orchid; T101P: cyan; S102N: dark green; S35C/T62C: navy; Y61M/N63L: yellow and
660
D65P/N66G: gold). WT and XylCmt9 are also shown as black and gray dotted lines,
661
respectively. All samples were adjusted to 1.0 mg/mL and dialyzed against 50 mM sodium
662
acetate buffer (pH 5.5) before the DSC measurements. Bar represents heat capacity (1 J K-1 g-1).
663 664
Figure 5. Superimposed models of 12 molecules of XylCmt9, and XylCmt9 (chain A) and WT
665
(closed form). A, 12 molecules (chain A-L, PDB: 5HXV) of XylCmt9 are indicated worm
666
models and colored by red, blue, pink, purple, yellow, light blue, cyan, green, coral, magenta,
667
gold, and tan, respectively. B, XylCmt9 (cyan) and WT (pink) are shown as ribbon models with
668
labeled ß-strands (B1-9/A2-6), one α-helix and thumb. Two glutamate residues (Glu119 and
669
Glu210) are indicated in the cylinder model. N- and C-termini are labeled.
670 671
Figure 6. New interactions in XylCmt9. A, Stereo view showing an overlay of the disulfide
672
bond and C-C contact with superimposed Ile36 of the WT enzyme. S35C, Ile36, T62C, and
673
N63L of XylCmt9 (italic) are shown as a cylinder model with atomic element colors (O atoms: 31 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
674
red; N atoms: blue; C atoms: cyan) and the disulfide bond in yellow. WT Ile36 is also shown in
675
the cylinder model with atomic element colors (O atoms: red; N atoms: blue; C atoms: pink).
676
The side chains of Ile36 and N63L of XylCmt9 are also highlighted cyan in the cylinder model
677
and the C-C contact is indicated with a black dashed line. Ile36, Ser35, Asn62, and Asn63 of
678
WT are colored pink in the cylinder model. B, Stereo view showing an overlay of a hydrogen
679
bond. N44H and Gly64 of XylCmt9 are shown in the cylinder model with atomic element
680
colors. The hydrogen bond between N44H and Gly64 is indicated by black dashed lines. C,
681
Stereo view showing an overlay of two hydrogen bonds. T101P, S102N, and Ser215 of
682
XylCmt9 are shown in the cylinder model with atomic element colors. The two hydrogen bonds
683
between S102N and Ser215 are indicated by black dashed lines.
684 685 686
SUPPLEMENTAL Fig. 1 SDS-PAGE analysis of purified WT and XylCmt9. M: molecular
687
weight markers; lane 1: WT; lane 2: XylCmt9. Both proteins are the final product after
688
purification using a Superdex 200 16/60 gel filtration column (GE Healthcare).
689 690
SUPPLEMENTAL Table 1 Summary of nucleotide primers used in this study.
32 ACS Paragon Plus Environment
Page 32 of 47
Page 33 of 47
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
691
Table 1 Summary of the relative activities and thermodynamic parameters for the heat
692
denaturation of TcXylC and mutants Relative activity (45°C) %
693 694 695 696 697
Tm o
Tm* (+X2)*
C
∆H(Tm)
∆S(Tm)
∆H(58.9°C)**
∆S(58.9°C)**
kJ/mol
kJ/(K mol)
kJ/mol
kJ/(K mol)
WT
100 ± 3
58.9
(59.8)
461
1.39
461
1.38
Y61M
94 ± 3
53.5
(54.1)
390
1.19
398
1.22
N63L
85 ± 3
50.6
(51.5)
349
1.08
362
1.12
Y61M/N63L
109 ± 3
56.7
(59.0)
483
1.55
486
1.56
n.d.
n.d.
n.d.
***
D65P
98 ± 3
50.0, 58.1
(58)
226
N66G
100 ± 4
49.5, 57.8
(58)
309
n.d.
n.d.
n.d.
D65P/N66G
102 ± 2
64.6
(65.4)
524
1.55
515
1.52
S35C
101 ± 2
56.6
(57.3)
413
1.25
417
1.26
T62C
86 ± 2
58.3
(57.9)
376
1.13
377
1.13
S35C/T62C
97 ± 3
74.5
(75.1)
469
1.35
445
1.28
T101P
99 ± 2
58.8
(60.5)
416
1.25
416
1.25
S102N
100 ± 2
60.0
(60.8)
414
1.24
412
1.23
N44H
101 ± 4
56.4
(57.3)
376
1.14
380
1.15
XylCmt9
114 ± 3
80.5
(81.7)
548
1.55
514
1.45
* Tm with 36 mM X2 **∆H(58.9oC) and ∆S(58.9oC) are the enthalpy and entropy changes at Tm of WT , respectively calculated by using the averaged heat capacity change ∆Cp=1.5 kJ/(K mol) and the equations of of ∂∆H/∂T= ∆Cp and ∂∆S /∂T= ∆Cp /T. *** not determined.
698 699 700 701 702 703 704 705 33 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
706
Table 2 Data collection and refinement statistics Data collection
XylCmt9
Wavelength (Å)
0.9
Space group
P21
Unit cell: a, b, c (Å), ß (°)
a=75.9, b=179.8, c=93.3, ß=103.8
Matthews coefficient (Å3Da-1)
2.5
Solvent content (%)
50.9
Resolution (Å)
50.0-2.0 (2.03-2.00)*
Rmerge a (%)
13.0 (28.0)*
Average I/σ(I)
8.6 (2.8)*
Completeness (%)
94.2 (88.2)*
Redundancy
2.7 (2.2)*
No. molecules/asym
12
Refinement Resolution (Å)
20.0-2.0
No. reflections (test reflections)
145,667 (7,697)
Rwork b/Rfree c (%)
18.0/22.8
No. atoms
18,554
Protein
17,352
Water
1,202
Mean overall B factor (Å2)
11.2
Root mean square deviations Bond lengths (Å)
0.01
Bond angles (°)
1.88
34 ACS Paragon Plus Environment
Page 34 of 47
Page 35 of 47
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
Ramachandran plot In most favored regions (%)
94.8
In allowed regions (%)
4.4
PDB accession #
5HXV
707
*
708
a
709
measurement of reflection hkl, including symmetry-related reflections, and is their
710
average. b Rwork = ΣhΣi ||Fo| – |Fc|| / Σ |Fo|. c Rfree is Rwork for approximately 5 % of the reflections
711
that were excluded from the refinement.
Outer shell (2.03-2.00 Å) Rmerge = ΣhklΣi | Ii(hkl) – | Σ hklΣiIi(hkl), where Ii(hkl) is the intensity of the ith
712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 35 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
728
Page 36 of 47
Table 3 Kinetic parameters of WT and XylCmt9 Enzyme (50°C)
Specific activity (U/mg)
Km (%)
kcat (sec−1)
kcat / Km (%-1 sec-1)
WT
3,991
4.7 ± 0.1
35.3 ± 0.3
7.5
XylCmt9
4,659
7.1 ± 0.1
49.8 ± 0.2
7.0
729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 36 ACS Paragon Plus Environment
Page 37 of 47
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
750
For Table of Contents Use Only
751
Manuscript Title: Construction of thermophilic xylanase and its structural analysis
752
Authors: Masahiro Watanabe, Harumi Fukada and Kazuhiko Ishikawa
753
754 755 756 757 758 759 760 761 762 763 764 765 766 767 768
37 ACS Paragon Plus Environment
Biochemistry
Page 38 of 47 B1
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
A2
A3
B5
B3
A5
B6
B7
B9
A6
B2
α-helix
B8
B4
A4
ACS Paragon Plus Environment
Fig. 1 Watanabe et al.
Page 39 of 47
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
Biochemistry
D65
T62
S35
S102
N66 N44
N63
Y61 T101
ACS Paragon Plus Environment
Fig. 2A Watanabe et al.
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
D65
T62
Page 40 of 47
S35
S102
N66
ACS Paragon Plus Environment
Fig. 2B Watanabe et al.
Page 41 of 47
WT
100
S35C N44H 80
Y61M T62C
Relative activity (%)
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
Biochemistry
N63L
60
D65P N66G 40
T101P S102N S35C/T62C
20
Y61M/N63L D65P/N66G 0 45
50
55
60
65
70
75
XylCmt9
Temperature (℃)
ACS Paragon Plus Environment
Fig. 3 Watanabe et al.
Biochemistry
1 J/(K g )
XylCmt9 S35C/T62C S102N WT
Heat Capacity
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
Page 42 of 47
T101P T62C S35C N44H Y61M Y61M/N63L N63L N66G D65P
30
40
50
60
70
D65P/N66G
80
90
o
Temperature / C
ACS Paragon Plus Environment
Fig. 4 Watanabe et al.
Page 43 of 47
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
Biochemistry
ACS Paragon Plus Environment
Fig. 5A Watanabe et al.
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
N
Page 44 of 47
A2 A3
B1 B2
C B3
Thumb
A5
A4
A6
B4 Glu210 B5 Glu119
B6 B7 B8
α-helix
ACS Paragon Plus Environment
B9
Fig. 5B Watanabe et al.
Page 45 of 47
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
Biochemistry
Tyr61
Y61M
Tyr61
Y61M
3.6
3.6 Asn63
Ile36
Asn63 Ile36
3.7 N63L
Ile36
T62C
3.7 N63L
Ile36
T62C
Thr62
S35C (N-ter.)
Thr62
S35C (N-ter.)
Ser35
Ser35
ACS Paragon Plus Environment
Fig. 6A Watanabe et al.
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
Page 46 of 47
Gly64
3.2
Gly64
3.2
Gly64
Asn44
Gly64
Asn44
N44H
N44H
ACS Paragon Plus Environment
Fig. 6B Watanabe et al.
Page 47 of 47
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
Biochemistry
Ser102
S102N
Ser215
3.3
Ser102
S102N
Ser215
Thr101
3.3
Thr101 2.8
2.8 Ser215
Ser215 T101P
ACS Paragon Plus Environment
T101P
Fig. 6C Watanabe et al.