Improvement of the Activity and Stability of Starch-Debranching

J. Agric. Food Chem. , Article ASAP. DOI: 10.1021/acs.jafc.8b06002. Publication Date (Web): November 30, 2018. Copyright © 2018 American Chemical Soc...
0 downloads 0 Views 769KB Size
Subscriber access provided by the Henry Madden Library | California State University, Fresno

Biotechnology and Biological Transformations

Tailoring of active sites lining the catalytic pocket improves the activity and stability of starch-debranching pullulanase from Bacillus naganoensis Xinye Wang, Yao Nie, and Yan Xu J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b06002 • Publication Date (Web): 30 Nov 2018 Downloaded from http://pubs.acs.org on November 30, 2018

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 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 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.

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 27

Journal of Agricultural and Food Chemistry

1

Tailoring of active sites lining the catalytic pocket improves the activity and stability of

2

starch-debranching pullulanase from Bacillus naganoensis

3 4

Xinye Wang†, Yao Nie†,‡*, Yan Xu†,‡

5 6

†School

7

Education, Jiangnan University, Wuxi 214122, China

8

‡Suqian

9

China

of Biotechnology and Key Laboratory of Industrial Biotechnology, Ministry of

Industrial Technology Research Institute of Jiangnan University, Suqian 223814,

10 11

*Correspondence and requests for materials should be addressed to Y.N. (email:

12

[email protected]).

13

Address: School of Biotechnology and Key Laboratory of Industrial Biotechnology, Ministry

14

of Education, Jiangnan University, 1800 Lihu Road, Wuxi 214122, China

15

E-mail: [email protected]

16

Tel.: +86-510-85197760; Fax: +86-510-85918201

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

17

ABSTRACT

18

Pullulanases are well-known debranching enzymes that hydrolyze α-1,6-glycosidic

19

linkages in starch and oligosaccharides. However, most of the pullulanases exhibit limited

20

activity for practical applications. Here, two sites (787 and 621) lining the catalytic pocket of

21

Bacillus naganoensis pullulanase were identified to be critical for enzymatic activity by

22

triple-code saturation mutagenesis. Subsequently, both sites were subjected to NNK-based

23

saturation mutagenesis to obtain positive variants. Among the variants showing enhanced

24

activity, the enzymatic activity and specific activity of D787C were 1.5-fold higher than

25

those of the wild-type (WT). D787C also showed a 1.8-fold increase in kcat and 1.7-fold

26

increase in kcat/Km. In addition, D787C maintained higher activity compared to WT at

27

temperatures over 60 °C. All the positive variants showed higher acid resistance, with D787C

28

maintaining 90 % residual activity at pH 4.0. Thus, enzymes with improved properties were

29

obtained by saturation mutagenesis at the active site.

30 31

Keywords: pullulanase; directed evolution; saturation mutagenesis; catalytic pocket; activity;

32

stability

2

ACS Paragon Plus Environment

Page 2 of 27

Page 3 of 27

34 35

Journal of Agricultural and Food Chemistry

INTRODUCTION Pullulanase (EC 3.2.1.41), an enzyme with a large molecular weight and a multi-domain

36

structure, is an important member of the glycoside hydrolase family 1. Pullulanase

37

specifically catalyzes the hydrolysis of α-1,6-glycosidic linkages of unmodified substrates

38

and is widely used in medicine, materials, organometallic chemistry, and the saccharification

39

industry 2. In the saccharification process, pullulanase, combined with glucoamylase, can

40

efficiently break down starch into monosaccharides. However, the specific activity of

41

pullulanase needs to be further improved for the saccharification reaction, while maintaining

42

its thermostability and acid resistance.

43

Directed evolution is an important method employed to engineer protein to obtain desired

44

characteristics. The most commonly used methods for directed evolution are epPCR, DNA

45

shuffling, and saturation mutagenesis (SM). Instead of iterative cycles of random amino-acid

46

changes in a protein, more attention is focused on improving the efficiency of directed

47

evolution by creating ‘small-and-smart’ libraries 3-10. The combinatorial active-site saturation

48

test and iterative SM have been developed to create mutant libraries with reduced amino acid

49

alphabets 3, 5, 7-8, 11. In addition, Sun et al. developed a structure-guided triple-code saturation

50

mutagenesis (TCSM) method for efficiently improving the stereoselectivity of limonene

51

epoxide hydrolase and enantioselectivity of alcohol dehydrogenase 9-10. This method involved

52

introduction of three rationally chosen amino acids as building blocks for SM at the

53

randomization sites lining the catalytic/binding pocket of pullulanase. These sites were

54

randomly grouped into smaller fragments, i.e., 3–4 sites in one group, to reduce screening

55

efforts to a minimum for 95 % library coverage. After grouping, TCSM was applied on the

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

56

first group to obtain a mutant, whose characteristics were superior to those of the original, as

57

a template for the next mutation of the second group, and so on 10. Although only three amino

58

acids were used as the degenerate codons for replacing native residues in groups, TCSM was

59

useful to identify the sites in the catalytic pocket that have a marked influence on enzyme

60

function.

61

Pullulanase has a complex structure with multiple domains 12-15. Based on the TCSM

62

method, it was feasible to construct a limited library of variants that focused on the catalytic

63

pocket of the enzyme and identify key sites affecting enzymatic activity and/or other

64

characteristics. Then, SM could be applied at the key sites for obtaining positive mutants.

65

Here, we utilized the TCSM approach to identify the key sites in the catalytic pocket that

66

significantly influence the catalytic activity of Bacillus naganoensis pullulanase (BnPUL).

67

Then, we applied NNK (N: adenine/cytosine/guanine/thymine; K: guanine/thymine) codon

68

degeneracy for encoding all 20 canonical amino acids for SM at the key sites to improve

69

enzymatic activity and stability.

70 71

MATERIALS AND METHODS

72

Materials

73

The recombinant plasmid pET-28a-PelB-pul bearing the pullulanase encoding gene pul

74

(GenBank Accession No. JN872757) from B. naganoensis was constructed as described

75

previously 16. Oligonucleotides for polymerase chain reaction (PCR) were synthesized by

76

Genewiz (Suzhou, China). The plasmid preparation kit I (D6943-02) was purchased from

77

Omega Bio-tek (Norcross, GA, USA). The ClonExpress II One Step Cloning Kit was

4

ACS Paragon Plus Environment

Page 4 of 27

Page 5 of 27

Journal of Agricultural and Food Chemistry

78

purchased from Vazyme Biotech Co., Ltd. (Nanjing, China). DNA sequencing was

79

performed by Genewiz (Suzhou, China). HisTrap HP affinity column and disposable PD-10

80

desalting columns were purchased from GE Healthcare Life Sciences (Beijing, China). All

81

commercial chemicals were purchased from Sigma-Aldrich (St. Louis, MO, USA), Takara

82

(Shiga, Japan), or Sinopharm Chemical Reagent (Shanghai, China). Pullulan was purchased

83

from Tokyo Chemical Industry (Chuo-ku, Tokyo).

84

Primer design and library construction

85

The recombinant plasmid pET-28a-PelB-pul was used as a template for TCSM at the key

86

sites in the catalytic pocket. The involved nine residues were divided into three groups: (A)

87

W504, Y506, and W650; (B) T585, S731, and N734; and (C) M621, D787, and Y790. The

88

short fragments of plasmids containing the target mutation sites were amplified using mixed

89

primers F1/R1, F2/R2, and F3/R3 for libraries I, II, and III, respectively (Figure S1). The

90

long fragments of plasmids were amplified by PCR as well. Next, the short and long

91

fragments were linked by homologous recombination using the ClonExpress II One Step

92

Cloning Kit (Figure S1). Primers used for plasmid construction are listed in Table S1.

93

Primers for SM at D787 and M621 sites were used with recombinant plasmid

94

pET-28a-PelB-pul as the template. The short fragments of plasmids containing the target

95

mutation site were amplified using mixed primers 787F/787R and 621F/621R for library

96

D787 and M621, respectively. The long fragments of plasmids were amplified by PCR as

97

well. Next, the short and long fragments were linked by homologous recombination using the

98

ClonExpress II One Step Cloning Kit. All of the reaction mixture was then used to transform

99

Escherichia coli BL21 (DE3) cells. After plating, individual colonies were selected and their

5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

100

plasmids were sequenced. Primers used for plasmid construction are listed in Table S2.

101

Screening and expression of recombinant proteins

102

Cell colonies containing mutant plasmids were picked and spotted at 45 positions on plates

103

containing solid auto-induction Studier 17 medium with 2 % agar, 1 % red pullulan, and 50

104

μg·mL–1 kanamycin, and then the plates were incubated overnight at 37 °C. Colonies forming

105

obvious transparent zones on the screening plates were selected for activity assay and

106

inoculated in 5 mL LB medium containing kanamycin (50 μg·mL–1). After incubation at

107

37 °C and 200 rpm for 10 h, 1 mL of the culture was transferred to a 250 mL flask containing

108

50 mL auto-induction medium (10 g·L–1 α-lactose, 1.0 g·L–1 glucose, 5.0 g·L–1 glycerol, 6.8

109

g·L–1 KH2PO4, 0.25 g·L–1 MgSO4, 10 g·L–1 tryptone, 5.0 g·L–1 yeast extract, 7.1 g·L–1

110

Na2HPO4, 0.71 g·L–1 Na2SO4, and 2.67 g·L–1 NH4Cl, pH 7.5) supplemented with kanamycin

111

(50 μg·mL–1). After cultivation at 37 °C and 200 rpm for the first 2–3 h, the culture was

112

incubated at 17 °C and 200 rpm for another 60 h for target protein expression.

113

Purification of recombinant proteins

114

The cells were lysed by sonication. The supernatant of the cell lysate, containing the crude

115

enzyme, was collected by centrifugation at 18,514 × g at 4 °C for 40 min, and purified by an

116

AKTAxpress system using HisTrap HP affinity column. Elution was carried out using 500

117

mM imidazole in the same buffer at a flow rate of 2.0 mL·min−1. Then, the purified fractions

118

were transferred into low salt buffer [10 mM Tris-HCl (pH 6.5), 0.1 M NaCl, 0.02 % NaN3,

119

and 5 mM DL-dithiothreitol] using disposable PD-10 desalting columns.

120 121

The molecular weight and concentration of the recombinant enzyme were estimated using 10 % (w/v) sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). Gels

6

ACS Paragon Plus Environment

Page 6 of 27

Page 7 of 27

Journal of Agricultural and Food Chemistry

122

were stained with Coomassie Brilliant Blue R250, and a protein ladder ranging from 10 to

123

200 kDa was used for electrophoresis.

124

Pullulanase activity assay

125

The supernatant obtained after cell lysis was assayed for pullulanase activity by measuring

126

the amount of aldehyde released during the enzymatic reaction from a mixture containing

127

pullulan solution and diluted enzyme sample. The reaction mixture, containing 0.2 mL 2 %

128

(w/v) pullulan in 0.1 M sodium acetate buffer (pH 4.5) and 0.2 mL enzyme solution diluted

129

in 0.1 M sodium acetate buffer (pH 4.5), was incubated at 60 °C for 20 min. Next, the amount

130

of released aldehyde was determined using dinitrosalicylic acid and then by measuring the

131

absorbance at 540 nm. One unit of pullulanase activity was defined as the amount of

132

pullulanase that released 1 μmol reducing sugar equivalents per min under the reaction

133

conditions. Enzymatic activity refers to the units of activity per volume of enzyme solution,

134

and specific activity is the units of activity per milligram protein that is measured after

135

protein purification. All assays were performed in triplicates.

136

Determination of Michaelis-Menten parameters

137

Km and kcat of the purified enzymes were determined based on a previously described

138

method 18 at the following pullulan concentrations: 0.1, 0.25, 0.5, 0.75, 1.0, 1.25, 1.5, 1.75,

139

2.0, 2.25, 2.5, 2.75, 3.0, 4.0, 5.0, 6.0, 8.0, 10.0, 15.0, and 20.0 mg·mL–1. The kcat and Km

140

values were obtained by fitting the initial rate data to the Michaelis-Menten equation using

141

nonlinear regression with the GraphPad Prism 19. Values are shown as means from three

142

replicates with standard deviation.

143

Homology modeling

7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

144

The homologous structure of the wild-type (WT) BnPUL was modeled using

145

SWISS-MODEL (https://www.swissmodel.expasy.org). The X-ray structure of pullulanase

146

from B. acidopullulyticus (PDB ID: 2WAN), which has 64 % sequence identity with BnPUL,

147

was used as the structural template.

148

Amino acid sequence analysis by DisMeta server

149

The DisMeta server (www-nmr.cabm.rutgers.edu/bioinformatics/disorder) employs a wide

150

range of disorder prediction tools and several sequence-based structural prediction tools 20.

151

The disorder regions in the amino acid sequence of the BnPUL were predicted using

152

DisMeta.

153 154

RESULTS AND DISCUSSION

155

Selection of sites in the catalytic pocket for TCSM

156

To date, there have been only a few reports focusing on engineering the catalytic pocket of

157

BnPUL 21-22. We employed the TCSM method to try to improve the catalytic efficiency of

158

BnPUL. Except for the catalytic triad, D619, E648, and D733, the functional residues were

159

still unclear even in the active center of BnPUL. In the Protein Data Bank (PDB), there are

160

four pullulanases with known protein structures showing high identity with BnPUL,

161

including B. acidopullulyticus pullulanase (64 % identity, PDB ID: 2WAN), Anoxybacillus

162

sp. LM18-11 pullulanase (43 % identity, PDB ID: 3WDJ), B. subtilis strain 168 pullulanase

163

(40 % identity, PDB ID: 2E8Y), and Klebsiella pneumoniae pullulanase (28 % identity, PDB

164

ID: 2FH8) 12-14, 18. Although B. acidopullulyticus pullulanase and B. subtilis strain 168

165

pullulanase share high identities with BnPUL, their apo structures, without the ligand for the

8

ACS Paragon Plus Environment

Page 8 of 27

Page 9 of 27

Journal of Agricultural and Food Chemistry

166

oligosaccharide molecule, make it difficult to identify the functional sites associated with

167

substrate binding 12, 18. Thus, Anoxybacillus sp. LM18-11 pullulanase and K. pneumoniae

168

pullulanase were selected to align with BnPUL to find the active sites possibly involved in

169

ligand binding and molecular interactions (Figure 1a). There were several residues in the

170

active sites that consistently appeared in 3WDJ and 2FH8, while some residues were unique

171

to each of the two pullulanases. Based on the conservation of the residues lining the catalytic

172

pocket, we chose nine residues (W504, Y506, T585, M621, W650, S731, N734, D787, and

173

Y790) for TCSM. They were randomly divided into three groups: (A) W504, Y506, and

174

W650; (B) T585, S731, and N734; and (C) M621, D787, and Y790 (Figure 1b and 1c).

175

In order to minimize the library size of variants and improve screening efficiency, three

176

conserved amino acids, as well as degeneracy of codons, needed to be determined for the

177

employment of TCSM. The amino acid sequence of the WT BnPUL was analyzed using

178

BLAST (https://blast.ncbi.nlm.nih.gov/Blast.cgi) by aligning the first 100 homologous

179

pullulanases (Figure S2). Based on the frequency of amino acids at the nine sites in

180

homologous sequences and the corresponding codon degeneracy, we chose tryptophan (W),

181

tyrosine (Y), and asparagine (N) as the triplet codon for SM (Figure S1).

182

Construction and screening of TCSM pullulanase libraries

183

For the three groups, A, B, and C, comprising nine residues, the corresponding SM

184

libraries were constructed as libraries I, II, and III, respectively. Catalytically active variants

185

were rapidly identified using the red pullulan plate assay (Figure 2). Library III contained

186

more variants that exhibited enzymatic activity as compared to the other two libraries.

187

Moreover, all of the variants in the library I failed to form transparent zones. It was presumed

9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

188

that the three amino acids in group A were rigidly conserved, and their mutations had a

189

negative impact on the enzymatic reaction.

190

Next, 5 and 14 colonies showing obvious transparent zones in the plate assay were picked

191

from libraries II and III, respectively, for further sequencing and activity assay. The

192

enzymatic activity of the variant was assayed in the supernatant obtained after cell lysis. The

193

variants from library II only maintained about half of the enzymatic activity of the WT, while

194

the variants from library III possessed about 70–80 % of the enzymatic activity of the WT.

195

Because the variants in library III were either active or inactive, the three sites of group C

196

might not be highly conserved and related to enzymatic reactions (Figure S3). Among the

197

three residues from group C, a mutation in Y790 was supposed to generate inactive variants,

198

and D787 and M621 could be potential sites exerting positive effects on enzymatic activity

199

and catalytic efficiency.

200

SM at potential active sites

201

We performed SM at the two potential sites, D787 and M621. For the mutation at site 621,

202

three variants showing larger transparent zones in the plate compared to other variants were

203

fermented by shake flask fermentation. The three variants exhibited about 400 U/mL activity,

204

which was not very different from that of WT. In the case of mutation at site 787, ten variants

205

showing larger transparent zones in the plate compared to other variants were fermented by

206

shake flask fermentation. Then, their enzymatic activity was assayed in the supernatant

207

obtained after cell lysis. Four variants, D787S, D787N, D787F, and D787C, exhibited higher

208

enzymatic activity than that of the WT (Figure 3). Among them, D787C exhibited the highest

209

enzymatic activity of 684 U·mL–1, which was 1.5-fold higher than that of the WT.

10

ACS Paragon Plus Environment

Page 10 of 27

Page 11 of 27

Journal of Agricultural and Food Chemistry

210

SDS-PAGE analysis of the cell-free extracts after expression indicated that the mutation in

211

D787 did not have a remarkable effect on protein expression (Figure 4).

212

Additionally, the variants D787S, D787N, D787F, and D787C were purified by Ni-NTA

213

affinity chromatography to measure specific activity and Michaelis-Menten parameters. The

214

four variants all exhibited higher specific activity than that of the WT, and among them

215

D787C showed the highest specific activity of 547 U·mg–1, which was 1.5-fold higher than

216

that of the WT (Figure 3). Moreover, the values of kcat and kcat/Km of the four variants all

217

showed more than 1.4-fold increase, compared to those of the WT (Figure 5). Among them,

218

the variant D787C exhibited a 1.8-fold increase in kcat and 1.7-fold increase in kcat/Km relative

219

to the WT (Figure 5). Therefore, the catalytic efficiency of the enzyme was improved by a

220

mutation in the residue D787.

221

According to the requirements of industrial saccharification process involving pullulanase,

222

the optimal pH and temperature of pullulanase should fit the industrial working condition.

223

Therefore, the effects of pH and temperature on enzymatic activity were investigated for

224

these four variants and the WT. All the tested variants exhibited the highest activities at pH

225

4.5 and 60 °C (Figure 6). Among them, the variant D787C maintained a higher activity than

226

the WT when the temperature was above 60 °C. On the other hand, all four variants showed

227

higher acid resistance than the WT, especially the variant D787C, retaining 90 % activity at

228

pH 4.0. Thermostability and acid resistance are known to be critical for the application of

229

pullulanase in starch saccharification. As compared to pullulanase from B. acidopullulyticus

230

(BaPUL), which has been used for commercial application in the starch industry 12, the

231

variant D787C has 75 % relative activity at 70 °C relative to 65 °C, while the variant from

11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

232

BaPUL has at least 60 % relative activity 23. The enzymatic activity of variants generated

233

from BaPUL was measured at pH 5.0 24, while in an even more acidic environment, the

234

variant D787C was assayed at pH 4.5 and was found to exhibit 85–90 % relative activity at

235

pH 4.0–5.0. Hence, an improved pullulanase variant was obtained with a potential for future

236

industrial application 22.

237

With respect to the change in enzymatic properties of BnPUL through mutations in the

238

functional sites identified by TCSM, we further analyzed the role of D787 in

239

structure-function relationship. Corresponding to the active site 889 of 2FGZ, D787 of

240

BnPUL could be one of the pullulanase-specific binding sites, which recognizes the

241

non-reducing ends of the main-chain sugar residues 13. The enhanced specific activity and

242

catalytic efficiency of the enzyme further confirmed the importance of D787 in the catalytic

243

reaction. As shown in Figure 7, after replacing aspartic acid with cysteine at site 787, the

244

hydrogen bond between Y735 and D787 as well as between Y735 and G785 was broken,

245

while a new bond was formed between Y735 and Q743. Y735 and one of the catalytic triad,

246

D733, are located in the same loop. The alteration in hydrogen bonds between Y735 and

247

D787 as well as between Y735 and Q743 induced conformational changes in the adjacent

248

loop and α-helix structures, leading to alterations in the steric conformation of D733 and

249

consequent molecular interactions between catalytic sites and ligands.

250

The amino acid sequence of the BnPUL was then analyzed using the Disorder Prediction

251

Meta-Server (DisMeta). Except for the N- and C-terminus of the BnPUL predicted to be

252

disordered by DisMeta, D787 was relatively more flexible than the other residues lining the

253

catalytic pocket of the enzyme (Figure S4). However, after the substitution of aspartic acid

12

ACS Paragon Plus Environment

Page 12 of 27

Page 13 of 27

Journal of Agricultural and Food Chemistry

254

with cysteine, the loop region close to the mutation site exhibited lower flexibility, compared

255

with the WT (Figure S5). As already known, the flexibility of the enzyme affects its stability.

256

Therefore, the improved stability of the variant D787C may be caused by the increased

257

rigidity. Based on the sequence alignment of homologous pullulanases (1,422 sequences) that

258

displayed more than 30% identity to the WT BnPUL (Figure S6), the conserved residues

259

corresponding to the site 787 in BnPUL are serine (S) and asparagine (N), but not

260

phenylalanine (F) and cysteine (C), suggesting that the obtained variant with improved

261

properties has not yet been generated by natural evolution.

262

In this work, we identified the active sites lining the catalytic pocket of BnPUL based on

263

the TCSM approach. Using red pullulan plate assay, two non-conserved sites, D787 and

264

M621, were discovered to have a positive effect on enzymatic activity. The improved

265

variants were then successfully obtained by further SM at these two sites. Among them,

266

D787S, D787N, D787F, and D787C showed enhanced specific activities and catalytic

267

efficiencies compared with the WT. Furthermore, these variants exhibited a broadened

268

temperature and pH profile. Hence, an improved enzyme variant exhibiting enhanced

269

performance in terms of both activity and stability was obtained, which is critical for the

270

industrial application of pullulanase.

271 272

ASSOCIATED CONTENT

273

Supporting Information

274 275

Construction of recombinant plasmids; Sequence alignment of B. naganoensis pullulanase; enzymatic activity of the WT and TCSM variants from libraries II and III; disorder region

13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

276

and secondary structure prediction of the WT and the variant D787C; and the list of primers.

277

ACKNOWLEDGMENT

278 279

We would like to thank Editage (www.editage.cn) for English language editing. Funding Resources

280

Financial support from the National Natural Science Foundation of China (NSFC)

281

(21336009, 21676120, 31872891), the Natural Science Foundation of Jiangsu Province

282

(BK20151124), the 111 Project (111-2-06), the High-end Foreign Experts Recruitment

283

Program (GDT20183200136), the Program for Advanced Talents within Six Industries of

284

Jiangsu Province (2015-NY-007), the National Program for Support of Top-notch Young

285

Professionals, the Fundamental Research Funds for the Central Universities (JUSRP51504),

286

the Project Funded by the Priority Academic Program Development of Jiangsu Higher

287

Education Institutions, the Jiangsu province "Collaborative Innovation Center for Advanced

288

Industrial Fermentation" industry development program, the Postgraduate Research &

289

Practice Innovation Program of Jiangsu Province, Top-notch Academic Programs Project of

290

Jiangsu Higher Education Institutions, and the National First-Class Discipline Program of

291

Light Industry Technology and Engineering (LITE2018-09) is greatly appreciated.

292

Conflict of Interest

293

The authors declare no competing financial interest.

294 295 296 297

14

ACS Paragon Plus Environment

Page 14 of 27

Page 15 of 27

Journal of Agricultural and Food Chemistry

298

REFERENCES

299

1. Henrissat, B.; Bairoch, A. New families in the classification of glycosyl hydrolases based

300

on amino acid sequence similarities. Biochem. J. 1993, 293, 781-788.

301

2. Bertoldo, C.; Antranikian, G. Starch-hydrolyzing enzymes from thermophilic archaea and

302

bacteria. Curr. Opin. Chem. Biol. 2002, 6, 151-160.

303

3. Reetz, M. T.; Bocola, M.; Carballeira, J. D.; Zha, D. X.; Vogel, A. Expanding the range of

304

substrate acceptance of enzymes: Combinatorial active-site saturation test. Angew. Chem. Int.

305

Edit. 2005, 44, 4192-4196.

306

4. Reetz, M. T.; Wang, L. W.; Bocola, M. Directed evolution of enantioselective enzymes:

307

Iterative cycles of CASTing for probing protein-sequence space. Angew. Chem. Int. Edit.

308

2006, 45, 1236-1241.

309

5. Reetz, M. T.; Wu, S. Greatly reduced amino acid alphabets in directed evolution: making

310

the right choice for saturation mutagenesis at homologous enzyme positions. Chem. Commun.

311

2008, 5499-5501.

312

6. Kille, S.; Acevedo-Rocha, C. G.; Parra, L. P.; Zhang, Z. G.; Opperman, D. J.; Reetz, M. T.;

313

Acevedo, J. P. Reducing Codon Redundancy and Screening Effort of Combinatorial Protein

314

Libraries Created by Saturation Mutagenesis. ACS Synth. Biol. 2013, 2, 83-92.

315

7. Parra, L. P.; Agudo, R.; Reetz, M. T. Directed evolution by using iterative saturation

316

mutagenesis based on multiresidue sites. Chembiochem 2013, 14, 2301-2309.

317

8. Sun, Z. T.; Lonsdale, R.; Kong, X. D.; Xu, J. H.; Zhou, J. H.; Reetz, M. T. Reshaping an

318

enzyme binding pocket for enhanced and inverted stereoselectivity: use of smallest amino

319

acid alphabets in directed evolution. Angew. Chem. Int. Edit. 2015, 54, 12410-12415.

15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

320

9. Sun, Z.; Lonsdale, R.; Ilie, A.; Li, G.; Zhou, J.; Reetz, M. T. Catalytic asymmetric

321

reduction of difficult-to-reduce ketones: Triple-code saturation mutagenesis of an alcohol

322

dehydrogenase. ACS Catal. 2016, 6, 1598-1605.

323

10. Sun, Z.; Lonsdale, R.; Wu, L.; Li, G.; Li, A.; Wang, J.; Zhou, J.; Reetz, M. T.

324

Structure-guided triple-code saturation mutagenesis: Efficient tuning of the stereoselectivity

325

of an epoxide hydrolase. ACS Catal. 2016, 6, 1590-1597.

326

11. Sun, Z.; Salas, P. T.; Siirola, E.; Lonsdale, R.; Reetz, M. T. Exploring productive

327

sequence space in directed evolution using binary patterning versus conventional mutagenesis

328

strategies. Bioresour. Bioprocess. 2016, 3, 44.

329

12. Turkenburg, J. P.; Brzozowski, A. M.; Svendsen, A.; Borchert, T. V.; Davies, G. J.;

330

Wilson, K. S. Structure of a pullulanase from Bacillus acidopullulyticus. Proteins 2009, 76,

331

516-9.

332

13. Mikami, B.; Iwamoto, H.; Malle, D.; Yoon, H. J.; Demirkan-Sarikaya, E.; Mezaki, Y.;

333

Katsuya, Y. Crystal structure of pullulanase: evidence for parallel binding of oligosaccharides

334

in the active site. J. Mol. Biol. 2006, 359, 690-707.

335

14. Xu, J.; Ren, F.; Huang, C. H.; Zheng, Y.; Zhen, J.; Sun, H.; Ko, T. P.; He, M.; Chen, C.

336

C.; Chan, H. C.; Guo, R. T.; Song, H.; Ma, Y. Functional and structural studies of pullulanase

337

from Anoxybacillus sp. LM18-11. Proteins 2014, 82, 1685-93.

338

15. Janecek, S.; Majzlova, K.; Svensson, B.; MacGregor, E. A. The starch-binding domain

339

family CBM41-An in silico analysis of evolutionary relationships. Proteins 2017, 85,

340

1480-1492.

341

16. Wang, X.; Nie, Y.; Mu, X.; Xu, Y.; Xiao, R. Disorder prediction-based construct

16

ACS Paragon Plus Environment

Page 16 of 27

Page 17 of 27

Journal of Agricultural and Food Chemistry

342

optimization improves activity and catalytic efficiency of Bacillus naganoensis pullulanase.

343

Sci. Rep. 2016, 6, 24574.

344

17. Studier, F. W. Protein production by auto-induction in high-density shaking cultures.

345

Protein Expres. Purif. 2005, 41, 207-234.

346

18. Malle, D.; Itoh, T.; Hashimoto, W.; Murata, K.; Utsumi, S.; Mikami, B. Overexpression,

347

purification and preliminary X-ray analysis of pullulanase from Bacillus subtilis strain 168.

348

Acta Crystallogr. Sect. F Struct. Biol. Cryst. Commun. 2006, 62, 381-4.

349

19. Swift, M. L. GraphPad prism, data analysis, and scientific graphing. J. Chem. Inf. Comp.

350

Sci. 1997, 37, 411-412.

351

20. Huang, Y. J.; Acton, T. B.; Montelione, G. T. DisMeta: a meta server for construct design

352

and optimization. Methods Mol. Biol. 2014, 1091, 3-16.

353

21. Wang, Q.-Y.; Xie, N.-Z.; Du, Q.-S.; Qin, Y.; Li, J.-X.; Meng, J.-Z.; Huang, R.-B. Active

354

hydrogen bond network (AHBN) and applications for improvement of thermal stability and

355

pH-sensitivity of pullulanase from Bacillus naganoensis. PLoS One 2017, 12.

356

22. Chang, M.; Chu, X.; Lv, J.; Li, Q.; Tian, J.; Wu, N. Improving the thermostability of

357

acidic pullulanase from Bacillus naganoensis by rational design. PLoS One 2016, 11,

358

e0165006.

359

23. Tomoko, M.; Akihiko, Y. Pullulanase variants and polynucleotides encoding same. U.S.

360

Patent 10,030,237, July 24, 2018.

361

24. Tomoko, M.; Suzanne, C.; Aki, T. Polypeptides having pullulanase activity suitable for

362

use in liquefaction. U.S. Patent 2,018,208,917, July 26, 2018.

17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

364

Figure Captions

365

Figure 1. (a) Sequence alignment of BnPUL with K. pneumoniae pullulanase (PDB ID:

366

2FGZ) and Anoxybacillus sp. LM18-11 pullulanase (PDB ID: 3WDH). (b, c) The spatial

367

location of the nine residues (W504, Y506, T585, M621, W650, S731, N734, D787, and

368

Y790) lining the catalytic pocket in the homology model of BnPUL that were identified

369

as mutational sites for TCSM. (a) The nine residues (W504, Y506, T585, M621, W650,

370

S731, N734, D787, and Y790) selected for TCSM are indicated by blue triangles. (b)

371

Backbone representations of the three groups: (A) W504, Y506, and W650; (B) T585, S731,

372

and N734; (C) M621, D787, and Y790 are shown in green, yellow, and blue, respectively.

373

The residues in the catalytic triad in the active site of the enzyme (D619, E648, and D733)

374

are indicated using red sticks. (c) Surface representation of BnPUL in the same orientation

375

and with the same color-coding as in panel (b). All images were obtained using the program

376

PyMOL.

377

Figure 2. Screening of variants with auto-induction of red pullulan plates. The substrate

378

is partially depolymerized pullulan, which is dyed with Procion Red MX-5B to an extent of

379

approximately one dye molecule per 30 sugar residues. When the enzyme variants are

380

incubated with red pullulan, the substrate is depolymerized to produce low-molecular-weight

381

dyed fragments, forming transparent zones and half-quantitatively determining the enzymatic

382

activity. (a) All of the variants in library I failed to form transparent zones. (b) Variants

383

forming transparent zones were selected for further activity measurement.

384

Figure 3. Enzymatic activity (purple bars, left scale) and specific activity (yellow bars,

385

right scale) of WT BnPUL and the variants. The values were measured in triplicates and

18

ACS Paragon Plus Environment

Page 18 of 27

Page 19 of 27

Journal of Agricultural and Food Chemistry

386

the standard deviations are indicated as error bars.

387

Figure 4. SDS-PAGE analysis of WT BnPUL and the variants. Lane M: the protein

388

marker.

389

Figure 5. kcat (purple bars, left scale) and kcat/Km (yellow bars, right scale) of WT

390

BnPUL and the variants. The values were measured in triplicates and the standard

391

deviations are indicated as error bars.

392

Figure 6. Effects of (a) temperature and (b) pH on the enzymatic activity of WT BnPUL

393

and the variants. Enzymatic activity was measured in 0.1 M sodium acetate buffer with pH

394

ranging from 3 to 5.5 or at temperatures ranging from 45 to 70 °C. The highest activity is

395

shown as 100 %. Error bars represent the standard deviations of three replicates.

396

Figure 7. The homology model of (a) WT BnPUL and (b) the variant D787C. The

397

enzyme backbone is represented as a cartoon in gray. The catalytic triad (D619, E648, and

398

D733) is represented by red lines. The substrate isomaltose is indicated by purple sticks. The

399

hydrogen bonds are indicated by yellow dashes. (a) The residue D787 is indicated by violet

400

lines. (b) The residue D787C is indicated by orange lines. All images were obtained using the

401

program PyMOL.

19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 1

20

ACS Paragon Plus Environment

Page 20 of 27

Page 21 of 27

Journal of Agricultural and Food Chemistry

Figure 2

21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 3

22

ACS Paragon Plus Environment

Page 22 of 27

Page 23 of 27

Journal of Agricultural and Food Chemistry

Figure 4

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 5

24

ACS Paragon Plus Environment

Page 24 of 27

Page 25 of 27

Journal of Agricultural and Food Chemistry

Figure 6

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 7

26

ACS Paragon Plus Environment

Page 26 of 27

Page 27 of 27

Journal of Agricultural and Food Chemistry

Table of Contents Graphic

27

ACS Paragon Plus Environment