Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)
Article
Directed evolution of carbonyl reductase from Rhodosporidium toruloides and its application in stereoselective synthesis of tert-butyl (3R,5S)-6-chloro-3,5-dihydroxyhexanoate Zhi-Qiang Liu, Lin Wu, Xiao-Jian Zhang, Ya-Ping Xue, and Yu-Guo Zheng J. Agric. Food Chem., Just Accepted Manuscript • Publication Date (Web): 20 Apr 2017 Downloaded from http://pubs.acs.org on April 23, 2017
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Journal of Agricultural and Food Chemistry 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 42
Journal of Agricultural and Food Chemistry
Title: Directed evolution of carbonyl reductase from Rhodosporidium toruloides and its application
in
stereoselective
synthesis
of
tert-butyl
(3R,5S)-6-chloro-
3,5-dihydroxyhexanoate
Authors and Affiliation: Zhi-Qiang Liua,b, Lin Wua,b, Xiao-Jian Zhang a,b, Ya-Ping Xue a,b and Yu-Guo Zheng a,b * a
Key Laboratory of Bioorganic Synthesis of Zhejiang Province, College of Biotechnology and
Bioengineering, Zhejiang University of Technology, Hangzhou 310014, China b
Engineering Research Center of Bioconversion and Biopurification of Ministry of Education,
Zhejiang University of Technology, Hangzhou 310014, China
*Corresponding author: Fax: +86-571-88320630; Tel: +86-571-88320630; E-mail:
[email protected] 1
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 2 of 42
1
ABSTRACT: tert-Butyl (3R,5S)-6-chloro-3,5-dihydroxyhexanoate ((3R,5S)-CDHH)
2
is a key intermediate of atorvastatin and rosuvastatin synthesis. Carbonyl reductase
3
RtSCR9 from Rhodosporidium toruloides exhibited excellent activity toward
4
tert-butyl (S)-6-chloro-3-hydoxy-5-oxohexanoate ((S)-CHOH). To improve the
5
activity of RtSCR9, random mutagenesis and site-saturation mutagenesis were
6
performed. Three positive mutants were obtained (mut-Gln95Asp, mut-Ile144Lys and
7
mut-Phe156Gln). These mutants exhibited 1.94-, 3.03- and 1.61-fold, and 1.93-, 3.15-
8
and 1.97-fold improvement on the specific activity and kcat/Km, respectively.
9
Asymmetric reduction of (S)-CHOH by mut-Ile144Lys coupled with glucose
10
dehydrogenase (GDH) was conducted. The yield and enantiomeric excess of
11
(3R,5S)-CDHH respectively reached 98% and 99% after 8-h bioconversion in a single
12
batch reaction with 1 M (S)-CHOH, and the space-time yield reached 542.83 mmol
13
L-1 h-1 g-1 wet cell weight. This study presents a new carbonyl reductase for efficient
14
synthesis of (3R,5S)-CDHH.
15
KEYWORDS:
16
reductase, random mutagenesis, site-saturation mutagenesis, single batch reaction
tert-butyl
(3R,5S)-6-chloro-3,5-dihydroxyhexanoate,
2
ACS Paragon Plus Environment
carbonyl
Page 3 of 42
Journal of Agricultural and Food Chemistry
17
INTRODUCTION
18
tert-Butyl (3R,5S)-6-chloro-3,5-dihydroxyhexanoate ((3R,5S)-CDHH) is a
19
promising intermediate applied in the side chain synthesis of atorvastatin and
20
rosuvastatin, cholesterol regulation drugs used in preventing cardiovascular diseases,
21
such as hypercholesterolemia, atherosclerosis and coronary heart disease.1-3 Chemical
22
synthesis of (3R,5S)-CDHH typically starts with (S)-epichlorohydrin followed by a
23
series of catalytic reactions involving expensive metal catalysts and environmentally
24
unfriendly organic solvents. To introduce the second chiral center in the chemical
25
synthesis of (3R,5S)-CDHH, sodium borohydride (NaBH4) is used to reduce the
26
carbonyl at C-3 under Parasad’s conditions (Et2BOMe, -80°C to -75°C).4,5 Although
27
(3R,5S)-CDHH can be totally synthesized within a short processing period by
28
chemical synthesis, difficulties remain in producing a final product with sufficient
29
enantiomeric purity. In addition, complex protection-deprotection processes, low
30
overall yield and high environmental pressure have made chemical synthesis less
31
competitive.6 Compared to chemical synthesis, biosynthesis has been an alternative in
32
the preparation of (3R,5S)-CDHH, along with the merits of low catalyst cost, high
33
enantioselectivities and a broad functional group tolerability, which is dominant in
34
large-scale pharmaceutical manufacture of important chiral intermediates.7-9
35
Carbonyl reductases (EC 1.1.1.148) have been demonstrated to be valuable
36
biocatalysts in the reduction of ketones to enantiopure alcohols with advantages
37
including mild catalytic conditions, chemo-, regio- and stereo-selectivity, and no
38
heavy metal contamination. So far, considerable studies have been performed on 3
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 4 of 42
39
biotransformation of tert-butyl 6-chloro-3,5-dioxohexanoate (CDOH) and tert-butyl
40
(S)-6-chloro-3-hydoxy-5-oxohexanoate ((S)-CHOH) by carbonyl reductases to
41
produce
42
enantioselectivity in the reduction of CDOH at C-5 to produce (S)-CHOH with over
43
99.5% enantiomeric excess (e.e.).11-13 Previously, a two-step biotransformation
44
process was established using LkADH1 and LkADH2 from Lactobacillus kefir in a
45
simple batch process to produce (3R,5S)-CDHH with >99% e.e. and 47.5% yield.14-18
46
Subsequently, fed batch processes and two-phase biotransformation systems were
47
established.19-21 A recombinant carbonyl reductase from Candida magnoliae
48
coexpressed with a glucose dehydrogenase highly efficiently catalyzed the reduction
49
of 200 g/L (S)-CHOH to prepare (3R,5S)-CDHH with 97.2% yield and 98.6%
50
diastereoisomeric excess (d.e.).22,23 Recently, liquid-core immobilized Saccharomyces
51
cerevisiae CGMCC No.2233 was used as catalyst to accomplish 100% conversion of
52
(S)-CHOH to produce (3R,5S)-CDHH with >99% d.e. when initial substrate
53
concentration was less than 50 g/L.24 Enzymatic asymmetric reductions for
54
commercial application have in general been, however, limited to poor substrate
55
tolerance, insufficient stability and low volumetric productivity. Thus, biocatalysts
56
with evolved characteristics are still urgently demanded to meet the large-scale
57
production of (3R,5S)-CDHH.
(3R,5S)-CDHH.10
Lactobacillus
brevis
(LbADH)
showed
a
high
58
Modification of carbonyl reductases is mainly focused on the improvement of
59
catalytic activity, stereoselectivity, stability and cofactor specificity.25-28 Codexis
60
Corporation (Redwood, CA) engineered a ketoreductase gene from Saccharomyces 4
ACS Paragon Plus Environment
Page 5 of 42
Journal of Agricultural and Food Chemistry
61
cerevisiae by rounds of directed evolution, resulting in a 99.8% conversion of 13 g/L
62
substrate (S)-CHOH with 99.2% d.e. in a 4-h reaction compared to wild type
63
ketoreductase with 37% yield, 97.3% d.e. in a 20-h reaction.29 A tetrad mutant of
64
LkADH through rational design was screened, exhibiting a 3.7- and 42-fold
65
improvement in specific activity toward CDOH over LbADH and wild-type LkADH,
66
respectively.30 Recently, the catalytic efficiency of aldo-keto reductase toward
67
tert-butyl 6-cyano-(5R)-hydroxy-3-oxohexanoate was improved 11.25-fold through
68
rational protein modification. The mutant (KIAKR-Tyr295Trp/Trp296Leu) catalyzed
69
the asymmetric reduction with tert-butyl 6-cyano-(3R,5R)-dihydroxyhexanoate
70
accumulated up to 162.7 mM with >99.5% d.e..31
71
In our previous work, a carbonyl reductase RtSCR9 (Genbank accession No.
72
EMS 18622.1) from Rhodosporidium toruloides was screened and showed
73
considerable activity toward (S)-CHOH.32 To improve the catalytic efficiency for
74
industrial application, random mutagenesis and site-saturation mutagenesis were
75
performed. Three positive variants showing 1.94-, 3.03- and 1.61-fold improvement in
76
specific activity were obtained. Apparent kinetic parameters, molecular modeling and
77
docking experiments were used to illustrate the mechanism of the changes in enzyme
78
characteristics. Bioconversion of 1 M (S)-CHOH was conducted in a single batch
79
reaction. The variant mut-Ile144Lys exhibited an enhanced catalytic ability
80
with >98% yield and >99% e.e. within 8 h of reaction, indicating its potential
81
application in upscale production of (3R,5S)-CDHH.
82
MATERIALS AND METHODS 5
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
83
Microorganism, media and chemicals. Escherichia coli BL21(DE3)
84
(Invitrogen, Carlsbad, CA) was used as the host cell and plasmid pET-28a(+)
85
(Novagen, Darmstadt, Germany) was used for cloning and expression. E. coli
86
BL21(DE3)/pET28a-RtSCR9 was used for expression of carbonyl reductase gene
87
RtSCR9. E. coli BL21(DE3)/pET28a-GDH carrying a glucose dehydrogenase (GDH)
88
gene was used for the “enzyme-coupled” cofactor regeneration system.33 E. coli cells
89
were cultured in Luria-Bertani (LB) medium containing 5 g of yeast extract, 10 g of
90
tryptone, and 10 g of NaCl per liter. Kanamycin (Kan) and the inducer
91
isopropylthio-β-galactoside (IPTG) were purchased from Sango Biotech (Shanghai,
92
China). Nicotinamide adenine dimucleotide phosphate (NADPH) was obtained from
93
Roche (Karlsruhe, Germany). Standards of (S)-CHOH and (3R,5S)-CDHH were
94
purchased from J&K Scientific Ltd. (Shanghai, China). All other chemicals were of
95
analytical grade purity and commercially available.
96
Analytical methods. The concentrations of (S)-CHOH and (3R,5S)-CDHH
97
were determined by high-performance liquid chromatography (HPLC) (Shimadzu Co.,
98
Kyoto, Japan) using an Agilent Zorbax SB-C8 column (150×4.6 mm, particle size 5
99
µm, Agilent Technologies Co., Santa Clara, CA) with UV detection at 210 nm. The
100
mobile phase was a mixture of acetonitrile and ultrapure water (30:70 v/v). The
101
retention times of (S)-CHOH and (3R,5S)-CDHH were 6.11 min and 9.66 min,
102
respectively, at 40°C, with a flow rate of 1 mL/min (Figure S1).
103
The e.e. values of (3R,5S)-CDHH were determined on a Chiracel OD-H column
104
(250×4.6 mm, particle size 5 µm, Daicel Chemical Industries, Tokyo, Japan) at 215 6
ACS Paragon Plus Environment
Page 6 of 42
Page 7 of 42
Journal of Agricultural and Food Chemistry
105
nm using hexane and isopropanol (85:15 v/v) as the mobile phase.30 The column
106
conditions were 40°C at a flow rate of 1 mL/min. Samples from the aqueous reaction
107
mixture containing hydroxyl keto ester were extracted into the mobile phase by
108
vigorous mixing. After centrifuging at 12,000×g for 3 min, the organic phase was
109
collected and filtered before HPLC analysis. The retention times of (S)-CHOH and
110
(3R,5S)-CDHH were 5.09 min and 5.90 min, respectively (Figure S2).
111
Construction of mutant library. A random mutagenesis library was
112
established through error-prone PCR and megaprimer PCR. The recombinant plasmid
113
pET28a-RtSCR9 was used as template DNA and the primers for error-prone PCR
114
were
115
5’-TCTACCATGGCAAGAACGTCC-3’. Appropriate amounts of Mg2+ and Mn2+
116
were added to the PCR reaction mixture in order to control the mismatch rate of base
117
pairing. The final error-prone PCR reaction mixture contained MgCl2 1.5 mM, MnCl2
118
0.3 mM, dATP 0.1 mM, dGTP 0.1 mM, dTTP 0.5 mM, dCTP 0.5 mM, 30 pmol each
119
primer, 10 ng template DNA, and 5 U Taq polymerase, diluted to a final volume of 50
120
µL by adding ddH2O. The error-prone PCR reaction condition started at 94°C for 3
121
min followed by amplification for 30 cycles: 95°C 30 s, 55°C 30 s, and 72°C 1 min,
122
followed by elongation at 72°C for 10 min. Subsequently, the target gene with
123
mutations from error-prone PCR was used as megaprimer to create random
124
mutagenesis libraries through megaprimer PCR of the whole plasmid (MEGAWHOP).
125
The 50 µL megaprimer PCR reaction mixture contained 2×PCR buffer 25 µL, 10 mM
126
dNTP 1 µL, 10 ng template DNA, 2 µL megaprimer, 2.5 U Phanta Max super-Fidelity
5’-TATGTCTTCGCCTACTCCCAAC-3’
7
ACS Paragon Plus Environment
and
Journal of Agricultural and Food Chemistry
127
DNA polymerase and 20.5 µL ddH2O following the procedure of denaturation at 98°C
128
for 3 min, 26 cycles for amplification 10 s at 98°C, 5 s at 55°C and 6 min at 72°C,
129
then elongation at 72°C for 10 min. The PCR products were treated with 40 U DpnI at
130
37°C for 3 h to remove the original template plasmid. The treated PCR products were
131
transformed to E. coli BL21(DE3) by heat-shock treatment. Transformants were
132
plated on LB-agar medium supplemented with 50 µg/mL of Kan and incubated for 12
133
h at 37°C. The mutations were determined by DNA sequencing.
134
High throughput screening and HPLC assay. A two-step screening
135
method based on the high-throughput screening combined with HPLC analysis was
136
designed to select positive mutants. Specifically, after transformants were obtained,
137
single colonies were picked into 800 µL LB medium (50 µg/mL Kan) in a well of a
138
96-well plate and incubated for 12 h at 37°C, 150 rpm. Fifty µL from each well of the
139
preculture was transferred to another 96-well plate with each well containing 800 µL
140
LB (50 µg/mL Kan), to which was added 300 µL 30% glycerol. After addition of
141
IPTG (0.1 mM final concentration), the main culture was incubated at 37°C for 3 h
142
(150 rpm) and then cultivated at 28°C for 16 h (150 rpm). Cells were harvested by
143
centrifugation at 1,500×g for 20 min, washed with phosphate buffer, pH 7.0, and
144
stored in a -80°C freezer. The cells were taken through three freeze-thaw cycles for
145
disruption and then treated with 2 g/L lysozyme at 28°C for 2 h. The lysate was
146
collected after centrifugation at 1,500×g for 20 min. The enzyme activity was
147
determined in a 200 µL reaction mixture containing 0.5 mM NADPH and 2 mM
148
(S)-CHOH within a quartz 96-well plate. Fifty µL of the resulting supernatant was 8
ACS Paragon Plus Environment
Page 8 of 42
Page 9 of 42
Journal of Agricultural and Food Chemistry
149
added and the decrease absorbance of optical density at 340 nm (OD340) in 3 min after
150
shaking 3 seconds was monitored on SpectraMax 5 (Molecular Devices, Sunnyvale,
151
CA). Positive mutants with improved activity consumed NADPH faster than negative
152
ones. This high-throughput screening method was used as preliminary step for
153
screening positive mutants, which shortened the screening time greatly.
154
Continually, the positive mutants obtained by high-throughput screening
155
methods were further tested according to the following procedures. Mutants with
156
higher activities during the first round of screening were inoculated into 8 mL LB
157
culture medium (50 µg/mL Kan, 37°C, 150 rpm, 8 h) from the primary seed plate and
158
transferred to 500 mL Erlenmeyer flasks containing 100 mL of LB medium (50
159
µg/mL Kan). IPTG (final concentration 0.1 mM) was added into the culture to induce
160
protein expression at 28°C on a rotary shaker at 150 rpm when OD600 reached 0.6-0.8.
161
After 12-h of induction, the cells were centrifuged at 8,000×g for 10 min and washed
162
with 30 mL phosphate buffer (pH 7.0). The bioconversion reaction system in 10 mL
163
contained 5 g/L variant, 5 g/L GDH, 100 mM (S)-CHOH and 140 mM glucose. The
164
bioconversion reactions were conducted at 30°C, 150 rpm on a water bath shaker for
165
30 min and terminated by adding 1 mL acetonitrile. The reaction mixture was
166
vigorously mixed before parallel samples were extracted and diluted to a proper
167
concentration before HPLC analysis of (3R,5S)-CDHH. Positive mutants were
168
preserved and sequenced.
169
Site-saturation mutagenesis. Amino acid sites 95, 144 and 156 were
170
subjected to site-saturation mutagenesis using the plasmid pET28a-RtSCR9 as the 9
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
171
template in order to determine the best variant at each position. The primers are listed
172
in Table S1. PCR reaction was carried out with Phanta Max super-Fidelity DNA
173
Polymerase in a 50 µL reaction volume containing 0.2 mM of both primers, 10 ng
174
template DNA, 1.0 mM MgCl2, 0.2 mM dNTP, and 2.5 U of Phanta DNA polymerase.
175
Twenty-six cycles of PCR were carried out at 98°C for 10 s, 55°C for 5 s and 72°C
176
for 6 min. The PCR product was digested with DpnI restriction enzyme to remove
177
template plasmid and then transformed into competent cells of E. coli BL21(DE3).
178
The transformants were sequenced to identify the variations at each site. The relative
179
activities of mutants with varied amino acid substitutions at each site were determined
180
in 10 mL reaction mixtures under the same condition as described above.
181
Expression and purification of recombinant carbonyl reductase. E.
182
coli cells were cultured in 50 mL LB medium containing 50 µg/mL Kan for 8 h at
183
37°C on a rotary shaker at 150 rpm, and then IPTG (0.1 mM final concentration) was
184
added into the culture when the OD600 reached 0.6-0.8 to induce the expression of
185
RtSCR9 at 28°C, 150 rpm. After 12-h of induction, the cells were centrifuged at 8,000
186
×g for 10 min, washed with buffer A (50 mM Na2HPO4-NaH2PO4, 300 mM NaCl, pH
187
8.0). The harvested cells were resuspended in buffer A and disrupted by sonication in
188
an ice water bath. The cell debris was removed by centrifugation at 12,000×g for 30
189
min. The supernatant was loaded onto a Nickel-NTA column (Bio-Rad, Hercules, CA)
190
at 1 mL/min, equilibrated with buffer A at 1 mL/min. Nonspecifically bound protein
191
was washed from the column with buffer B (buffer A containing 15 mM imidazole,
192
pH 8.0), and target protein was eluted with buffer C (buffer A containing 400 mM 10
ACS Paragon Plus Environment
Page 10 of 42
Page 11 of 42
Journal of Agricultural and Food Chemistry
193
imidazole, pH 8.0). The collected protein was washed with ddH2O and concentrated
194
with a 10 kDa Millipore filter. The purity and molecular weight were analyzed by
195
sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE).34
196
Enzyme activity assay and kinetic parameters measurement. The
197
enzyme activities of purified recombinant RtSCR9 and variants were measured at
198
30°C in 100 mM phosphate buffer pH 7.0 containing 1 mM NADPH and 5 mM
199
(S)-CHOH. The reaction mixtures were preheated for 3 min at 30°C. The reactions
200
were performed at 30°C for 3 min and terminated by adding 1 mL of 30% aqueous
201
acetonitrile. One mL of the reacting solution was extracted and centrifuged at
202
12,000×g for 3 min. The supernatant was filtered and analyzed by HPLC. One unit (U)
203
of enzyme activity was defined as the amount of enzyme required to catalyze the
204
production of 1 µmol (3R,5S)-CDHH per minute under the standard enzyme activity
205
assay conditions. To determine kinetic parameters, (S)-CHOH was used as substrate,
206
with concentrations ranging from 0 to 10 mM, and the concentration of the cofactor
207
NADPH was varied from 0.5-2.5 mM. The maximal reaction rate (Vmax) and apparent
208
Michaelis-Menten constant (Km) were calculated based on the Lineweaver-Burk plot
209
according to the following equation:
210
v=
Vmax ⋅ [A][B] [A][B] + [B] K mA + [A]K Bm + K sA K Bm
211
where [A] and [B] are concentrations of NADPH and (S)-CHOH; Vmax is the
212
maximal reaction rate; K Am and K Bm correspond to apparent kinetic parameters of
11
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
213
RtSCR9 towards NADPH and (S)-CHOH; K sA represents the dissociation constant
214
between RtSCR9 and NADPH.
215
Effects of pH and temperature on recombinant carbonyl reductase.
216
To determine the optimal catalytic conditions of wild and mutant enzymes, the effect
217
of temperature and pH were investigated under standard enzyme activity assay
218
conditions. The optimal temperatures for recombinant RtSCR9 and its variants were
219
determined at the range of 25°C to 65°C. The thermostability of recombinant RtSCR9
220
was examined by preincubating the purified enzymes in 100 mM phosphate buffer,
221
pH 7.0, at 30°C, 40°C and 50°C for required time intervals. The residual activity was
222
detected with non-preheated enzyme as control (100% of activity). The optimum pH
223
was evaluated at 30°C with the assay mixture in different buffers including
224
citrate-Na2HPO4 (pH 4.0-6.5), KH2PO4-K2HPO4 (pH 6.0-8.0) and Tris-HCl (pH
225
7.5-9.0). The pH stabilities were examined by preincubating the purified enzyme in
226
different KH2PO4-K2HPO4 (100 mM, pH 5.0-9.0) for 1 h and then the residual
227
activities of the enzymes were detected. The non-preinbubated enzyme was taken as
228
control (100% of activity).
229
Molecular modeling. The crystal structure of Thermotoga maritima SCR
230
(PDB-code: 1 VL8) in complex with NADPH was used as template for
231
three-dimensional (3D) analysis of RtSCR9 structure. The MODELLER 9.12 program
232
(http://www.salilab.rog) was used to construct the relaxed models of RtSCR9. Models
233
with the highest scores were selected and further evaluated using the PROCHECK
234
program.35 The final models were used for docking analysis with the AutoDock 4.2.1 12
ACS Paragon Plus Environment
Page 12 of 42
Page 13 of 42
Journal of Agricultural and Food Chemistry
235
program (Scripps Res. Inst., La Jolla, CA). The 3D structure of (S)-CHOH and
236
NADPH were created using ChemDraw 8.0 (CambridgeSoft, Cambridge, UK). The
237
structures were visualized with the PyMOL program (http://www.pymol.org).
238
Asymmetric bioreduction of (S)-CHOH in a single batch reaction. The
239
bioconversion of (S)-CHOH to (3R,5S)-CDHH using RtSCR9 and mut-Ile144Lys was
240
investigated using whole E. coli cells in single batch reactions. Recombinant GDH
241
was used to complete the “enzyme-coupled” cofactor regenerating system (Figure S1).
242
The reaction mixture (30 mL) containing 0.9 g RtSCR9 (or mut-Ile144Lys) wet cell,
243
0.9 g GDH wet cell, 8.25 g (S)-CHOH and 8.77 g glucose was preheated for 5 min at
244
30°C on a magnetic stirring apparatus. During the process of biotransformation, 50 µL
245
samples were extracted per hour and diluted 200 times with 30% acetonitrile
246
(acetonitrile: water, 30:70 v/v) to a proper concentration and filtered before HPLC
247
analysis. The pH of the reaction mixture was controlled by an automatic regulation
248
system 902 Titrando Titrator (Metrohm Inc., Herisau, Switzerland) through titrating
249
an 8.0 M NaOH solution at a flow rate of 0.2 mL/min. It is noteworthy that the
250
volume of (S)-CHOH and titrated NaOH solution were calculated into the total
251
volume of the whole biotransformation reaction. The amount of products were
252
calculated according to the standard curves of (S)-CHOH and (3R,5S)-CDHH. The
253
yield represents the ratio between the formed (3R,5S)-CDHH and the original
254
(S)-CHOH.
255
After filtering, the reaction mixture was saturated by 5 g NaCl and then extracted
256
three times with 30 mL ethyl acetate. The product in organic phase was thoroughly 13
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
257
dried with 5 g anhydrous Na2SO4. The solvent was removed by rotary evaporation
258
under 40°C until the weight of product maintains to a constant, and then the yield of
259
pure product was calculated according to the following equation:
260
Yield (%) =
purified product (g) ×100 8.32 (g)
261
Statistical analysis. If not specifically noted, all experiments in this study
262
were performed in triplicate. Analysis of variance (ANOVA) was carried out using
263
the SAS program version 8.1 (SAS Institute Inc., Cary, NC). Least significant
264
difference (LSD) was computed at p < 0.05. All the Figures in this study were drawn
265
using the origin software version 8.0 (OriginLab Corp., Northampton, MA).
266
RESULTS AND DISCUSSION
267
Construction and screening of mutant libraries. Directed evolution does
268
not require detailed information of secondary structure or catalytic mechanism of
269
proteins.36 Site saturation mutagenesis at hotspot is a unique method for creating
270
focused libraries whereby each amino acid is substituted by the other 19 naturally
271
occurring amino acids.37
272
To improve the enzyme activity of RtSCR9, random mutagenesis strategy was
273
initially adopted through error-prone PCR and megaprimer PCR to mimic in vitro
274
Darwinian evolution. Random mutations were introduced into the gene of RtSCR9 by
275
error-prone PCR. The genes carrying the mutations were then cloned into the
276
expression vector by megaprimer PCR (Figure S3). Over 10,000 colonies were picked
277
in an initial round of mutagenesis. The majority of transformants carrying negative 14
ACS Paragon Plus Environment
Page 14 of 42
Page 15 of 42
Journal of Agricultural and Food Chemistry
278
mutations scarcely showed any detectable activity, while only a few numbers of
279
mutants exhibited similar or improved activity compared with wild type strain (Figure
280
S4). Mutants with higher activities were selected for a secondary screening by
281
determining the bioconversion of (S)-CHOH to (3R,5S)-CDHH in a 10-mL reaction
282
mixture and further determined using HPLC. Three positive mutations (Gln95Asp,
283
Ile144Leu and Phe156Asn) at three different residual sites were obtained, displaying
284
1.96-, 1.20- and 1.65-fold increased specific activity, respectively (Table S2).
285
Though three hotspots were identified through random mutagenesis, molecular
286
understanding of the improved properties of mutants was still vague and the best
287
substitutions at sites 95, 144 and 156 could not be identified. Site-saturation
288
mutagenesis was performed therefore to further explore the most beneficial
289
substitutions in this study (Table S3). After site-saturation mutagenesis, enzyme
290
activities of different mutants were also determined by HPLC. Through site-saturation
291
mutagenesis, there was more than one mutant that showed improved activity at each
292
site (Figure 1). Among the mutants, Gln95Cys, Gln95Asp, Gln95Glu and Gln95Asn
293
showed a more significant improvement of activity (Figure 1a); Ile144Lys was
294
obviously superior to other substitutions in the conversion of the substrate (S)-CHOH.
295
In addition, Ile144Ala, Ile144Leu and Ile144Gln also showed a little improvement of
296
the enzyme activity (Figure 1b). For substitutions of Phe156, Phe156Cys, Phe156Lys,
297
Phe156Asn, Phe156Gln, Phe156Ser, Phe156Thr and Phe156Arg all showed improved
298
enzyme activity (Figure 1c). Based on the above, Gln95Asp, Ile144Lys and
299
Phe156Gln were finally selected for further studies. 15
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 16 of 42
300
Kinetic studies. The kinetic parameters were determined using (S)-CHOH as
301
substrate at different concentrations (0-10 mM) and cofactor NADPH with a varied
302
concentration (0.5-2.5 mM). Compared with the wild-type RtSCR9, mut-Gln95Asp
303
and mut-Ile144Lys showed similar Km values, mut-Phe156Gln gave a 1.23-fold
304
decrease of Km. The Vmax of Gln95Asp, Ile144Lys and Phe156Gln were improved
305
1.94-, 3.03- and 1.61-fold, respectively, compared to wild type (Table 1). All three
306
mutants exhibited increased kcat values, giving 1.94-, 3.03- and 1.61-fold improvement.
307
The kcat/Km values increased to 1.93-, 3.15- and 1.97-fold. Improved kcat and kcat/Km
308
values mainly contributed to the enhancement of the catalytic efficiency.
309
Purification and biochemical properties of purified wild-type and
310
evolved RtSCR9. Recombinant RtSCR9 and its variants were purified through
311
immobilized metal affinity chromatography. Both crude protein and the purified
312
enzyme were examined by SDS-PAGE analysis (Figure 2). There are no big
313
differences
314
mut-Ile144Lys and mut-Phe156Gln. Single site amino acid substitution might not
315
affect the correct protein folding and soluble expression. In addition, purified proteins
316
gave only one band on the SDS-PAGE gel without any undesired protein
317
contamination. The protein molecular weights of mut-Gln95Asp, mut-Ile144Lys and
318
mut-Phe156Gln were found nearly 27 kDa, which was identical to the wild-type
319
RtSCR9.
in
the
protein
expression
levels
of
RtSCR9,
mut-Gln95Asp,
320
The effects of different metal ions and the metal chelator EDTA on the activity
321
of RtSCR9 were reported in our previous work32, indicating the non-metal 16
ACS Paragon Plus Environment
Page 17 of 42
Journal of Agricultural and Food Chemistry
322
dependence of RtSCR9, which differs from the most representative carbonyl reductase
323
LbADH.38 The effects of temperature and pH on the enzymes catalyzing the
324
conversion of (S)-CHOH to (3R,5S)-CDHH were investigated. Temperature
325
influences molecule movement and the protonation process during interactions
326
between (S)-CHOH molecules and amino acid residures.39 The optimum catalytic
327
temperature of both RtSCR9 and its variants appeared to be 30°C. The activities of
328
wild type RtSCR9 and mutants were maintained at over 90% between 25°C and 40°C,
329
decreased rapidly between 45°C and 55°C and were completely lost when the
330
temperature went over 60°C (Figure 3a). The studies showed that the amino acid
331
substitutions at position 95 and 144 did not change the thermostability of the enzyme.
332
Phe156Gln did, however, retain >80% activity after incubation in phosphate buffer,
333
pH 7.0, for 24 h at 30°C, which was better than other mutants. Thermostability tests
334
under 40°C also demonstrated that mut-Phe156Gln had a slightly better stability
335
(Figure 3b). Besides protolysis, pH conditions also influence the hydrolysis of
336
ketoester substrates.40 To investigate the effect of pH on enzyme activity, reactions
337
were performed in buffers at pH 4.0 to 9.0. RtSCR9, mut-Gln95Asp, mut-Ile144Lys
338
and mut-Phe156Gln showed almost the same trends of activities under different pH
339
conditions. The activities of RtSCR9 and mutants increased slowly from pH 4.0 to 6.0
340
and the maximum activities were observed at 6.5 and 7.0, then began to decrease at
341
7.5 to 9.0 (Figure 3c). The investigation of pH stability of wild type RtSCR9 and its
342
variants showed that there was no apparent difference before and after protein
343
engineering. RtSCR9, Gln95Asp and Ile144Lys showed good stability between pH 17
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
344
6.5 and pH 8.5, retaining >90% of original activity, while Phe156Gln had a wider
345
range of pH tolerance, from pH 5.5 to pH 8.5 (Figure 3d). The effects of temperature
346
and pH indicated that mutant Phe156Gln was more thermostable and had a wider pH
347
catalytic condition than the others.
348
Homology modeling and analysis of mutants based on molecular
349
simulation. To elucidate the mechanism for the enhancement of activity regarding
350
the amino acid substitutions, homology modeling of the enzyme structures and
351
docking between enzymes and substrates were performed. Based on structure (PDB: 1
352
VL8), the three-dimensional structures of RtSCR9 and mutants were built.41
353
Validation with the PROCHECK program revealed that 92.8% of the residues were
354
presented in the most favored regions of the structure (Figure S5), and no residues
355
were in disallowed regions, suggesting that the built model was of high quality and
356
suitable for further analysis. The catalytic tetrads of RtSCR9 were Asn115, Thr143,
357
Tyr161 and Lys165. The cofactor NADPH was embedded in the coenzyme binding
358
area (Figure S6). The traits of substrate binding pockets play a key role in binding
359
process between ligands and receptors and thus impact the catalytic efficiency of
360
enzymes.42 Generally, an appropriately sized pocket is beneficial for substrate binding
361
and product release. Large ligands encountering a narrow pocket might not enter the
362
catalytic center, while a loose pocket environment can cause wrong positioning of
363
ligands, leading to the opposite stereo-configuration of the products.43
364
It was observed that the chloromethyl group of (S)-CHOH entered the substrate
365
pocket first and embedded into a hydrophobic cavity composed of Ile144, Ala145, 18
ACS Paragon Plus Environment
Page 18 of 42
Page 19 of 42
Journal of Agricultural and Food Chemistry
366
Ala149 and Ser150 (Figure 4a, e). The terminal ester group with its large steric
367
hindrance pointed away from the substrate pocket, which consisted of Ser94, Gln95,
368
Ser97, Met196, Ile197 and Gln200. Thr143 and Tyr161 stabilized (S)-CHOH by
369
forming H-bonds with C-3 carbonyl oxygen atom, which was within attacking
370
distance from a hydrogen atom NADPH C-4 at the si-face.
371
Gln95, located on the inlet of the substrate pocket, influenced the entrance of
372
substrate molecules. When replaced with Asp, the carboxyl of Asp oriented down
373
toward the entryway of ligands and reduced the inlet size of the pocket (Figure 4b, f).
374
This change could lead to a well-orientated tert-butyl ester in the binding pocket, as
375
the motility of the terminal ester group impacts the correct positioning of the
376
C3-carbonyl. Positive substitutions (Cys, Asp, Glu and Asn) from the site-saturation
377
mutagenesis experiments had similar spatial structures with the original Gln. It is
378
concluded that the dramatic change in steric hindrance of the amino acid in this site
379
could cause irregular features of the outer pocket, blocking the entrance of substrate
380
molecules.
381
When (S)-CHOH molecules entered the pocket in the simulations, the
382
chloromethyl group positioned next to the Ile144, and the methyl group substitution
383
on the β’ carbon atom of Ile increased the steric hindrance. When the chlorine atom
384
stretched upward towards the deeper hydrophobic cavity, the carbonyl moiety of the
385
substrate was put in a point where interactions between C-3 carbonyl oxygen and
386
active residues (Thr143 and Tyr156) were weakened. Substitution of Ile with the long
387
chain amino acid Lys showed that a small arc could form between the β’ and γ’ 19
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
388
carbon atoms, making room for the chloromethyl group (Figure 4c, g). The distance
389
from the hydrogen atom of NADPH to the substrate carbonyl oxygen atom shortened
390
from 3.3 Å to 3.2 Å after substitution of Ile to Lys. Ala, Leu, and Gln substitutions at
391
site 144 also improved the activity of the enzyme. Ala reduced the steric hindrance of
392
hydrophobic cavity, which facilitated the entrance of chloromethyl group. Leu has
393
similar steric hindrance but with a methyl substitution at γ’ carbon atom. It is
394
presumed that the β’-substitution has bigger impact on the positioning of
395
chloromethyl group than γ’-substitution. Most substitutions at this site were, however,
396
negative. The reason might be that site 144 is next to the catalytic tetrad and amino
397
acid variations at this site influence the enzyme function.
398
The third mutation, Phe156Gln, was on the outer surface of the enzyme (Figure
399
4d), which is associated with the stability of RtSCR9. When Phe156 bound by
400
substrate was mutated with the uncharged polar amino acids Gly, Ser, Thr, Cys, Asn,
401
Tyr, and Gln, substitutions of Ser, Thr, Cys, Asn, and Gln exhibited improved enzyme
402
activity. Gly156 had no activity, which might be due to the tremendous difference in
403
both the structure and properties of the two amino acids. Substitutions of Phe156 with
404
nonpolar amino acids Ala, Vla, Leu, Ile, Pro, Trp, and Met lost activity completely
405
except for Phe156Trp. Since Phe, Tyr and Trp are all aromatic amino acids sharing
406
close conformations and physicochemical characteristics, Phe156Tyr and Phe156Trp
407
could retain the original activity of the wild enzyme. In addition, substitutions with
408
negatively charged amino acids, Phe156Asp and Phe156Glu, also weakened the
409
enzyme activity. Oppositely, positively charged amino acids Lys and Arg enhanced 20
ACS Paragon Plus Environment
Page 20 of 42
Page 21 of 42
Journal of Agricultural and Food Chemistry
410
the catalytic activity of the enzyme. Thermostability tests (Figure 3b) and B-factor
411
analysis (Table S4) of Phe156Gln also confirmed that it displayed more stability than
412
the other enzymes.44-46 Pro157 and Val158 linked to the Phe156, as well as several
413
other amino acids, constituted the inner surface of the substrate pocket (Figure 4 h).
414
Substitution at site 156 also influenced the conformations of sites 157 and 158, thus
415
further changing the shape of the inner surface.
416
Whole cell asymmetric reduction of (S)-CHOH in a single batch
417
reaction. To verify the potential application of mut-Ile144Lys, a time course
418
bioconversion process was conducted to verify its catalytic efficiency in single batch
419
reactions. GDH was used as a coupled enzyme to complete the cofactor regeneration
420
(Figure S7).47-49 Glucose was added as co-substrate. The reaction pH was controlled at
421
approximately 7.0 by continuous titration with 8.0 M NaOH. Both RtSCR9 and
422
mut-Ile144Lys displayed outstanding catalytic ability towards (S)-CHOH (Figure 5).
423
During the first 2 h, the yield of product grew slowly because of mass transfer
424
resistance, and then the catalytic efficiency increased rapidly. When the yield of
425
(3R,5S)-CDHH was over 90%, the reaction rate slowed down gradually and finally
426
leveled at >95% after 12-h biotransformation. In comparison, the asymmetric
427
reduction process for mut-Ile144Lys presented a higher product yield, reaching 98.9%
428
after an 8-h reaction. The highest space-time yield of (3R,5S)-CDHH increased from
429
245.17 mmol L-1 h-1 g-1 wet cell weight (WCW) by RtSCR9 to 542.83 mmol L-1 h-1 g-1
430
WCW by mut-Ile144Lys. After isolation, 7.84 g (3R,5S)-CDHH was finally harvested,
21
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
431
total yield of product reached 95.1%. These results demonstrated the high catalytic
432
efficiency and the potential for industrial application of mut-Ile144Lys.
433
In conclusion, compared with wild type enzyme, the mut-Ile144Lys displayed
434
3.03- and 3.15-fold improvement of specific activity and kcat/Km respectively.
435
Potential mechanisms for the improvements in catalytic efficiency were studied
436
through homology modeling and molecular docking, indicating that there is still
437
opportunity for further evolution on catalytic activity and stabilities. Mut-Ile144Lys
438
catalyzed 542.83 mmol L-1 h-1 g-1 WCW during a single batch reaction, demonstrating
439
its great potential for industrial application.
440
ASSOCIATED CONTENT
441 442 443
The Supporting Information is available free of charge on the ACS Publications website at DOI:
444
Figure S1: HPLC detection of (3R,5S)-CDHH and (S)-CHOH. Figure S2:
445
enantiomeric excess detection of (3R,5S)-CDHH and (S)-CHOH. Figure S3: agarose
446
gel electrophoresis of error-prone PCR and megaprimer PCR. Figure S4: screening of
447
clones in a 96-well plate by measuring the absorbance of NADPH at OD340. Figure S5:
448
ramachandran plot of RtSCR9 chain A. Figure S6: homology modeled structure of
449
RtSCR9 monomer and the location of mutant sites. Figure S7: asymmetric reduction
450
of (S)-CHOH by recombinant E. coli BL21(DE3)/pET28a-RtSCR9 coupled with E.
451
coli BL21(DE3)/pET28a-GDH. Table S1: primers for PCR amplification. Table S2:
452
specific activities of positive mutants obtained from random mutagenesis. Table S3: 22
ACS Paragon Plus Environment
Page 22 of 42
Page 23 of 42
Journal of Agricultural and Food Chemistry
453
relative activity and enantioselectivity towards (3R,5S)-CDHH of mutants generated
454
by saturation mutagenesis. Table S4: the 20 amino acids of RtSCR9 chain A with the
455
highest B-values.
456
AUTHOR INFORMATION
457 458
*Phone: +86-571-88320630, E-mail:
[email protected] 459 460
This study was financially supported by the National Natural Science Foundation of
461
China (no. 21672190) and the Program of Science and Technology of Zhejiang
462
Province (no. 2015C33137).
463 464
The authors declare no competing financial interest.
465
23
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Page 24 of 42
466
REFERENCES
467
(1) Časar, Z. Historic overview and recent advances in the synthesis of super-statins.
468
Curr. Org. Chem. 2010, 14, 816-845.
469
(2) Dunn, P. J. The importance of green chemistry in process research and
470
development. Chem. Soc. Rev. 2012, 41, 1452-1461.
471
(3) Sirtori, C. R. The pharmacology of statins. Pharmacol. Res. 2014, 88, 3-11.
472
(4)
473
3-hydroxy-3-methylglutaryl–coenzyme A reductase inhibitors. Tetrahedron. 2015, 71,
474
8487-8510.
475
(5) Xiong, F. J.; Li, J.; Chen, X. F.; Chen, W. X.; Chen, F. E. An improved process
476
for chiron synthesis of the atorvastatin side chain. Tetrahedron: Asymmetry 2014, 25,
477
1205-1208.
478
(6) Gupta, P.; Mahajan, N.; Taneja, S. C. Recent advances in the stereoselective
479
synthesis of 1,3-diols using biocatalysts. Catal. Sci. Technol. 2013, 3, 2462-2480.
480
(7) Bornscheuer, U. T.; Huisman, G. W.; Kazlauskas, R. J.; Lutz, S.; Moore, J. C.;
481
Robins, K. Engineering the third wave of biocatalysis. Nature 2012, 485, 185-194.
482
(8) Reetz, M. T. Biocatalysis in organic chemistry and biotechnology: past, present,
483
and future. J. Am. Chem. Soc. 2013, 135, 12480-12496.
484
(9) Liu, Z. Q.; Dong, S. C.; Yin, H. H.; Xue, Y. P.; Tang, X. L.; Zhang, X. J.; He, J.
485
Y.; Zheng, Y. G. Enzymatic synthesis of an ezetimibe intermediate using carbonyl
486
reductase coupled with glucose dehydrogenase in an aqueous-organic solvent system.
487
Bioresour. Technol. 2017, 229, 26-32.
Wu,
Y.;
Xiong,
F.
J.;
Chen,
F.
E.
Stereoselective
24
ACS Paragon Plus Environment
synthesis
of
Page 25 of 42
Journal of Agricultural and Food Chemistry
488
(10) Muller, M. Chemoenzymatic synthesis of building blocks for statin side chains.
489
Angew. Chem. Int. Ed. 2005, 44, 362-365.
490
(11) Wolberg, M.; Hummel, W.; Muller, M. Biocatalytic reduction of β, δ-diketo
491
esters: a highly stereoselective approach to all four stereoisomers of a chlorinated β,
492
δ-dihydroxy hexanoate. Chem. Eur. J. 2001, 7, 4562-4571.
493
(12) Wolberg, M.; Hummel, W.; Wandrey, C.; Muller, M. Highly regio- and
494
enantioselective reduction of 3,5-dioxocarboxylates. Angew. Chem. 2000, 112,
495
4476-4478.
496
(13) Wolberg, M.; Hummel, W.; Wandrey, C.; Muller, M. Highly regio-and
497
enantioselective reduction of 3, 5-dioxocarboxylates. Angew. Chem. Int. Ed. 2000, 39,
498
4306-4308.
499
(14) Bradshaw, C. W.; Hummel, W.; Wong, C. H. Lactobacillus kefir alcohol
500
dehydrogenase: a useful catalyst for synthesis. J. Orc. Chem. 1992, 57, 1532-1536.
501
(15) Haberland, J.; Kriegesmann, A.; Wolfram, E.; Hummel, W.; Liese, A.
502
Diastereoselective synthesis of optically active (2R, 5R)-hexanediol. Appl. Microbiol.
503
Biotechnol. 2002, 58, 595-599.
504
(16) Hummel, W. Enzyme-catalyzed synthesis of optically pure R (+)-phenylethanol.
505
Biotechnol. Lett. 1990, 12, 403-408.
506
(17) Hummel, W. Reduction of acetophenone to R (+)-phenylethanol by a new
507
alcohol dehydrogenase from Lactobacillus kefir. Appl. Microbiol. Biotechnol. 1990,
508
34, 15-19.
25
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
509
(18) Pfruender, H.; Amidjojo, M.; Hang, F.; Weuster-Botz, D. Production of
510
Lactobacillus kefir cells for asymmetric synthesis of a 3,5-dihydroxycarboxylate.
511
Appl. Microbiol. Biotechnol. 2005, 67, 619-622.
512
(19) Pfruender, H.; Amidjojo, M.; Kragl, U.; Weuster-Botz, D. Efficient whole-cell
513
biotransformation in a biphasic ionic liquid/water system. Angew. Chem. Int. Ed.
514
2004, 43, 4529-4531.
515
(20) Amidjojo, M.; Franco-Lara, E.; Nowak, A.; Link, H.; Weuster-Botz, D.
516
Asymmetric synthesis of tert-butyl (3R,5S) 6-chloro-dihydroxyhexanoate with
517
Lactobacillus kefir. Appl. Microbiol. Biotechnol. 2005, 69, 9-15.
518
(21) Pfruender, H.; Jones, R.; Weuster-Botz, D. Water immiscible ionic liquids as
519
solvents for whole cell biocatalysis. J. Biotechnol. 2006, 124, 182-190.
520
(22) Kizaki, N.; Yamada, Y.; Yasohara, Y.; Hasegawa, J. Carbonyl reductase, gene
521
thereof and method of using the same. US Patent 6645746 B1, 2003.
522
(23) Kizaki, N.; Yamada, Y.; Yasohara, Y.; Nishiyama, A.; Miyazaki, M.; Mitsuda,
523
M.; Kondo, T.; Ueyama, N.; Inoue, K. Process for the preparation of optically active
524
2-[6-(hydroxymethyl)-1,3-dioxan-4yl] acetic acid derivatives. US Patent 6472544 B1,
525
2002.
526
(24) Sun, X. Y.; Shi, H. B.; Bi, H. X.; Ou, Z. M. Bioconversion process for synthesis
527
of tert-butyl (3R, 5S)-6-chloro-3,5-dihydroxyhexanoate using liquid-core immobilized
528
Saccharomyces cerevisiae CGMCC No 2233. Korean J. Chem. Eng. 2013, 30,
529
166-171.
26
ACS Paragon Plus Environment
Page 26 of 42
Page 27 of 42
Journal of Agricultural and Food Chemistry
530
(25) Ni, Y.; Xu, J. H. Biocatalytic ketone reduction: a green and efficient access to
531
enantiopure alcohols. Biotechnol. Adv. 2012, 30, 1279-1288.
532
(26) Zhang, D. L.; Chen, X.; Chi, J.; Feng, J. H.; Wu, Q. Q.; Zhu, D. M.
533
Semi–rational engineering a carbonyl reductase for the enantioselective eeduction of
534
β-amino ketones. ACS Catal. 2015, 5, 2452-2457.
535
(27) Zheng, M. M.; Chen, K. C.; Wang, R. F.; Li, H.; Li, C. X; Xu, J. H. Engineering
536
7β-hydroxysteroid dehydrogenase for enhanced ursodeoxycholic acid production by
537
multiobjective directed evolution. J. Agric. Food Chem. 2017, 65, 1178-1185.
538
(28) Cao, L. C.; Chen, R.; Xie, W.; Liu, Y. H. Enhancing the thermostability of
539
feruloyl esterase EstF27 by directed evolution and the underlying structural basis. J.
540
Agric. Food Chem. 2015, 63, 8225-8233.
541
(29) Giver, L. J.; Newman, L. M.; Mundorff, E.; Huisman, G. W.; Jenne, S. J.; Zhu,
542
J.; Behrouzian, B.; Grate, J. H.; Lalonde, J. Polynucleotides encoding ketoreductases
543
for producing stereoisomerically pure statins and synthetic intermediates therefor. US
544
Patent 0040364 A1, 2013.
545
(30) He, X. J.; Chen, S. Y.; Wu, J. P.; Yang, L. R.; Xu, G. Highly efficient enzymatic
546
synthesis of tert-butyl (S)-6-cholro-5-hydroxy-3-oxohexanoate with a mutant alcohol
547
dehydrogenase of Lactobacillus kefir. Appl. Microbiol. Biotechnol. 2015, 99,
548
8963-8975.
549
(31) Luo, X.; Wang, Y. J.; Shen, W.; Zheng, Y. G. Activity improvement of a
550
Kluyveromyces lactic aldo-keto reductase KIAKR via rational design. J. Biotecnol.
551
2016, 224,20-26. 27
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
552
(32) Chen, X.; Liu, Z. Q.; Lin, C. P.; Zheng, Y. G. Chemoenzymatic synthesis of
553
(S)-duloxetine using carbonyl reductase from Rhodosporidium toruloides. Bioorg.
554
Chem. 2016, 65, 82-89.
555
(33) You, Z. Y.; Liu, Z. Q.; Zheng, Y. G. Characterization of a newly synthesized
556
carbonyl reductase and construction of a biocatalytic process for the synthesis of ethyl
557
(S)-4-chloro-3-hydroxybutanoate with high space-time yield. Appl. Microbiol.
558
Biotechnol. 2014, 98, 1671-1680.
559
(34) Laemmli, U. K. Cleavage of structure proteins during the assembly of head of
560
bacteriophage T4. Nature 1970, 227, 680-685.
561
(35) Laskowski, R. A.; Macarthur, M. W.; Moss, D. S.; Thornton, J. M. PROCHECK:
562
a program to check the stereochemical quality of protein structures. J. Appl. Cryst.
563
1993, 26, 283-291.
564
(36) Joshi, S., Satyanarayana, T. In vitro engineering of microbial enzymes with
565
multifarious application: Prospects and perspectives. Bioresour. Technol. 2015, 176,
566
273-283.
567
(37) Yang, H. Q.; Li, J. H.; Shin, H. D.; Du, G. C.; Liu, L.; Chen, J. Molecular
568
engineering of industrial enzymes: recent advances and future prospects. Appl.
569
Microbiol. Biotechnol. 2014, 98, 23-29.
570
(38) Noey, E. L.; Tibrewal, N.; Jiménez-Osés, G.; Osuna, S.; Park, J.; Bond, C. M.;
571
Cascio, D.; Liang, J.; Zhang, X.; Huisman, G. W.; Tang, Y.; Houk, K. N. Origins of
572
stereoselectivity in evolved ketoreductases. Proc. Natl. Acad. Sci. 2015, 112,
573
7065-7072. 28
ACS Paragon Plus Environment
Page 28 of 42
Page 29 of 42
Journal of Agricultural and Food Chemistry
574
(39) Niefind, K.; Müller, J.; Riebel, B.; Hummel, W.; Schomburg, D. The crystal
575
structure of R-specific alcohol dehydrogenase from Lactobacillus brevis suggests the
576
structural basis of its metal dependency. J. Mol. Biol. 2003, 327, 317-328.
577
(40) Cha, J., Batt, C. A. Lowering the pH optimum of D-xylose isomerase: the effect
578
of mutations of negatively charged residues. Mol. Cells 1998, 8, 374-382.
579
(41) Takeshita, D.; Kataoka, M.; Miyakawa, T.; Miyazono, K.; Kumashiro, S.; Nagai,
580
T.; Urano, N.; Uzura, A.; Nagata, K.; Shimizu, S.; Tanokura, M. Structural basis of
581
stereospecific reduction by quinuclidinone reductase. AMB Express 2014, 4, 6.
582
(42) Zheng, G. W.; Xu, J. H. New opportunities for biocatalysis: driving the synthesis
583
of chiral chemicals. Curr. Opin. Biotechnol. 2011, 22, 784-792.
584
(43) Musa, M. M.; Lott, N.; Laivenieks, M.; Watanabe, L.; Vieille, C.; Phillips, R. S.
585
A single point mutation reverses the enantiopreference of Thermoanaerobacter
586
ethanolicus secondary alcohol dehydrogenase. ChemCatChem 2009, 1, 89-93.
587
(44) Parthasarathy, S.; Murthy, M. R. N. Protein thermal stability: insights from
588
atomic displacement parameters (B values). Protein Eng. 2000, 13, 9-13.
589
(45) Reetz, M. T.; Carballeira, J. D.; Vogel, A. Iterative saturation mutagenesis on the
590
basis of B factors as a strategy for increasing protein thermostability. Angew. Chem.
591
Int. Ed. 2006, 45, 7745-7751.
592
(46) Reetz, M. T.; Carballeira, J. D. Iterative saturation mutagenesis (ISM) for rapid
593
directed evolution of functional enzymes. Nat. Protoc. 2007, 2, 891-903.
29
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
594
(47) Dascier, D.; Kambourakis, S.; Hua, L.; Rozzell, J. D., Stewart, J. D. Influence of
595
cofactor regeneration strategies on preparative-scale, asymmetric carbonyl reductions
596
by engineered Escherichia coli. Org. Process Res. Dev. 2014, 18, 793-800.
597
(48) Chenault, H. K., Whitesides, G. M., Regeneration of nicotinamide cofactors for
598
use in organic synthesis. Appl. Biochem. Biotechnol. 1986, 14, 147-197.
599
(49) Pennacchio, A.; Giordano, A.; Rossi, M.; Raia, C. A. Asymmetric reduction of
600
α-keto esters with Thermus thermophilus NADH-dependent carbonyl reductase using
601
glucose dehydrogenase and alcohol dehydrogenase for cofactor regeneration. Eur. J.
602
Org. Chem. 2011, 23, 4361-4366.
30
ACS Paragon Plus Environment
Page 30 of 42
Page 31 of 42
Journal of Agricultural and Food Chemistry
FIGURE CAPTIONS Figure 1 Relative activities of mut-Gln95 (a), mut-Ile144 (b) and mut-Phe156 (c) mutants generated by site-saturation mutagenesis, comparing the wild-type, are presented in the bar graph.
Figure 2 SDS-PAGE analysis of wild type and mutants. Lane M: Protein Marker; Lane 1: Supernatant extract of BL21-pET28a; Lane 2: Supernatant extract of wild type RtSCR9; Lane 3: Supernatant extract of mut-Gln95Asp; Lane 4: Supernatant extract of mut-Ile144Lys; Lane 5: Supernatant extract of mut-Phe156Gln; Lane 6: Purified wild type RtSCR9. Lane 7: Purified mut-Gln95Asp; Lane 8: Purified mut-Ile144Lys; Lane 9: Purified mut-Phe156Gln.
Figure 3 Effects of temperature and pH on recombinant enzyme activities. a. Effects of temperature on enzymatic activities; b. Thermostabilities of RtSCR9 and mutants; c. Effects of pH on enzyme activities; d. pH stability of RtSCR9 and mutants.
Figure 4 Molecular docking analysis. Conformations of (S)-CHOH (cyan sticks) within RtSCR9 (a, e), mut-Gln95Asp (b, f), mut-Ile144Lys (c, g), mut-Phe156Gln (e, h). a, b, c, and d show the interaction of (S)-CHOH with active sites and NADPH. H-bonds are shown as red dashes, and the distance of (S)-CHOH and the C4 atom of the nicotinamide ring are shown as black dashes. Three mutants (green), active sites (white) and NADPH (gray) are displayed as sticks. Key amino acids in the substrate binding pocket are drawn in purple lines. e, f, g and h show the general conformation of the substrate pocket when (S)-CHOH is embedded.
31
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Figure 5 Time course of asymmetric reduction of (S)-CHOH in a batch reaction catalyzed by wild-type RtSCR9 and mut-Ile144Lys.
32
ACS Paragon Plus Environment
Page 32 of 42
Page 33 of 42
Journal of Agricultural and Food Chemistry
Table 1. Apparent kinetic parameters of RtSCR9 and mutants. a
kcat
kcat/Km
e.e.
(mM)
(s-1)
(mM-1 s-1)
(%)
62.84 ± 4.54
0.83 ± 0.05
28.28 ± 4.83
33.99 ± 2.06 >99
mut-Gln95Asp
122.15 ± 4.73
0.84 ± 0.06
54.97 ± 4.67
65.44 ± 3.27 >99
mut-Ile144Lys
190.55 ± 7.71
0.80 ± 0.03
85.75 ± 5.12
107.19 ± 3.58 >99
mut-Phe156Gln
101.14 ± 5.02
0.68 ± 0.05
45.51 ± 3.29
67.12 ± 4.14 >99
Vmax
Km
(µmol min-1) RtSCR9
Enzyme
a
kcat, where [E] is the molar concentration of the enzymes.
33
ACS Paragon Plus Environment
AL CYA AS S G P L PHU G E LY H I IL S LYE LE S M U E AS T PR N G O L A N RG SE THR VAR TRL TY P R
Relative activity (%) AL CYA AS S G P L PHU G E LY H I IL S LYE LE S M U E AS T PR N G O L A N RG SE THR VAR TRL TY P R
Relative activity (%)
Journal of Agricultural and Food Chemistry
FIGURE GRAPHICS
400 Wild Variants
300
200
100
0
Figure 1a
400 Wild Variants
300
200
100
0
Figure 1b
34
ACS Paragon Plus Environment
Page 34 of 42
Page 35 of 42
Journal of Agricultural and Food Chemistry
400 Wild Variants
Relative activity (%)
300
200
100
AL CA Y AS S G P L PHU G E LY H I IL S LYE LE S M U E AS T PR N G O L AR N SEG THR VAR TRL TY P R
0
Figure 1c
35
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
Figure 2
36
ACS Paragon Plus Environment
Page 36 of 42
Page 37 of 42
Journal of Agricultural and Food Chemistry
120 Relative activity
e.e.
100
80
60
60
40 30 20
0
0 20
30
40
50
60
o
Temperature ( C)
Figure 3a
120
Relative activity (%)
90
60
30
o
o
RtSCR9-30 C o Gln95Asp-30 C o Ile144Lys-30 C o Phe156Gln-30 C
RtSCR9-40 C o Gln95Asp-40 C o Ile144Lys-40 C o Phe156Gln-40 C
0 0
4
8
12
16
20
24
Time (h)
Figure 3b
37
ACS Paragon Plus Environment
e.e. (%)
Relative activity (%)
90
Journal of Agricultural and Food Chemistry
120
Citrate buffer
Phosphate buffer
Tris-HCl buffer
Page 38 of 42
e.e.
90
75
60
50
30
25
0
0 4
5
6
7
8
9
pH of aqueous phase
Figure 3c
120
Relative activity (%)
90
60
RtSCR9 Gln95Asp Ile144Lys Phe156Gln
30
0 5
6
7
8
9
pH of aqueous phase
Figure 3d
38
ACS Paragon Plus Environment
10
e.e. (%)
Relative activity (%)
100
Page 39 of 42
Journal of Agricultural and Food Chemistry
a
b
c
d
Figure 4a-d
39
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
e
f
g
h
Figure 4e-h
40
ACS Paragon Plus Environment
Page 40 of 42
Page 41 of 42
Journal of Agricultural and Food Chemistry
Yield of (3R,5S)-CDHH (%)
120
90 Wild-type RtSCR9 RtSCR9-Ile144Lys
60
30
0
0
4
8
12
16
20
24
Time (h)
Figure 5
41
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
TOC Graphic
42
ACS Paragon Plus Environment
Page 42 of 42