<<

bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

1 Identification of NAD-dependent xylitol dehydrogenase from Gluconobacter

2 oxydans WSH-003

3

4 Li Liu1,2,4, Weizhu Zeng1, Guocheng Du1,3, Jian Chen1,2,4, Jingwen Zhou1,2,4*

5

6 1 School of Biotechnology and Key Laboratory of Industrial Biotechnology, Ministry

7 of Education, Jiangnan University, 1800 Lihu Road, Wuxi, Jiangsu 214122, China.

8 2 National Engineering Laboratory for Cereal Fermentation Technology, Jiangnan

9 University, 1800 Lihu Road, Wuxi, Jiangsu 214122, China.

10 3 The Key Laboratory of Carbohydrate Chemistry and Biotechnology, Ministry of

11 Education, Jiangnan University, 1800 Lihu Road, Wuxi, Jiangsu 214122, China.

12 4 Jiangsu Provisional Research Center for Bioactive Processing Technology,

13 Jiangnan University, 1800 Lihu Road, Wuxi, Jiangsu 214122, China.

14

15 * Corresponding author: Jingwen Zhou

16 Mailing address: School of Biotechnology, Jiangnan University, 1800 Lihu Road, Wuxi,

17 Jiangsu 214122, China

18 Phone: +86-510-85914317, Fax: +86-510-85914317

19 E-mail: [email protected].

20

21 bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

22 Abstract

23 Gluconobacter oxydans plays important role in conversion of D- to L-sorbose,

24 which is an essential intermediate for industrial-scale production of vitamin C. In the

25 fermentation process, some D-sorbitol could be converted to D-fructose and other

26 byproducts by uncertain dehydrogenases. Genome sequencing has revealed the

27 presence of diverse genes encoding dehydrogenases in G. oxydans. However, the

28 characteristics of most of these dehydrogenases remain unclear. Therefore, analyses of

29 these unknown dehydrogenases could be useful for identifying those related to the

30 production of D-fructose and other byproducts. Accordingly, dehydrogenases in G.

31 oxydans WSH-003, an industrial strain used for vitamin C production, were examined.

32 An NAD-dependent dehydrogenase, which was annotated as xylitol dehydrogenase 2,

33 was identified, codon-optimized, and expressed in Escherichia coli BL21 (DE3) cells.

+ 34 The exhibited high preference for NAD as the , while no activity

35 with NADP+, FAD, or PQQ was noted. Although this enzyme presented high

36 similarity with NAD-dependent xylitol dehydrogenase, it showed high activity to

37 catalyze D-sorbitol to D-fructose. Unlike the optimum temperature and pH for most of

38 the known NAD-dependent xylitol dehydrogenases (30°C–40°C and about 6–8,

39 respectively), those for the identified enzyme were 57°C and 12, respectively. The Km

40 and Vmax of the identified dehydrogenase towards L-sorbitol were 4.92 μM and 196.08

41 μM/min, respectively. Thus, xylitol dehydrogenase 2 can be useful for cofactor

42 NADH regeneration under alkaline conditions or its knockout can improve the

43 conversion ratio of D-sorbitol to L-sorbose.

44

45 Keywords: D-Fructose; D-sorbitol; cofactor regeneration; metabolic pathway.

46 bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

47 Importance

48 Production of L-sorbose from D-sorbitol by Gluconobacter oxydans is the first step

49 for industrial scale production of L-ascorbic acid. G. oxydans contains a lot of

50 different dehydrogenases, among which only several are responsible for the

51 conversion of D-sorbitol to L-sorbose, while others may responsible for the

52 accumulation of byproducts, thus decreased the yield of L-sorbose on D-sorbitol.

53 Therefore, a new xylitol dehydrogenase has been identified from 44 dehydrogenases

54 of G. oxydans. Optimum temperature and pH of the xylitol dehydrogenase are

55 different to most of the known ones. Knock-out of the dehydrogenase may improve

56 the conversion ratio of D-sorbitol to L-sorbose. Besides, the enzyme exhibits high

57 preference for NAD+ and have potential to be used for cofactor regeneration.

58

59

60 bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

61 Introduction

62 The genus Gluconobacter is a part of the group of acetic acid bacteria, which are

63 characterized by their ability to incompletely oxidize a wide range of carbohydrates

64 and alcohols (1). Gluconobacter strains have been successfully used for the industrial

65 production of food-related products, pharmaceuticals, and cosmetics, such as vitamin

66 C (2), miglitol (3), dihydroxyacetone (DHA) (4), and ketogluconates (5). In particular,

67 Gluconobacter oxydans has applications in the production of food additives and

68 sweeteners owing to its ability to synthesize flavoring ingredients from aromatic

69 alcohols, aliphatic alcohols, and 5-ketofructose (6, 7). Besides, G. oxydans ,

70 cell membranes, and whole cells are also used as sensor systems for the detection of

71 polyols, sugars, and alcohols (8-10). In recently years, some G. oxydans strains have

72 been employed for the production of enantiomeric pharmaceuticals and platform

73 compounds; for example, G. oxydans DSM2343 has been employed for the reduction

74 of various ketones used in pharmaceutical, agrochemical, and natural products (11),

75 Gluconobacter sp. JX-05 has been utilized for D-xylulose and xylitol production (12),

76 and G. oxydans DSM 2003 has been used for 3-hydroxypropionic acid production (13).

77 As all of these products are related to the dehydrogenases of G. oxydans, identification

78 of these enzymes in G. oxydans can expand the application of this bacterium.

79 Gluconobacter strains possess numerous dehydrogenases, some of which have

80 been identified, such as that could convert ethanol to

81 acetaldehyde (14, 15), NADP-dependent acetaldehyde dehydrogenase that could

82 convert acetaldehyde to acetate (16), PQQ-dependent glucose dehydrogenase that

83 could convert D-glucose to D-gluconate (14), gluconate dehydrogenase that could

84 convert D-gluconate to 2- or 5-ketogluconate (17), 2-ketogluconate dehydrogenase

85 that could convert 2-ketogluconate to 2,5-diketogluconate (14), D-sorbitol bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

86 dehydrogenase that could convert D-sorbitol to L-sorbose or fructose (14, 18-21),

87 sorbose/sorbosone dehydrogenase that could convert L-sorbose to L-sorbosone or

88 2-KLG (22, 23), dehydrogenase that could convert mannitol to fructose (24,

89 25), quinate dehydrogenase that could convert quinic acid to shikimic acid (26-28),

90 glycerol dehydrogenase that could convert glycerol to DHA (28), etc. In 2005, the

91 complete genome of G. oxydans 621H was sequenced (29), which revealed 75 open

92 reading frames (ORFs) that encode putative dehydrogenases/ of

93 unknown functions. Identification of the functions of these unknown

94 dehydrogenases/oxidoreductases is important to expand the application of G. oxydans.

95 For instance, carbonyl reductase (GoKR) from G. oxydans DSM2343 has been

96 employed for the reduction of various ketones (11), and membrane-bound alcohol

97 dehydrogenase (mADH) and membrane-bound aldehyde dehydrogenase from G.

98 oxydans DSM 2003 have been employed for 3-hydroxypropionic acid production

99 (13).

100 In G. oxydans, the central metabolic pathway, such as citrate cycle and

101 Embden-Meyerhof-Parnas pathway (EMP), is incomplete because of the absence of

102 some genes encoding succinate dehydrogenase, phosphofructokinase,

103 phosphotransacetylase, acetate kinase, succinyl-CoA synthetase, succinate

104 dehydrogenase, isocitratelyase, and malate synthase (30, 31), which may be the reason

105 for the low biomass of G. oxydans when cultured in rich medium. In a previous study,

106 sdhCDABE genes encoding succinate dehydrogenase and flavinylation factor SdhE,

107 ndh gene encoding a type II NADH dehydrogenase, and sucCD from

108 Gluconacetobacter diazotrophicus encoding succinyl-CoA synthetase were expressed

109 in G. oxyda ns to increase its biomass yield (31). However, G. oxydans biomass only

110 increased by 60%, suggesting the presence of some unknown bottleneck. Except for bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

111 the TCA cycle, all the genes were identified to encode enzymes involved in oxidative

112 pentose phosphate and Entner-Doudoroff (ED) pathways (29). The pentose phosphate

113 pathway is believed to be the most important route for phosphorylative breakdown of

114 sugars and polyols to CO2 and provide carbon skeleton. It has been speculated that G.

115 oxydans has the capability to uptake and channelize several polyols, sugars, and sugar

116 derivatives into the oxidative pentose phosphate pathway; however, the gene involved

117 in this process is still unknown. Hence, in the fermentation of sorbitol to sorbose for

118 vitamin C production, some sorbitol must get converted to fructose or other byproduct

119 to enter the pentose phosphate pathway for cell growth. Therefore, it is crucial to

120 balance and control the conversion of sorbitol to fructose for cell growth and sorbose

121 production.

122 Gene disruption and complementation experiments are often used to verify one

123 gene function. Some G. oxydans genes have been identify by using this method, such

124 as PQQ-dependent D- responsible for the oxidation of

125 1-(2-hydroxyethyl) amino-1-deoxy-D-sorbitol to 6-(2-hydroxyethyl)

126 amino-6-deoxy-L-sorbose, which is the precursor of an antidiabetic drug miglitol (3),

127 pyruvate decarboxylase that catalyzes the conversion of pyruvate to acetaldehyde by

128 decarboxylation (32), mADH, membrane-bound inositol dehydrogenase,

129 membrane-bound PQQ-dependent glucose dehydrogenase, etc. (33, 34). However,

130 some G. oxydans genes encoding dehydrogenases are necessary for cell growth, and

131 their knockout resulted in absence of growth (unpublished data). Besides, G. oxydans

132 comprises numerous dehydrogenases, some of which are isoenzymes, such as SldAB1

133 and SldAB2 of G. oxydans WSH-003 (35), or are often associated with a broad range

134 of substrates such as GoKR (11). Hence, the use of gene knockout strategy to identify

135 the functions of some dehydrogenases of G. oxydans, especially the numerous bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

136 unknown dehydrogenases of G. oxydans WSH-003, may not be appropriate. Therefore,

137 in the present study, we expressed numerous unknown dehydrogenases of G. oxydans

138 WSH-003 in Escherichia coli BL21 (DE3) cells and purified the products by one-step

139 affinity chromatography with Ni-NTA agarose column to identify their functions. The

140 results revealed a new xylitol dehydrogenase (NAD-dependent xylitol dehydrogenase

141 2) that could convert sorbitol to fructose. Kinetics analysis of the novel enzyme

142 revealed some unique traits that were quite different from the known xylitol

143 dehydrogenases. The optimum temperature and pH of the identified xylitol

144 dehydrogenase 2 was 57°C and 12, respectively. This novel enzyme provides new

145 insights into G. oxydans dehydrogenases and could have potential applications in

146 xylitol production.

147

148 Results

149 Gene expression and purification of the identified dehydrogenase

150 The selected dehydrogenase from G. oxydans WSH-003 was successfully

151 expressed and purified. Sequence analysis revealed that the purified enzyme,

152 annotated as xylitol dehydrogenase 2, contained a NAD(P)-binding motif and a

153 classical motif belonging to the short-chain dehydrogenase family.

154 SDS-PAGE analysis showed an expected single band with a molecular weight of

155 about 38 kDa (Fig. 1A), which was consistent with the calculated molecular mass

156 based on the deduced amino acid sequence (36.6 kDa). The optimum pH and

157 temperature for the purified xylitol dehydrogenase 2 were determined to be pH 12 (50

158 mM glycine-NaOH buffer) and 57°C, respectively (Fig. 1B, C), which are different

159 from those for known xylitol dehydrogenases.

160 bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

161 Identification of cofactor of xylitol dehydrogenase 2

162 In general, dehydrogenases require some cofactors as electron acceptor, such as

163 NAD(P), FAD/FMN, or PQQ. Most of the previously identified membrane

164 dehydrogenases from G. oxydans have been reported to utilize PQQ or FAD as the

165 cofactor. According to the prediction of transmembrane domains, xylitol

166 dehydrogenase 2 from G. oxydans WSH-003 was noted to lack transmembrane

167 domain. Therefore, the cofactor of the identified dehydrogenase was verified by using

168 the purified enzyme to catalyze reactions with different cofactors. The results showed

169 that xylitol dehydrogenase 2 was highly specific for NAD+, and no detectable enzyme

170 activity was observed with NADP+, FAD, or PQQ as the cofactor (Fig. 2).

171

172 Effect of EDTA and metal ions on enzyme activity

173 To determine the effects of chelator and metal ions on NAD-dependent xylitol

174 dehydrogenase 2, EDTA and various ions (0.5 mM Ca2+, Mg2+, Cu2+, Fe2+, Zn2+, Co2+,

175 Ni2+, Mn2+, Cr3+, and Fe3+) were respectively added to the reaction system. EDTA

176 elicited no obvious effect on NAD-dependent xylitol dehydrogenase 2, indicating that

177 the enzyme does not require chelator for its activity. However, the enzyme could be

178 activated by Zn2+, Co2+, and Mn2+, among which Zn2+ improved the enzyme activity

179 by 1.8 times. In contrast, Cu2+ could almost completely inhibit the activity of

180 NAD-dependent xylitol dehydrogenase 2 (Fig. 3), while the rest of the examined

181 metal ions had no obvious impact on the enzyme activity.

182

183 specificity and kinetic constants

184 In recent years, xylitol dehydrogenase has been used for the industrial production of

185 xylitol, and under enhanced NADH supply, NAD-dependent xylitol dehydrogenase bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

186 can reduce D-xylulose to desired xylitol. In the present study, substrate specificity

187 analysis of NAD-dependent xylitol dehydrogenase 2 revealed that the enzyme was

188 highly specific towards D-sorbitol and xylitol, but showed limited activity towards

189 D-mannitol, sorbose, and glycerol. Moreover, the enzyme showed no activity when

190 glucose, inositol, galactose, sorbitol, mannose, rhamnose, xylose, fructose, glucuronic

191 acid, glucolactone, 2-KLG, gluconic, propanol, isopropanol, methanol, and ethanol

192 were used as substrate (Fig. 4). To determine the kinetic constants, the initial

193 velocities of the enzyme were determined in glycine-NaOH buffer (pH 12) with

194 D-sorbitol (at increasing concentrations from 1 to 500 mM) under standard assay

195 conditions, and the Km and Vmax were noted to be 4.92 μM and 196.08 μM/min,

196 respectively.

197

198 Discussion

199 In this study, a xylitol dehydrogenase from 44 uncharacterized dehydrogenases

200 of G. oxydans WSH-003 was identified and characterized. This novel NAD-dependent

201 xylitol dehydrogenase 2 could convert D-sorbitol to D-fructose, indicating certain

202 correlation of this enzyme with pentose phosphate pathway (31). The optimum

203 temperature and pH for the identified xylitol dehydrogenase 2 revealed its unique

204 characteristics, when compared with some of the previously identified xylitol

205 dehydrogenases. It has been reported that D-fructose is the major byproduct formed

206 during the conversion of D-sorbitol to L-sorbose by G. oxydans in industrial-scale

207 vitamin C production (36), and that knockout of genes involved in D-fructose

208 production can further improve the conversion rate of D-sorbitol to L-sorbose. Owing

209 to its unique characteristics, the NAD-dependent xylitol dehydrogenase 2 identified in

210 the present study can be applied for the production of D-xylitol (12). To characterize bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

211 all the dehydrogenases of G. oxydans WSH-003, the enzymes were predicted and

212 heterologously overexpressed in E. coli BL21 (DE3) cells. Then, the expressed

213 dehydrogenases were purified by one-step affinity chromatography with Ni-NTA

214 agarose column. While most of the dehydrogenases with obvious expression levels in

215 E. coli showed no activities, NAD-dependent xylitol dehydrogenase 2 could

216 efficiently convert D-sorbitol to D-fructose.

217 Previous studies have indicated that majority of the numerous dehydrogenases in

218 G. oxydans are membrane-bound, PQQ- or FAD-dependent enzymes with more than

219 one subunit; for example, alcohol dehydrogenases have three subunits (37), aldehyde

220 dehydrogenases have two subunits (38), D-sorbitol dehydrogenases have one or three

221 subunits (18, 21, 39), and polyol dehydrogenase have two subunits (40). Most of the

222 cytoplasmic soluble polyol dehydrogenases are NADP-dependent with more than one

223 subunit; for instance, NADP-dependent D-sorbitol dehydrogenase have four subunits,

224 NADP-dependent D-sorbitol dehydrogenase have two subunits (41), and

225 NAD-dependent ribitol dehydrogenase have four subunits (42). However, the xylitol

226 dehydrogenase 2 identified in the present study was noted to be NAD-dependent with

227 only one subunit. The amino acid sequence of the NAD-dependent xylitol

228 dehydrogenase 2 showed similarity to those of the enzymes in the MDR superfamily.

229 However, the optimum pH and temperature for the oxidation activity of the

230 NAD-dependent xylitol dehydrogenase 2 were observed to be slightly higher than

231 those reported in earlier studies for the same reaction of xylitol dehydrogenases

232 isolated from different strains of G. oxydans (43). The reason for this variation in the

233 optimum pH and temperature for xylitol dehydrogenase activity could be owing to the

234 different source strains from which the enzymes were isolated.

235 With regard to the substrate specificity of xylitol dehydrogenases, xylitol bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

236 dehydrogenase from G. oxydans ATCC 621 has been noted to present higher catalytic

237 activity towards sorbitol and xylitol (44), whereas xylitol dehydrogenase from G.

238 thailandicus CGMCC1.3748 has been demonstrated to exhibit catalytic activity

239 towards xylitol, D-sorbitol, D-mannitol, and D-fructose (43). Besides, while most of

240 the known xylitol dehydrogenases have been reported to be dependent on cofactor

241 NAD+, an NADP+-dependent xylitol dehydrogenase has been found to increase

242 ethanol production from xylose in recombinant Saccharomyces cerevisiae though

243 protein engineering (45).

244 In most of the identified G. oxydans strains, glycolysis and citric acid cycle are

245 incomplete owing to the lack of phosphofructokinase and succinate dehydrogenase

246 (29), which is the main reason for the low biomass yield of G. oxydans, when

247 compared with other common bacteria, and a major limitation to the use of G. oxydans

248 whole cell biotransformation. It has been reported that pentose phosphate pathway

249 and ED pathway are the main catabolic routes for biomass and energy supply in

250 Gluconobacter strains (46). Despite its industrial application for several decades, the

251 metabolic pathways and regulatory mechanisms of Gluconobacter spp. are not yet

252 fully elucidated (47-49). To improve the biomass of G. oxydans, Krajewski et al.

253 knocked out the membrane-bound glucose dehydrogenase and soluble glucose

254 dehydrogenase, and improved the biomass by 271% (50). An understanding of the

255 mechanisms of catabolism of polyols, sugars, and sugar derivatives into the pentose

256 phosphate pathway is essential for increasing the biomass and efficiency of G.

257 oxydans strains. As D-sorbitol and L-sorbose cannot directly enter into the pentose

258 phosphate pathway, they must be catabolized via some intermediates. The

259 NAD-dependent xylitol dehydrogenase 2 identified in the present study can catalyze

260 D-sorbitol to D-fructose, which can directly enter the pentose phosphate pathway bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

261 through phosphorylation, suggesting that overexpression of this enzyme may increase

262 the biomass of G. oxydans by utilizing more D-sorbitol.

263 In conclusion, a novel NAD-dependent xylitol dehydrogenase 2 from G. oxydans

264 WSH-003 was identified in this study. Owing to its unique characteristics, such as

265 optimum pH and temperature, the identified dehydrogenase could be used in the

266 production of xylitol or fructose, or in regeneration of cofactor under specific

267 conditions. Although G. oxydans WSH-003 has been mutated from wild-type strain at

268 least 90 times by different methods with reliable records to improve L-sorbose

269 production and tolerance to saccharides and alditols such as L-sorbose and D-sorbitol,

270 generation of D-fructose as the byproduct of the strain could not be resolved.

271 However, knockout of xylitol dehydrogenase and similar dehydrogenases could

272 facilitate further increase in the yield of D-sorbitol to L-sorbose, which could be

273 important for the current industrial-scale production of vitamin C.

274

275 Materials and methods

276 Genes, plasmids, and strains

277 The vector pMD19-T Simple and pET-28a(+) were used for vector construction and

278 protein expression, respectively. E. coli JM109 cells were employed for plasmid

279 construction and E. coli BL21 (DE3) cells were used for protein expression. The

280 dehydrogenase gene (GenBank Accession No.: 29878874) was PCR-amplified from

281 the genomic DNA of G. oxydans WSH-003 using the primer pair

282 CCGGAATTCATGGCTCAAGCTTTGGTTCTGGAAC/CCGCTCGAGTCAGCCT

283 GGAAGCTTAATTTGTAGCTTC, purified, digested, and inserted into EcoRI/XhoI

284 sites of pET-28a(+) to obtain pET-28a-XDH. The recombinant plasmid pET-28a-XDH

285 was transformed into E. coli BL21 (DE3) cells for protein expression. All the bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

286 sequences were verified by Sanger sequencing (Sangon Biotech, Shanghai, China).

287 The transmembrane domains of the protein were predicted by using TMHMM

288 (http://www.cbs.dtu.dk/services/TMHMM/).

289

290 Gene expression and purification of dehydrogenase

291 The recombinant strain was cultured in 250-mL shake flasks containing 25 mL of

292 Terrific broth (TB) medium. After growth to log phase (OD600=0.6), the cells were

293 pre-cooled to 20°C. Then, 0.5 mM isopropyl-β-D-thiogalactopyranoside (IPTG) was

294 added to induce protein expression, and the cells were incubated at 20°C for another

295 16 h for protein expression.

296 Subsequently, the cells were collected by centrifugation at 5,000 rpm for 5 min,

297 washed twice with binding buffer (50 mM phosphate buffer), and lysed by sonication

298 at 4°C. The lysate was centrifuged for 20 min at 7,000 rpm at 4°C to obtain a clear

299 supernatant. The supernatant was passed through a 0.45-μm filter, and then applied to

300 a 5-mL nickel-charged Hi-Trap column pre-equilibrated with binding buffer. The

301 column was washed with 15 mL of binding buffer and then with washing buffer (50

302 mM phosphate buffer, 150 mM NaCl, and 50 mM imidazole; pH adjusted to 7.0) until

303 no more protein was eluted. The column was eluted with 20 mL of eluting buffer (50

304 mM phosphate buffer, 150 mM NaCl, and 500 mM imidazole), and the pH of the

305 eluent was adjusted to 7.0. The fractions were combined and dialyzed against dialysis

306 buffer (50 mM phosphate buffer).

307

308 Enzyme assay and identification of cofactor

309 The enzyme activity was measured by determining the increase in absorbance of

310 NADH at 340 nm. The reaction mixture contained 2 mM NAD+, 20 mM sorbitol, 50 bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

311 mM phosphate buffer (pH 12), and enzyme solution to a total volume of 200 μL. One

312 unit of enzyme activity was defined as the amount of enzyme catalyzing the formation

313 of 1 μmol of reduced NAD+ per minute at 30°C under the given conditions.

314

315 Effect of metal ions and EDTA

316 In order to determine the effect of the metal ions and the EDTA on the enzyme,

317 various metal ions (0.5 mM) and EDTA (5 mM) were added individually to the

318 reaction mixture. Relative activity was used to investigate, while the reaction mixture

319 without any additional treatment served as a control (100%).

320

321 Substrate specificity and determination of kinetic constants

322 Substrate specificity of the identified dehydrogenase was tested using 20 mM xylitol,

323 glucose, D-mannitol, inositol, sorbose, galactose, sorbitol, mannose, rhamnose, xylose,

324 fructose, glucuronic acid, glucolactone, 2-KLG, gluconic acid, propanol, glycerol,

325 inopropanol, methanol, and ethanol in the above-mentioned buffers. For kinetics

326 experiments, the substrate concentrations were varied between 1 and 500 mM and the

327 cofactor concentration was 2 mM.

328

329 Determination of optimum temperature and pH for the identified dehydrogenase

330 To determine the optimum pH, the enzyme activity was assessed in a pH range of

331 3–13 in the following buffers (50 mM): NaAc-HAc (pH 3.0–5.0), PBS (pH 5.0–9.0)

332 Tris-HCl (pH 9.0–10.0), and glycine-NaOH (pH 9.0–13). Similarly, the optimal

333 temperature for the identified dehydrogenase was determined by analyzing the

334 enzyme activity from 20°C to 70°C.

335 bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

336 Acknowledgements

337 This work was supported by grants from the National Natural Science Foundation of

338 China (Key Program, 31830068), the National Science Fund for Excellent Young

339 Scholars (21822806), the Fundamental Research Funds for the Central Universities

340 (JUSRP51701A), the National First-class Discipline Program of Light Industry

341 Technology and Engineering (LITE2018-08), the Distinguished Professor Project of

342 Jiangsu Province, and the 111 Project (111-2-06).

343

344 References

345 1. Deppenmeier U, Hoffmeister M, Prust C. 2002. Biochemistry and

346 biotechnological applications of Gluconobacter strains. Appl. Microbiol.

347 Biotechnol. 60:233-242.

348 2. Giridhar R, Srivastava AK. 2000. Model based constant feed fed-batch

349 L-sorbose production process for improvement in L-sorbose productivity.

350 Chem. Biochem. Eng. Q. 14:133-140.

351 3. Yang X-P, Wei L-J, Lin J-P, Yin B, Wei D-Z. 2008. A membrane-bound

352 PQQ-dependent dehydrogenase in Gluconobacter oxydans M5, responsible for

353 production of 6-(2-hydroxyethyl) amino -6-deoxy-L-sorbose. Appl. Microbiol.

354 Biotechnol. 74 5250-5253.

355 4. Poljungreed I, Boonyarattanakalin S. 2017. Dihydroxyacetone production

356 by Gluconobacter frateurii in a minimum medium using fed ‐ batch

357 fermentation. J. Appl. Chem. Biotechnol. 92:2635–2641.

358 5. Li K, Mao X, Liu L, Lin J, Sun M, Wei D, Yang S. 2016. Overexpression of

359 membrane-bound gluconate-2-dehydrogenase to enhance the production of

360 2-keto-D-gluconic acid by Gluconobacter oxydans. Microb. Cell Fact. 15:121 bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

361 6. Siemen A, Kosciow K, Schweiger P, Deppenmeier U. 2018. Production of

362 5-ketofructose from fructose or sucrose using genetically modified

363 Gluconobacter oxydans strains. Appl. Microbiol. Biotechnol. 102:1699-1710.

364 7. Rabenhorst J, DR., Gatfield I, DR., Hilmer J-M, DR. 2001-02-28 2001.

365 Natural, aliphatic and thiocarboxylic acids obtainable by fermentation and a

366 microorganism therefore EP1078990 patent EP1078990.

367 8. Bertokova A, Bertok T, Filip J, Tkac J. 2015. Gluconobacter sp cells for

368 manufacturing of effective electrochemical biosensors and biofuel cells. Chem.

369 Pap. 69:27-41.

370 9. Macauley S, McNeil B, Harvey LM. 2001. The genus Gluconobacter and its

371 applications in biotechnology. Crit. Rev. Biotechnol. 21:1-25.

372 10. Schenkmayerova A, Bertokova A, Sefcovicova J, Stefuca V, Bucko M,

373 Vikartovska A, Gemeiner P, Tkac J, Katrlik J. 2015. Whole-cell

374 Gluconobacter oxydans biosensor for 2-phenylethanol biooxidation

375 monitoring. Anal. Chim. Acta 854:140-144.

376 11. Chen R, Liu X, Wang J, Lin J, Wei D. 2015. Cloning, expression, and

377 characterization of an anti-Prelog stereospecific carbonyl reductase from

378 Gluconobacter oxydans DSM2343. Enzyme Microb. Technol. 70:18-27.

379 12. Qi X, Zhang H, Magocha TA, An Y, Yun J, Yang M, Xue Y, Liang S, Sun

380 W, Cao Z. 2017. Improved xylitol production by expressing a novel

381 D-arabitol dehydrogenase from isolated Gluconobacter sp. JX-05 and

382 co-biotransformation of whole cells. Bioresour. Technol. 235:50-58.

383 13. Zhu J, Xie J, Wei L, Lin J, Zhao L, Wei D. 2018. Identification of the

384 enzymes responsible for 3-hydroxypropionic acid formation and their use in

385 improving 3-hydroxypropionic acid production in Gluconobacter oxydans bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

386 DSM 2003. Bioresour. Technol. 265:328-333.

387 14. Matsushita K, Toyama H, Adachi O. 1994. Respiratory chains and

388 bioenergetics of Acetic Acid Bacteria. Adv. Microb. Physiol. 36:247-301.

389 15. Matsushita K, Yakushi T, Toyama H, Adachi O, Miyoshi H, Tagami E,

390 Sakamoto K. 1999. The quinohemoprotein alcohol dehydrogenase of

391 Gluconobacter suboxydans has ubiquinol oxidation activity at a site different

392 from the ubiquinone reduction site. Biochim. Biophys. Acta, Bioenerg.

393 1409:154-164.

394 16. Schweiger P, Volland S, Deppenmeier U. 2007. Overproduction and

395 characterization of two distinct aldehyde-oxidizing enzymes from

396 Gluconobacter oxydans 621H. J. Mol. Microbiol. Biotechnol. 13:147-155.

397 17. Shinagawa E, Matsushita K, Toyama H, Adachi O. 1999. Production of

398 5-keto-D-gluconate by acetic acid bacteria is catalyzed by pyrroloquinoline

399 quinone (PQQ)-dependent membrane-bound D-gluconate dehydrogenase J.

400 Mol. Catal. B: Enzym. 6:341-350.

401 18. Choi ES, Lee EH, Rhee SK. 1995. Purification of a membrane-bound

402 sorbitol dehydrogenase from Gluconobacter suboxydans. FEMS Microbiol.

403 Lett. 125:45-49.

404 19. Kim T-S, Patel SKS, Selvaraj C, Jung W-S, Pan C-H, Kang YC, Lee J-K.

405 2016. A highly efficient sorbitol dehydrogenase from Gluconobacter oxydans

406 G624 and improvement of its stability through immobilization. Sci. Rep. 6:

407 33438.

408 20. Shibata T, Ichikawa C, Matsuura M, Takata Y, Noguchi Y, Saito Y,

409 Yamashita M. 2000. Cloning of a gene for D-sorbitol dehydrogenase from

410 Gluconobacter oxydans G624 and expression of the gene in Pseudomonas bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

411 putida IFO3738. J. Biosci. Bioeng. 89:463-468.

412 21. Sugisawa T, Hoshino T. 2002. Purification and properties of

413 membrane-bound D-sorbitol dehydrogenase from Gluconobacter suboxydans

414 IFO 3255. Biosci., Biotechnol., Biochem. 66:57-64.

415 22. Asakura A, Hoshino T. 1999. Isolation and characterization of a new

416 quinoprotein dehydrogenase, L-sorbose/L-sorbosone dehydrogenase. Biosci.,

417 Biotechnol., Biochem. 63:46-53.

418 23. Saito Y, Ishii Y, Hayashi H, Imao Y, Akashi T, Yoshikawa K, Noguchi Y,

419 Soeda S, Yoshida M, Niwa M, Hosoda J, Shimomura K. 1997. Cloning of

420 genes coding for L-sorbose and L-sorbosone dehydrogenases from

421 Gluconobacter oxydans and microbial production of 2-keto-L-gulonate, a

422 precursor of L-ascorbic acid, in a recombinant G. oxydans strain. Appl.

423 Environ. Microbiol. 63:454-460.

424 24. Zahid N, Deppenmeier U. 2016. Role of mannitol dehydrogenases in

425 osmoprotection of Gluconobacter oxydans. Appl. Microbiol. Biotechnol.

426 100:9967-9978.

427 25. Adachi O, Toyama H, Matsushita K. 1999. Crystalline NADP-dependent

428 D-mannitol dehydrogenase from Gluconobacter suboxydans. Bioscience

429 Biotechnology and Biochemistry 63:402-407.

430 26. Vangnai AS, Promden W, De-Eknamkul W, Matsushita K, Toyama H.

431 2010. Molecular characterization and heterologous expression of quinate

432 dehydrogenase gene from Gluconobacter oxydans IFO3244. Biochemistry

433 (Moscow) 75:452-459.

434 27. Yakushi T, Komatsu K, Matsutani M, Kataoka N, Vangnai AS, Toyama H,

435 Adachi O, Matsushita K. 2018. Improved heterologous expression of the bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

436 membrane-bound quinoprotein quinate dehydrogenase from Gluconobacter

437 oxydans. Protein Expression Purif. 145:100-107.

438 28. Yakushi T, Terada Y, Ozaki S, Kataoka N, Akakabe Y, Adachi O,

439 Matsutani M, Matsushita K. 2018. Aldopentoses as new substrates for the

440 membrane-bound, pyrroloquinoline quinone-dependent glycerol (polyol)

441 dehydrogenase of Gluconobacter sp. Appl. Microbiol. Biotechnol.

442 102:3159-3171.

443 29. Prust C, Hoffmeister M, Liesegang H, Wiezer A, Fricke WF, Ehrenreich A,

444 Gottschalk G, Deppenmeier U. 2005. Complete genome sequence of the

445 acetic acid bacterium Gluconobacter oxydans. Nat. Biotechnol. 23:195-200.

446 30. Greenfield S, Claus GW. 1972. Nonfunctional tricarboxylic acid cycle and

447 the mechanism of glutamate biosynthesis in Acetobacter suboxydans. J.

448 Bacteriol. 112:1295-1301.

449 31. Kiefler I, Bringer S, Bott M. 2017. Metabolic engineering of Gluconobacter

450 oxydans 621H for increased biomass yield. Appl. Microbiol. Biotechnol.

451 101:5453-5467.

452 32. Peters B, Junker A, Brauer K, Mühlthaler B, Kostner D, Mientus M,

453 Liebl W, Ehrenreich A. 2013. Deletion of pyruvate decarboxylase by a new

454 method for efficient markerless gene deletions in Gluconobacter oxydans.

455 Appl. Microbiol. Biotechnol. 97:2521-2530.

456 33. Peters B, Mientus M, Kostner D, Junker A, Liebl W, Ehrenreich A. 2013.

457 Characterization of membrane-bound dehydrogenases from Gluconobacter

458 oxydans 621H via whole-cell activity assays using multideletion strains. Appl.

459 Microbiol. Biotechnol. 97:6397-6412.

460 34. Mientus M, Kostner D, Peters B, Liebl W, Ehrenreich A. 2017. bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

461 Characterization of membrane-bound dehydrogenases of Gluconobacter

462 oxydans 621H using a new system for their functional expression. Appl.

463 Microbiol. Biotechnol. 101:3189-3200.

464 35. Gao L, Zhou J, Liu J, Du G, Chen J. 2012. Draft genome sequence of

465 Gluconobacter oxydans WSH-003, a strain that is extremely tolerant of

466 saccharides and alditols. J. Bacteriol. 194:4455-4456.

467 36. Macauley-Patrick S, McNeil B, Harvey LM. 2005. By-product formation in

468 the D-sorbitol to L-sorbose biotransformation by Gluconobacter suboxydans

469 ATCC 621 in batch and continuous cultures. Process Biochem. 40:2113-2122.

470 37. Adachi O, Tayama K, Shinagawa E, Matsushita K, Ameyama M. 1978.

471 Purification and characterization of particulate alcohol dehydrogenase from

472 Gluconobacter suboxydans. Agric. Biol. Chem. 42:2045-2056.

473 38. Adachi O, Tayama K, Shinagawa E, Matsushita K, Ameyama M. 1980.

474 Purification and characterization of membrane-bound aldehyde dehydrogenase

475 from Gluconobacter suboxydans. Agricultural and Biological Chemistry

476 44:503-515.

477 39. Shinagawa E, Matsushita K, Adachi O, Ameyama M. 1982. Purification

478 and characterization of D-sorbitol dehydrogenase from membrane of

479 Gluconobacter suboxydans var. α. Agric. Biol. Chem. 46:135-141.

480 40. Matsushita K, Fujii Y, Ano Y, Toyama H, Shinjoh M, Tomiyama N,

481 Miyazaki T, Sugisawa T, Hoshino T, Adachi O. 2003. 5-keto-D-gluconate

482 production is catalyzed by a quinoprotein glycerol dehydrogenase, major

483 polyol dehydrogenase, in Gluconobacter species. Appl. Environ. Microbiol.

484 69:1959-1966.

485 41. Adachi O, Ano Y, Moonmangmee D, Shinagawa E, Toyama H, Theeragool bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

486 G, Lotong N, Matsushita K. 1999. Crystallization and properties of

487 NADPH-dependent L-sorbose reductase from Gluconobacter melanogenus

488 IFO 3294. Biosci., Biotechnol., Biochem. 63:2137-2143.

489 42. Adachi O, Fujii Y, Ano Y, Moonmangmee D, Toyama H, Shinagawa E,

490 Theeragool G, Lotong N, Matsushita K. 2001. Membrane-bound sugar

491 alcohol dehydrogenase in acetic acid bacteria catalyzes L-ribulose formation

492 and NAD-dependent ribitol dehydrogenase is independent of the oxidative

493 fermentation. Biosci., Biotechnol., Biochem. 65:115-125.

494 43. Zhang H, Yun J, Zabed H, Yang M, Zhang G, Qi Y, Guo Q, Qi X. 2018.

495 Production of xylitol by expressing xylitol dehydrogenase and alcohol

496 dehydrogenase from Gluconobacter thailandicus and co-biotransformation of

497 whole cells. Bioresour. Technol. 257:223-228.

498 44. Sugiyama M, Suzuki S, Tonouchi N, Yokozeki K. 2003. Cloning of the

499 xylitol dehydrogenase gene from Gluconobacter oxydans and improved

500 production of xylitol from D-arabitol. Biosci., Biotechnol., Biochem.

501 67:584-591.

502 45. Matsushika A, Watanabe S, Kodaki T, Makino K, Inoue H, Murakami K,

503 Takimura O, Sawayama S. 2008. Expression of protein engineered

504 NADP+-dependent xylitol dehydrogenase increases ethanol production from

505 xylose in recombinant Saccharomyces cerevisiae. Appl. Microbiol. Biotechnol.

506 81:243-255.

507 46. De Muynck C, Pereira CSS, Naessens M, Parmentier S, Soetaert W,

508 Vandamme EJ. 2007. The genus Gluconobacter oxydans: Comprehensive

509 overview of biochemistry and biotechnological applications. Crit. Rev.

510 Biotechnol. 27:147-171. bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

511 47. Klasen R, Bringer-Meyer S, Sahm H. 1992. Incapability of Gluconobacter

512 oxydans to produce tartaric acid. Biotechnol. Bioeng. 40:183-186.

513 48. Rauch B, Pahlke J, Schweiger P, Deppenmeier U. 2010. Characterization of

514 enzymes involved in the central metabolism of Gluconobacter oxydans. Appl.

515 Microbiol. Biotechnol. 88:711-718.

516 49. Gupta A, Qazi GN, Verma V. 1997. Transposon induced mutation in

517 Gluconobacter oxydans with special reference to its direct-glucose oxidation

518 metabolism. FEMS Microbiol. Lett. 147:181-188.

519 50. Krajewski V, Simić P, Mouncey NJ, Bringer S, Sahm H, Bott M. 2010.

520 Metabolic engineering of Gluconobacter oxydans for improved growth rate

521 and growth yield on glucose by elimination of gluconate formation. Appl.

522 Environ. Microbiol. 76:4369–4376.

523

524

525

526

527

528

529

530

531

532

533

534

535 bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.

536 Figure legends

537

538 Fig. 1 Optimum pH and temperature for xylitol dehydrogenase 2

539 SDS-PAGE of the identified xylitol dehydrogenase 2 purified from E. coli BL21 (DE3)

540 cells containing pET-28a-XDH. Lane 1: E. coli BL21 containing pET-28a after

541 induction for 16 h at 20°C. Lane 2: Recombinant strain E. coli BL21 containing

542 pET-28a-XDH after induction for 16 h at 20°C. Lane 3: Purified recombinant enzyme.

543 Lane M: Molecular mass markers. (B) Effect of pH on the activity of purified xylitol

544 dehydrogenase 2. (C) Effect of temperature on the activity of purified xylitol

545 dehydrogenase 2.

546

547 Fig. 2 Determination of cofactor of xylitol dehydrogenase 2

548 Catalytic reaction of purified xylitol dehydrogenase 2 (A) without cofactor, (B) with

549 NAD+, (C) with NADP+, (E) with FAD, and (E) with PQQ.

550

551 Fig. 3 Effect of metal ions on the activity of NAD-dependent xylitol

552 dehydrogenase 2

553 Relative activities of the enzyme in the presence of various metal ions, when

554 compared with the control without metal ions.

555

556 Fig. 4 Substrate specificity of NAD-dependent xylitol dehydrogenase 2

557 Relative enzyme activity towards (A) xylitol, (B) glucose, (C) D-mannitol, (E)

558 sorbose, (F) galactose, (G) sorbitol, (H) mannose, (I) rhamnose, (J) xylose, (K)

559 fructose, (L) glucuronic acid, (M) glucolactone, (N) 2-KLG, (O) gluconic acid, (P)

560 propanol, (Q) glycerol, (R) inopropanol, (S) methanol, and (T) ethanol. bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license. bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license. bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license. bioRxiv preprint doi: https://doi.org/10.1101/634238; this version posted May 10, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC 4.0 International license.