Characterization and thermal denaturation kinetic analysis of

Oct 2, 2018 - Stenotrophomonas maltophilia HS1 exhibits L-amino acid ester hydrolase (SmAEH) activity, which can synthesize dipeptides such as Ile–T...
1 downloads 0 Views 884KB Size
Subscriber access provided by University of Sunderland

Biotechnology and Biological Transformations

Characterization and thermal denaturation kinetic analysis of recombinant L-amino acid ester hydrolase from Stenotrophomonas maltophilia Md Saddam Hossain, Takahiro Tanaka, Junji Hayashi, Kazuyoshi Takagi, Yoichi Takeda, and Mamoru Wakayama J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b04573 • Publication Date (Web): 02 Oct 2018 Downloaded from http://pubs.acs.org on October 3, 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 41

Journal of Agricultural and Food Chemistry

1

1

Characterization and thermal denaturation kinetic analysis of recombinant

2

L

3

Md Saddam Hossaina, Takahiro Tanakaa, Junji Hayashia, Kazuyoshi Takagib, Yoichi Takedaa,

4

Mamoru Wakayamaa,*

5 6 7

a

-amino acid ester hydrolase from Stenotrophomonas maltophilia

Dept of Biotechnology, College of Life Sciences, Ritsumeikan University, Kusatsu, Shiga

525-8577, Japan b

Dept of Applied Chemistry, College of Life Sciences, Ritsumeikan University, Kusatsu,

8

Shiga 525-8577, Japan

9

*Corresponding author: E-mail: [email protected]

10

TEL: 077-561-276

11 12 13 14 15 16 17 18 19 20 21 22 23 24

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

2

25

Abstract

26

Stenotrophomonas maltophilia HS1 exhibits L-amino acid ester hydrolase (SmAEH) activity,

27

which can synthesize dipeptides such as Ile–Trp, Val–Gly, and Trp–His from the

28

corresponding amino acid methyl esters and amino acids. The gene encoding SmAEH was

29

cloned and expressed in Escherichia coli, and was purified and characterized. SmAEH shared

30

77% sequence identity with a known amino acid ester hydrolase (AEH) from Xanthomonas

31

citri, which belongs to a class of β-lactam antibiotic acylases. The thermal stability of

32

SmAEH was evaluated using various mathematical models to assess its industrial potential.

33

First-order kinetics provided the best description for the inactivation of the enzyme over a

34

temperature range of 35°C to 50°C. Decimal reduction time ranged from 212.76 to 3.44 min,

35

with a z value of 8.06°C, and the deactivation energy was 204.1 kJmol−1.

36 37

Keywords: Stenotrophomonas maltophilia HS1; L-amino acid ester hydrolase; Cloning;

38

Purification; Inactivation kinetics

39 40 41 42 43 44 45 46 47 48 49

ACS Paragon Plus Environment

Page 2 of 41

Page 3 of 41

Journal of Agricultural and Food Chemistry

3

50

Introduction

51

In recent years, the use of peptides has led to major progress in medicine and the chemical

52

industry, and they have gradually replaced classical drug approaches for regulating many

53

physiological processes1. Dipeptides and tripeptides, in particular, have attracted considerable

54

attention because of the possibility of their oral administration and the ease of performing

55

structural, molecular, and quantitative functional characterization2. The isolation and

56

characterization of new biologically active peptides are an ongoing endeavor and will

57

undoubtedly continue in the future3. For example, γ-Glu–Val–Gly is found in natural food

58

products, and received increased attention as a novel food additive, because it has a strong

59

taste of kokumi. Kokumi enhances the five basic tastes, particularly sweet, salty, and umami,

60

and modifies the thickness and richness of food4. Several food-derived anti-oxidant peptides

61

such as glutathione (γ-Glu–Cys–Gly), carnosine (β-Ala–His), and His–Pro have recently

62

received increased attention as novel food additives1. Val–Tyr, Ile–Trp, Ile–Tyr, and Lys–Trp

63

are strong anti-hypertensive peptides and are derived from hydrolysates of fish-meat,

64

seaweeds, etc5. Because of this growing interest, it is equally important to develop methods

65

for the synthesis of dipeptides on a commercial scale. Consequently, significant research

66

efforts have been expended to synthesize peptides using chemical or enzymatic approaches6.

67

Development of cost-effective methods for synthesizing di- and tripeptides will be crucial in

68

food science, since such small peptides have the ability to be absorbed through the

69

gastrointestinal tract and reach target organs via the circulatory system1.

70

L-Amino

acid ester hydrolases (AEHs) are well known for their application to the

71

biocatalytic synthesis of semisynthetic β-lactam antibiotics since 19727. Only α-amino acid

72

derivatives act as substrates, and owing to its high preference for esters over amides, the

73

enzyme has been formally named α-amino acid ester hydrolase (EC 3.1.1.43)8. AEHs were

74

never used for peptide synthesis until 2001 because, in addition to having only a few reports

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

4 75

on such enzymes to that date, no data was available on their activity with respect to peptide

76

synthesis and hydrolysis9. However, several recent studies on AEHs that successively and

77

effectively produced dipeptides have revealed that they can synthesize oligo-peptides as

78

well10–11. Owing to their preference for esters, it is conceivable that greater product

79

accumulation could be achieved in synthesis reactions using these enzymes rather than with

80

L

81

reports on natural microbial strains that produce this enzyme have been published10. By

82

contrast, however, reports on recombinant AEHs are limited because the attempts made to

83

clone an aeh gene have been unsuccessful12–13. To date, recombinant AEHs from

84

Xanthomonas citri14, A. turbidans15, and X. campestris pv. campestris16 have been

85

characterized, but there remains great interest in further expanding the number of

86

characterized species isoforms in this class of enzymes to enable improvement in their

87

properties by combinatorial or data-driven protein engineering techniques. The cloning of aeh

88

genes and the overproduction of their products would be particularly worthwhile since their

89

expression level in their natural hosts is low, varying from 0.5% to 3% of total cellular

90

protein17–18.

-amino acid ligase (Lal)5. Since the initial study of AEH from Acetobacter turbidans8, many

91

Previously, we screened microorganisms exhibiting AEH activity that can synthesize

92

dipeptides such as Ile–Trp from Ile–OMe and Trp and isolated several bacteria with this

93

property. Among the isolates, Stenotrophomonas maltophilia HS1 exhibited higher AEH

94

activity than the other isolates. The fermentation conditions for AEH production were

95

investigated18, but AEH production was insufficient to produce a material for further

96

investigation, such as conducting detailed kinetic studies with highly purified protein or

97

assessing its application to dipeptide synthesis. In this study, we thus aimed to clone the aeh

98

gene from S. maltophilia HS1 and to characterize its recombinant AEH protein, which could

99

potentially be useful for the synthesis of the dipeptide Ile–Trp, an important anti-hypertensive

ACS Paragon Plus Environment

Page 4 of 41

Page 5 of 41

Journal of Agricultural and Food Chemistry

5 100

agent. A BLAST search revealed that the 16s RNA sequence of wild-type S. maltophilia HS1

101

had 99% sequence identity with S. maltophilia R551-3. Furthermore, an analysis of the NCBI

102

protein database revealed that the sequence of a putative X-pro dipeptidyl-peptidase domain

103

protein (Smal_2367) of S. maltophilia R551-3 is 80% identical to a known AEH from X. citri.

104

All of the key catalytic residues are conserved, including the catalytic triad, the carboxylate

105

cluster, and the oxyanion hole (Figure S1(a), Supporting information). On the basis of these

106

observations, the gene corresponding to Smal_2367 was cloned from the genomic DNA of S.

107

maltophilia HS1, and the encoded protein was expressed in Escherichia coli. Since this

108

protein exhibited properties of an AEH, the gene encoding it was named aehS. We succeeded

109

in producing active AEH from S. maltophilia HS1 (SmAEH) in E. coli and characterized it.

110

AEH is useful for producing value-added products16–17; however, its proper utilization in

111

industrial applications requires engineering to facilitate optimization, design simulation, and

112

control of industrial projects. The low thermal stability of AEH is the main obstacle to its

113

industrial application19. Therefore, it is important to understand its thermal denaturation

114

behavior in a properly controlled reaction to achieve maximum production efficiency; this

115

includes evaluating the mutant enzymes from the view point of thermal stability. In the

116

present work, we aimed at elucidating the thermal denaturation kinetics of recombinant

117

SmAEH in detail. Various mathematical models that can predict enzyme residual activity as a

118

function of time were analyzed statistically, and the kinetic model that best explains the

119

thermal inactivation of SmAEH was selected. This model will facilitate inferences on the

120

enzyme’s behavior during heat treatment, thereby improving procedure precision and the

121

quality of the final products.

122

123

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 41

6

124

Materials and Methods

125

Materials

126

L-Isoleucine

127

methyl ester (L-Trp–OMe), L-tryptophan (L-Trp), L-glycine (L-Gly), and L-histidine (L-His)

128

were purchased from Wako Pure Chemicals (Tokyo, Japan). Isoleucyl–tryptophan (Ile–Trp),

129

valine–glycine (Val–Gly), and tryptophan–histidine (Trp–His) were from Watanabe

130

Chemical Industry (Hiroshima, Japan). Diethylaminoethyl (DEAE)-Cellufine and DEAE-

131

Sepharose were from JNC (Tokyo, Japan). The oligonucleotides for cloning of the aehS and

132

DNA sequences were obtained from FASMAC (Kanagawa, Japan). The chemicals used in

133

the DNA manipulation procedures were purchased from TaKaRa (Kusatsu, Japan) and used

134

as recommended by the manufacturer. The other reagents were of analytical grade.

135

Bacterial strains and plasmids

136

S. maltophilia HS1 was grown at 30°C as described previously18. Chromosomal DNA was

137

obtained using phenol/chloroform extraction and used as the template for PCR. The pET21a

138

plasmid (Novagen) was used as an expression vector, and E. coli XL10-gold (Agilent

139

Technologies) and E. coli Rosetta-gami B(DE3) (Novagen) were used for cloning and

140

recombinant protein expression, respectively.

141

Cloning and protein expression

142

Primers were initially designed based on the gene information of Smal_2367 (Figure S1(b),

143

Supporting information) in order to isolate the complete open reading frame of the aehS gene

144

from the chromosomal DNA of S. maltophilia HS1. The oligonucleotide primers used were

145

5 ′ -ATGCTTTCCTGCCGGGAGGGGTTGTC-3 ′

146

TCAGTACACCGGCAGATCGATGTAGCTGG-3′ (reverse). To construct the expression

methyl ester (L-Ile–OMe), L-valine methyl ester (L-Val–OMe), L-tryptophan

ACS Paragon Plus Environment

(forward)

and

5 ′ -

Page 7 of 41

Journal of Agricultural and Food Chemistry

7

147

vector, PCR was used with the 5′-GGAATTCCATATGCTTTCCTGCCGGGAGG-3′ as

148

the forward primer and 5′-CCCAAGCTTTCAGTACACCGGCAGATCGAT-3′ as the

149

reverse. These primers contained Nde I and Hind III restriction sites (underlined),

150

respectively. After DNA denaturation, the amplifications were carried out in 30 cycles of 10 s

151

at 98°C, 30 s at 55°C, and 1 min at 68°C using KOD-plus neo as DNA polymerase according

152

to the supplier’s instruction (TOYOBO, Osaka, Japan). The amplified DNA fragment was

153

digested with Nde I and Hind III and ligated into pET21a linearized with the same restriction

154

enzymes. This plasmid, pETSM, was used as an expression vector. The newly constructed

155

plasmids were amplified in E. coli XL10-gold and then used to transform in E. coli Rosetta-

156

gami B(DE3) cells for expression. The transformants were cultured at 30°C in 0.5 L of Luria

157

Bertani medium containing 100 μg/mL ampicillin while shaking at 100 rpm. When the

158

optical density at 600 nm reached about 0.8, isopropyl-β-D-thiogalactopyranoside was added

159

to the culture medium at a final concentration of 0.2 mM. The cultures were incubated further

160

at 20°C for 12 h, and cells were harvested by centrifugation (10,000 g for 10 min at 4°C).

161

Protein purification

162

The harvested cells were suspended in 20 mM Tris-HCl buffer (pH 8.0; buffer I) and treated

163

with 0.5 mg/mL lysozyme at 4°C for 30 min. The cells were disrupted by sonication, and cell

164

debris was removed by centrifugation. The resulting supernatant was dialyzed against buffer I

165

and then loaded on a DEAE-Cellufine column pre-equilibrated with buffer I. SmAEH was

166

eluted in the nonbinding fraction from the column and was dialyzed against buffer I. The

167

SmAEH was then applied to a DEAE-Sepharose fast flow column equilibrated with buffer I

168

and eluted with linear gradient from 0 to 150 mM NaCl. The SmAEH-containing fractions

169

were collected and dialyzed against buffer I, and the purified enzyme was analyzed using

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

8 170

sodium dodecyl sulfate–polyacrylamide gel electrophoresis (SDS–PAGE) and 10.5% gel.

171

Enzymes were concentrated by ultrafiltration and stored at 4°C.

172

Enzyme assays and determination of kinetic constants

173

The amino-group transferring activity of AEH was routinely assayed at 20°C in 100 mM

174

borate buffer (pH 10.5), 50 mM L-Ile–OMe as the acyl donor, and 100 mM Trp (pH 10.5) as

175

the acyl acceptor. The formation of the Ile–Trp product was monitored by high-performance

176

liquid chromatography (HPLC) with a COSMOSIL 5C18-MS-II column (4.6 × 250 mm;

177

Nacalai Tesque, Inc., Kyoto, Japan) using the method of Tanaka et al17. Before analysis, the

178

samples (5 µL) were quenched and diluted 100-fold by the addition of a derivatizing-

179

fluorescent compound in the form of 10 mM Boc-Cys-OH and 20 mM phthalaldehyde in 400

180

mM borate buffer (pH 9.0). Five µL of this mixture was then analyzed by HPLC. One unit of

181

enzyme activity was defined as the amount required to produce 1 µmol Ile–Trp per minute

182

under standard assay conditions. To determine the kinetics parameters, the enzyme was

183

incubated with various concentrations of acyl donors in the range of 0–60 mM L-Ile–OMe, L-

184

Val–OMe, or L-Trp–OMe, and acyl acceptors at 0–120 mM Trp and 0–250 or 0–250 mM His.

185

The rates of dipeptide synthesis were estimated by measuring product formation by HPLC.

186

Data were fitted using nonlinear regression using a Michaelis–Menten kinetics model, and

187

the calculated kinetic parameters are presented with their standard deviations.

188

MS analysis of SmAEH reaction products

189

The reaction mixture was diluted 50-fold with Milli-Q water for matrix-assisted laser

190

desorption ionization time-of-flight mass spectrometry (MALDI-TOF MS) analysis. One µL

191

of the sample was mixed with 1 µL of matrix solution on a plate, and the plate was dried and

192

loaded onto the MALDI-TOF MS (autoflex speed, Bruker Daltonics K.K., Kanagawa, Japan).

ACS Paragon Plus Environment

Page 8 of 41

Page 9 of 41

Journal of Agricultural and Food Chemistry

9 193

The matrix (2,5-dihydroxybenzoic acid; DHB) was dissolved in 0.1% trifluoroacetic

194

acid:acetonitrile = 1:2.

195

Heat inactivation experiments

196

Thermal inactivation tests were performed in sealed tubes containing 1 mL of SmAEH. The

197

tubes were immersed in a thermostatically controlled water bath (Personal-11 SM; Taitec

198

Corporation, Japan), and the enzyme solutions were incubated at 35°C, 40°C, 45°C, or 50°C.

199

After incubation, the tubes were immediately immersed in an ice bath, and SmAEH activity

200

was measured using a mixture of 100 mM borate buffer (pH 10.5), 50 mM L-Ile–OMe, and

201

100 mM Trp (pH 10.5). The activity after 1 min of heating was considered to be the initial

202

activity (A0) for the purpose of calculating the residual activities in order to eliminate the

203

effects of warming and to ensure evaluation of an isothermal process20. Assays were

204

performed in triplicate. Residual activity data with respect to time at different temperatures

205

were fitted to several models (Table 1) using nonlinear regression (Statistica 13.0; StatSoft,

206

Inc., Tulsa, UK).

207

Kinetic models of enzyme inactivation

208

Kinetic models commonly used to describe the behavior of an enzyme during heat treatment

209

were employed in this study (Table 1). In the equations, A represents the SmAEH activity at

210

time t (min), A0 is the initial activity, and k (min−1) is the inactivation rate constant at a given

211

temperature. Generally, first-order kinetics (Eq. 1) have been reported to describe the thermal

212

inactivation of enzymes21. The temperature-dependent parameters for the first-order model

213

are given by the Arrhenius equation, which describes the relationship of the thermal

214

inactivation rate (k) and temperature and can be expressed algebraically:

215

ln ( k ) = ln ( A ) −

Ea , RT

ACS Paragon Plus Environment

(6)

Journal of Agricultural and Food Chemistry

Page 10 of 41

10 216

where A is the Arrhenius constant, Ea is the activation energy, R is the universal gas constant

217

(8.31 Jmol−1K−1), and T is the absolute temperature. However, we also considered other

218

kinetics models; for equations 2 and 3, it has been suggested that the residual activity of an

219

enzyme could be the sum of two exponential decays, signifying the presence of a mixture of

220

enzymes with different heat sensitivities and/or catalytic properties. These models were tested

221

to check whether the SmAEH solution was composed of labile and resistant heat fractions,

222

and kL and kR are the first-order reaction rate constants for the labile and resistant heat

223

fractions, respectively. Moreover, coefficient “a” represents the active fraction of the

224

thermolabile fraction in terms of the total activity, and Ar implies the nonzero activity of an

225

enzyme after prolonged heating and assumes the presence of an extremely heat-resistant

226

fraction22. Descriptions of nth-order decay (Eq. 4) for thermal inactivation of enzymes are

227

also not exceptional23. The Weibull distribution (Eq. 5) is based on the assumption that, under

228

the conditions examined, the momentary rate of thermal sensitivity to heat is only a factor of

229

the transient heating intensity and residual activity but not of the rate at which the residual

230

activity has been reached24. The values n and b characterize this model and represent the

231

shape of the distribution curve and its scaling, respectively. Its temperature dependence can

232

be expressed by a log-logistic equation (Eq. 7)25:

233

b(T) = loge{1+ exp[k′(T −Tc )]},

234

where T marks the temperature level of the inactivation’s onset and K ' is the slope of b(T ) . c

235

Comparison of kinetic models

236

In general, the statistical standards for the assessment of models are the coefficient of

237

determination(r2), chi-square (χ2), and standard error mean (SEM)26. χ2 was used to compare

238

models of enzyme thermal inactivation20,25 and was determined as follows:

ACS Paragon Plus Environment

(7)

Page 11 of 41

Journal of Agricultural and Food Chemistry

11

χ2 =

239

240

∑ ( ameasured − a predicted ) m− p

2

.

(8)

The SEM was determined using 2  (a  ∑ measured − a predicted ) . SEM =    m  

241

(9)

242

Each model was evaluated by an optimization criterion based on their sum of squares and the

243

parameters, where m is the number of observations and p is the number of parameters.

244

To discriminate between kinetic models, statistical and physical criteria should be considered.

245

The model with the lowest χ2 and SEM and highest r2 for the residual enzyme activity is the

246

best choice from a statistical perspective20. Estimation of a negative value for parameters is a

247

criterion to reject the inactivation model27.

248

Thermodynamic analysis

249

Activation energy (Ea) can be estimated from the Arrhenius equation (Eq. 6). Using this value

250

of Ea, the inactivation enthalpy (∆H#) for each temperature was calculated by

251

252

253

∆H # = Ea − RT .

(10)

The free energy of inactivation (∆G#) can be determined using the following expression:

 kh  ∆G # = − RT ln  ,  kBT 

(11)

254

where h (6.6262 × 10−34 Js) is the Planck’s constant, KB (1.3806 × 10−23 JK−1) is the

255

Boltzmann’s constant, and k (s−1) is the inactivation rate constant for each temperature.

256

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 41

12 257

From equations 10 and 11, it is possible to calculate the inactivation entropy (∆S#) as follows:

∆S # =

258

∆H # − ∆G # . T

(12)

259

Nucleotide sequence accession number

260

The nucleotide sequence for SmAEH has been submitted to the DNA Database of Japan and

261

has been assigned the accession no. LC319754.

262

263

Results and discussion

264

Since 1974, amino acid sequence information on only two AEHs that have been

265

experimentally characterized has been available in the NCBI database, although they have

266

been repeatedly assessed for use in biocatalysts. To clone the gene encoding the AEH of S.

267

maltophilia (SmAEH), we designed oligonucleotide primers based on the first six amino

268

acids of the putative X-pro dipeptidyl-peptidase domain protein (NCBI GenBank ID

269

ACF52070.1) and used them with the chromosomal DNA of S. maltophilia to produce a PCR

270

product of 2.0-kb. Sequencing indicated that this product shared approximately 94%

271

sequence identity with a putative X-pro dipeptidyl-peptidase domain protein of S. maltophilia

272

R551-3 encoded with a polypeptide of 645 amino acids with a calculated molecular mass of

273

72,019 Da. Database searches for homologous proteins using position-specific iterated PSI-

274

BLAST28 found that the deduced amino acid sequence of SmAEH was homologous with

275

several protein sequences, the majority of which are from genome-sequencing projects and

276

have an unidentified function; however, protein information retrieved from the UniProt

277

database suggested that most of the homologous proteins were either AEHs or glutaryl-7-

278

ACA acylases (Table S1, Supporting information). Moreover, these homologous proteins

279

encoded the consensus sequence GXSYXG, where X is a nonconserved amino acid,

ACS Paragon Plus Environment

Page 13 of 41

Journal of Agricultural and Food Chemistry

13 280

suggesting that the proteins are serine hydrolases. Surprisingly, this motif was also conserved

281

in SmAEH, suggesting that it functions as a serine hydrolase; this was further supported by

282

the observation that SmAEH is inactivated by PMSF (data not shown). Until a few years ago,

283

it was believed that all lipases and carboxyl esterases contain the consensus sequence motif

284

GXSXG around the active site serine29. However, some exceptions to this have recently been

285

described.

286

To verify the identity of the ester hydrolase gene, the recombinant plasmid pETSM

287

was introduced into E. coli RGB DE3 competent cells for heterologous expression of the

288

enzyme, and cell extracts were analyzed with an AEH activity assay. The extracts had a high

289

level of AEH activity compared with those produced from cells containing the pET21(a)

290

plasmid. The recombinant SmAEH was purified by chromatography on DEAE-Cellufine,

291

followed by DEAE-Sepharose resins. The specific activity of SmAEH increased at each step

292

to approximately 11.3 U/mg, representing a seven-fold purification (Table 2). A clear single

293

band corresponding to a protein with a molecular mass of around 72 kDa was observed by

294

SDS–PAGE (Figure 1). This is consistent with the masses of purified native AEHs from A.

295

turbidans and the Xanthomonas family, which are 74 and 72 kDa protein, respectively14,15.

296

The activity of SmAEH was measured at different temperatures and pH using

297

isoleucine methyl ester and tryptophan as substrates. The optimal pH for SmAEH activity

298

was 10.5, and the optimal temperature was 20°C (Figures S2a and S2b, Supporting

299

information). The activity of the enzyme was almost stable between pH 9.0 and 11.0, and it

300

did not exhibit any high requirements for any metals for activity (Figure S2c, Supporting

301

information). A slight inhibition was observed in the presence of Fe3+ and Co2+ ions, with a

302

decrease of around 30% of its original activity, and one of 20% in the presence of K+. A

303

similar study showed that AEH from Elizabethkingia sp. has no significant requirement for

304

metal ions, and its maximum activity occurred at pH 9.0 and a temperature of 25°C17. In

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

14 305

another study, AEH from Empedobacter brevis ATCC14234 exhibited the highest activity at

306

pH 8.5 and a temperature of 30°C30. However, Bacillus mycoides AEH has a maximum

307

activity at pH 8.0 and a temperature of 45°C9.

308

HPLC analyses of SmAEH reaction mixtures (Figure S3, Supporting information)

309

showed that products were formed in a time-dependent manner. In particular, Ile–Trp (peak

310

no. 3) was detected after 30 min and increased further after 1 h. Moreover, in comparisons

311

with standards and zero-min HPLC peaks, time-dependent reaction studied showed a new

312

peak with a retention time of 34 min, presumably reflecting presence of Ile-Trp; however, no

313

peaks corresponding to the formation of Ile–Ile–OMe were observed. For checking

314

hydrolysis of the product by the enzyme, enzymatic reaction using standard Ile-Trp as a

315

substrate instead of Ile-OMe was performed under the same condition as synthetic reaction.

316

No breakdown of di-peptide, Ile-Trp was observed after 2-h reaction time. The reaction

317

mixture was also analyzed by MALDI-TOF MS, and a peak was detected at m/z 318.41,

318

which corresponds to [Ile–Trp+H]+, indicating the formation of Ile–Trp under these reaction

319

condition (Figure S4, Supporting information). Furthermore, other peaks such as those at m/z

320

132.18, 205.23, 227.23, 243.33, 281.36, 396.64, and 544.77, which correspond to [Ile+H]+,

321

[Trp+H]+, [Trp+Na]+, [Trp+K]+, [Ile–Ile–OMe+Na]+, [Ile–Ile–Ile+K]+, and [Ile–Ile–Ile–

322

Trp+H]+, were also detected, suggesting the formation of peptides other than Ile–Trp as well

323

as simple hydrolysis of the donor substrate. HPLC analysis estimated that the Ile–Trp

324

concentration reached approximately 5 mM from 50 mM L-Ile–OMe and 100 mM Trp and

325

that of dipeptide Val–Gly and Trp–His reached 8 and 4 mM, each of 50 mM acyl donor and

326

200 mM acyl acceptor, respectively. Yokozeki and Hara found that AEH from E. brevis

327

ATCC produced 12 and 4 mM Val–Gln and Ile–Gln, each of 100 mM acyl donor and 200

328

mM acyl acceptor, respectively10, and in another study, AEH from Sphingobacterium

329

siyangensis synthesized 10 and 2 mM Val–Gln and Ile–Gln30, where each of 100 mM acyl

ACS Paragon Plus Environment

Page 14 of 41

Page 15 of 41

Journal of Agricultural and Food Chemistry

15 330

donor and 200 mM acyl acceptor, respectively. To study the substrate specificity of

331

recombinant SmAEH, steady-state kinetic parameters were measured for a variety of

332

substrates, and Michaelis–Menten kinetics were analyzed. The Km values for the acyl donors

333

L-Ile–OMe, L-Val–OMe,

334

respectively, and those for the acyl acceptors L-Trp, L-Gly, and L-His were 35, 65, and 55

335

mM (Table 3), respectively. However, AEHs from X. citri and A. turbidans have Km of 90

336

and 7 mM for D-phenylglycine methyl ester and 1.8 and 0.34 mM for cephalexin,

337

respectively14–15. In our study, SmAEH exhibited variability in substrate concentration

338

requirements for different acyl donors and acceptors, which supports a previous finding that

339

the nature of the acyl group has a large influence on AEH enzyme kinetics15. Furthermore,

340

the hydrolysis capability of SmAEH for various acyl donors was investigated using a

341

previously described colorimetric method17. The hydrolytic activity assay for SmAEH

342

showed that it had a relatively strong substrate preference for hydrophobic amino acid methyl

343

esters (Figure S5, Supporting information). Unfortunately, owing to the lack of standard

344

dipeptides in our laboratory, the preferences of SmAEH for acceptor substrates could not be

345

evaluated in this study. Examination of the acceptor preference of SmAEH will be performed

346

in the future, since it will provide useful information for the synthesis of other valyl

347

derivatives.

and L-Trp–OMe were determined to be 10.4, 22, and 27 mM,

348

The most closely related proteins whose activities have been described are the AEHs

349

from X. citri (XcAEH) and A. turbidans (AtAEH); the activity of these enzymes against α-

350

amino acid esters has been reported, and their crystal structures have been determined14,31. In

351

our study, the deduced amino acid sequence of SmAEH shared 77% and 58% sequence

352

identity with XcAEH (PDB ID 1MPX) and AtAEH (PDB ID 2V9B). Homology modeling of

353

SmAEH was performed based on XcAEH as a template and using the Swiss model server32,

354

and the model was validated using an Internal Coordinate Mechanics (ICM)-based computer

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

16 355

program via a method adapted from Hossain et al33. The model structure resembled the

356

experimental structure (Figure S6, Supporting information), and the catalytic triad residues of

357

SmAEH were Ser-180, Asp-261, and His-294, which correspond to residues Ser-174, Asp-

358

255, and His-288 of XcAEH. A hypothesis on the AEH enzyme family by Barends et al.14

359

suggests that a carboxylate group cluster near the catalytic triad is responsible for AEH

360

specificity toward Nα-free amino acid esters. We aligned the sequences of experimentally

361

characterized AEHs with SmAEH and one unreported AEH from Elizabethkingia sp.

362

identified in our lab and found that this also supports the hypothesis of the presence of a

363

carboxylate cluster group (Figure 2a). We judged that the catalytic site of SmAEH is

364

structurally related to XcAEH and AtAEH, and hence, we superimposed the catalytic triad

365

and carboxylate cluster residues and found that the root-mean-square deviation values of

366

SmAEH were 0.07 Å with XcAEH and 0.20 Å with AtAEH (Figure 2b and 2c), suggesting

367

that the AEH catalytic sites are structurally unique.

368

From both scientific and technological perspectives, inactivation kinetics are essential

369

to enzyme characterization. AEHs catalyze the synthesis and hydrolysis of α-amino β-lactam

370

antibiotics and are used as alternatives to the more commonly used penicillin G acylase14–16.

371

Semi-synthetic β-lactam antibiotics make up approximately 65% of the total world market for

372

antibiotics of $15 billion34. Moreover, AEHs are of particular interest for use in the one-pot

373

synthesis of semi-synthetic antibiotics such as ampicillin directly from Penicillin G, thus

374

eliminating intermediate states16; however, inactivation behavior studies of AEH would be

375

essential for its incorporation into such industrial processes. This would lead to lower costs

376

and more environmentally friendly production of semi-synthetic antibiotics and peptides

377

compared to traditional chemical synthesis and existing enzymatic synthesis. In 2010, Blum

378

and Bommarius16 tried to determine the inactivation pattern of AEH from X. campestris pv.

379

campestris ATCC33913, but their data did not follow first-order kinetics, and this issue

ACS Paragon Plus Environment

Page 16 of 41

Page 17 of 41

Journal of Agricultural and Food Chemistry

17 380

remains unresolved. Therefore, from an industrial perspective on the AEH enzyme, it is

381

essential to assess the effects of heat treatment on residual activity without performing

382

numerous trial runs; with this in mind, several mathematical equations have been proposed to

383

explain the behavior of SmAEH during thermal degradation. A graphical representation of

384

the residual activities during heat inactivation fitted to a first-order equation is presented in

385

Figure 3, showing a satisfactory fit of the experimental data to the predicted curves. Table 4

386

presents the results from five inactivation kinetic models that were tested to fit the

387

experimental data for heat treatments of SmAEH. Among these, fractional conversion (Eq. 3)

388

produced negative parameters, which are physical criteria to reject the models, since the

389

residual activity for SmAEH must be positive. And, in the analysis of two fractions (Eq. 2),

390

the data showed equal parameters and were rejected as well. The nth order model (Eq. 4)

391

produced low r2 and high SEM and χ2 values compared with the other models, so it was also

392

rejected. In general, enzyme inactivation is characterized by a first-order model. However, in

393

recent years, the appropriateness and utility of the Weibull distribution have been widely

394

discussed35, and it has been proposed to explain the thermal inactivation of both enzymes and

395

microorganisms27. Thus, the Weibull distribution was tested for its suitability for fitting the

396

experimental data from SmAEH inactivation. Both first-order and Weibull distribution

397

models produced good fits for the data (Table 4) with very similar r2, χ2, and SEM values.

398

The r2 values for the Weibull distribution model ranged from 0.980 to 0.997, and those for

399

the first-order model ranged from 0.989 to 0.998. The χ2 values for the first-order model were

400

lower than those for the Weibull pattern, ranging from 0.022 to 0.00027 and from 0.0482 to

401

0.00137, respectively. The SEM ranged from 0.026 to 0.00077 for the first-order model and

402

from 0.027 to 0.00224 for the Weibull. The data obtained for the analyzed conditions were

403

very similar for both models, making it difficult to choose between them. Therefore, the

404

temperature-dependent parameters obtained for the first-order (k values) and Weibull

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

18 405

distribution (b values) models were fitted to the Arrhenius (Eq. 6) and log-logistic (Eq. 7)

406

equations, as described elsewhere20. The r2 for the Arrhenius equation gave a good fit to the

407

data, with a value of 0.984 (Figure 4). For the Weibull model parameters, the log-logistic

408

equation produced a comparatively low r2 value of 0.74 (data not shown). This comparatively

409

poor fit of the Weibull model to the experimental data supports the conclusion that the first-

410

order equation is the best model to explain the thermal inactivation of SmAEH.

411

Thermal inactivation concepts (D and z values) are generally used to represent a first-

412

order reaction. The decimal reduction time (D value) is the time needed for a 10-fold

413

reduction of the initial activity at a given temperature, and it is obtained by plotting activity

414

on a log scale against inactivation time. The z value is the temperature needed to reduce the D

415

value one log unit and is obtained by plotting D values on a log scale against the

416

corresponding temperatures20. The data obtained for the k, D, and z values for SmAEH for

417

heat treatments between 35°C and 50°C are shown in Table 5. In general, the rate constant

418

increased with higher heating temperatures while D values decreased, indicating a faster

419

inactivation at higher temperatures. D values ranged between 212.76 and 3.44 min at

420

temperatures between 35°C and 50°C, respectively. The effect of the temperature on the D

421

values is shown in Figure 5, and the calculated z value for the temperature range studied was

422

approximately 8.06°C.

423

In our study, data for the residual SmAEH activity best fit the first-order model and

424

yielded decent statistical criteria for the conditions evaluated. The values for r2, SEM, and χ2

425

were comparable to those calculated for model acceptance in other investigations20–21,27,35.

426

This is the first of such report on an AEH enzyme, although there are broad classes of

427

enzymes36,37, bio-active substances20, and peptides38 that obey first-order kinetics. Figure 3

428

signified data that SmAEH exhibits a single-step exponential behavior and provides

429

suggestions for a possible deactivation mechanism. Since alterations in pH do not strongly

ACS Paragon Plus Environment

Page 18 of 41

Page 19 of 41

Journal of Agricultural and Food Chemistry

19 430

interfere with the enzyme deactivation rate (data not shown), it is unlikely that inactivation is

431

due to reactions such as reshuffling or β-elimination of cysteine residues, which are

432

extremely pH dependent36, although SmAEH enzymes have four cysteine residues. The pH

433

range studied also ruled out inactivation from the hydrolysis of the peptide chain. In our study,

434

enzyme activity decreased as heating time increased. It is apparent in single step first-order

435

reaction that only one reason controlling overall inactivation rate leading to enzyme

436

denaturation37. In this study, the most likely cause of inactivation with increasing temperature

437

is the deamidation of the asparagine or glutamine residues. The sequence of SmAEH consists

438

of approximately 7% glutamine (4.5%) and asparagine (2.5%). This high percentage may be

439

one of the causes of enzyme inactivation at elevated temperatures. Furthermore, it has been

440

reported that a temperature increase of 15°C above the optimum resulted in a 75% reduction

441

in the half-life of an asparagine peptide39. These observations support the hypothesis that

442

deamidation is temperature sensitive and that proteins exposed to high temperatures are

443

especially prone to deamidation and subsequent enzyme inactivation40. Another study found

444

that, during an increase in temperature from 5°C to 65°C, peptide chain deamidation behaves

445

in accordance with the Arrhenius equation41. However, a direct biochemical analysis is

446

required to validate that deamidation actually occurs during the heat inactivation of SmAEH.

447

The temperature dependence of the k-values obtained for the first-order model was

448

fitted to the Arrhenius equation (Eq. 6), and the result is shown in Figure 4. The r2 value was

449

0.984, indicating that 98.4% of the total variation is explained by the linear regression. The

450

deactivation energy (Ea) can be considered to represent the energy that needs to be absorbed

451

or released for the molecules to be able to react42 and it can be estimated from the Arrhenius

452

equation. In general, the higher the Ea, the higher the energy barrier that must be crossed for

453

enzyme inactivation to occur; higher values therefore indicate increased stability. The

454

estimated Ea for SmAEH inactivation was 204.1 kJ mol−1, suggesting that energy needed to

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

20 455

be absorbed from the external medium to initiate inactivation at temperatures between 35°C

456

and 50°C; the data fitted well, which enabled the calculation of thermodynamic parameters

457

(Table 5) using transition state theory (Eqs. 10–12). ∆H# (enthalpy) increased, and ∆G# (free

458

energy) decreased with increasing temperature, suggesting that destabilization of SmAEH

459

occurred42. Positive ∆H# values are indicative of the endothermic character of the inactivation

460

process43. However, the ∆S# (entropy) values did not exhibit a continuous behavior,

461

suggesting that the increase of temperature destroyed ordered structures, leading to a fragile

462

state. ∆H# and ∆S# represent heat and entropy changes, respectively, and these two

463

parameters also provide a measure of the number of noncovalent bonds broken and the net

464

change in enzyme/solvent disorder associated with the formation of the transition state44.

465

Lals or proteases have also been reported as other enzymatic methods for synthesizing

466

dipeptides. Peptide synthesis using proteases has been extensively studied by protecting or

467

freeing either one or both N and C-terminal ends of acyl donors and acceptors45. Despite

468

numerous trials, proteases fail to meet industrial requirements in terms of productivity,

469

reaction rate, yield, proportion of acyl acceptors to acyl donors, and breadth of applicability

470

to a sufficiently wide range of peptides. In Lal-based methods, processes suffer from the low

471

accumulation and productivity of the target dipeptide5, reflecting technological nascence of

472

dipeptide formulations. By contrast, AEHs synthesize dipeptides efficiently and irreversibly

473

using Nα-free amino acid esters as acyl donors and a Nα- and Cα-free amino acids as an acyl

474

acceptors9,17. Although several reports on the chemical and enzymatic synthesis of dipeptides

475

are available, all of these processes have acknowledged limitations with respect to

476

incorporation of hydrophobic amino acids. Synthesis of highly hydrophobic dipeptides is a

477

difficult and challenging task because of their low solubility in both aqueous and organic

478

solvents. SmAEH in our study had optimum pH at 10.5, which makes it particularly

479

advantageous for dissolving hydrophobic amino acids, and the enzyme exhibited a wide

ACS Paragon Plus Environment

Page 20 of 41

Page 21 of 41

Journal of Agricultural and Food Chemistry

21 480

specificity toward hydrophobic substrates. This is because the alkaline buffer facilitated

481

dissolution of hydrophobic amino acids. Nothing has been published on the synthesis of the

482

dipeptide Ile–Trp, and this is the first report on the synthesis of this important anti-

483

hypertensive agent. Our study covered the characterization of SmAEH and the analysis of

484

inactivation kinetics, which should prove valuable for industrial applications of its bioproduct

485

synthesis capabilities. Our analysis will be useful for the future engineering of AEH enzymes.

486

Competing interests

487

The authors have declared that no competing interest exists.

488

Acknowledgments

489

We are grateful to Monbukagakusho, The Ministry of Education, Culture, Sports, Science

490

and Technology in Japan (MEXT), for financial support (MEXT scholarship).

491

Supporting information

492

Figure S1 Multiple sequence alignments and primer sequence information data

493

Table S1 Amino acid sequence similarities of the AEH of S. maltophilia HS1

494

Figure S2 Characterization of SmAEH

495

Figure S3 High-performance liquid chromatography (HPLC) profiles of SmAEH reactions

496

over time.

497

Figure S4 MALDI-TOF MS analysis of the SmAEH reaction

498

Figure S5 Substrate specificity of SmAEH for hydrolysis reaction

499

Figure S6 Homology model of SmAEH

500

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

22

501

References

502

(1)

Samaranayaka, A. G. P.; Li-Chan, E. C. Y. Food-Derived Peptidic Antioxidants: A

503

Review of Their Production, Assessment, and Potential Applications. J. Funct. Foods

504

2011, 3 (4), 229–254.

505

(2)

Devappa, R. K.; Makkar, H. P. S.; Becker, K. Nutritional, Biochemical, and

506

Pharmaceutical Potential of Proteins and Peptides from Jatropha. J. Agric. Food Chem.

507

2010, 58 (11), 6543–6555.

508

(3)

Fan, X.; Bai, L.; Zhu, L.; Yang, L.; Zhang, X. Marine Algae-Derived Bioactive

509

Peptides for Human Nutrition and Health. J. Agric. Food Chem. 2014, 62 (38), 9211–

510

9222.

511

(4)

Kuroda, M.; Kato, Y.; Yamazaki, J.; Kai, Y.; Mizukoshi, T.; Miyano, H.; Eto, Y.

512

Determination and Quantification of γ-Glutamyl-Valyl-Glycine in Commercial Fish

513

Sauces. J. Agric. Food Chem. 2012, 60 (29), 7291–7296.

514

(5)

and Perspectives. Appl. Microbiol. Biotechnol. 2008, 81 (1), 13–22.

515 516

(6)

(7)

Takahashi, T.; Yamazaki, Y.; Kato, K.; Isono, M. Enzymic Synthesis of Cephalosporins. J. Am. Chem. Soc. 1972, 94 (11), 4035–4037.

519 520

Tsuchiya, K.; Numata, K. Chemoenzymatic Synthesis of Polypeptides for Use as Functional and Structural Materials. Macromol. Biosci. 2017, 17 (11).

517 518

Yagasaki, M.; Hashimoto, S. Synthesis and Application of Dipeptides; Current Status

(8)

Takahashi, T.; Yamazaki, Y.; Kato, K. Substrate Specificity of an α-Amino Acid Ester

521

Hydrolase Produced by Acetobacter turbidans ATCC 9325. Biochem. J. 1974, 137 (3),

522

497–503.

523

(9)

Sugihara, A.; Shimada, Y.; Sugihara, S.; Nagao, T.; Watanabe, Y.; Tominaga, Y. A ACS Paragon Plus Environment

Page 22 of 41

Page 23 of 41

Journal of Agricultural and Food Chemistry

23 524

Novel Alpha-Amino-Acid Esterase from Bacillus mycoides Capable of Forming

525

Peptides of DD- and DL-Configurations. J. Biochem. 2001, 130 (1), 119–126.

526

(10)

Peptides from Unprotected Starting Materials. J. Biotechnol. 2005, 115 (2), 211–220.

527 528

Yokozeki, K.; Hara, S. A Novel and Efficient Enzymatic Method for the Production of

(11)

Hirao, Y.; Mihara, Y.; Kira, I.; Abe, I.; Yokozeki, K. Enzymatic Production of L-

529

Alanyl-L-Glutamine by Recombinant E. coli Expressing alpha;-Amino Acid Ester

530

Acyltransferase from Sphingobacterium siyangensis. Biosci. Biotechnol. Biochem.

531

2013, 77 (3), 7–10.

532

(12)

Nam, D. H.; Dewey, D. Y. Molecular Cloning of the Gene for Alpha-Acylamino Beta-

533

Lactam Acylhydrolase from Acetobacter turbidans by Immunochemical Detection

534

Method. Microbiol. Biotechnol. Lett. 1988, 16 (5), 363–368.

535

(13)

Pv. citri. Microbiology 1996, 142 (10), 2951–2957.

536 537

Alonso, J.; García, J. L. Proline Iminopeptidase Gene from Xanthomonas campestris

(14)

Barends, T. R. M.; Polderman-Tijmes, J. J.; Jekel, P. A.; Hensgens, C. M. H.; de Vries,

538

E. J.; Janssen, D. B.; Dijkstra, B. W. The Sequence and Crystal Structure of the α-

539

Amino Acid Ester Hydrolase from Xanthomonas citri Define a New Family of β-

540

Lactam Antibiotic Acylases. J. Biol. Chem. 2003, 278 (25), 23076–23084.

541

(15)

Polderman-Tijmes, J. J.; Jekel, P. A.; de Vries, E. J.; van Merode, A. E. J.; Floris, R.;

542

van der Laan, J.-M.; Sonke, T.; Janssen, D. B. Cloning, Sequence Analysis, and

543

Expression in Escherichia coli of the Gene Encoding an Alpha-Amino Acid Ester

544

Hydrolase from Acetobacter turbidans. Appl. Environ. Microbiol. 2002, 68 (1), 211–

545

218.

546

(16)

Blum, J. K.; Bommarius, A. S. Amino Ester Hydrolase from Xanthomonas campestris

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

24 547

Pv. campestris, ATCC 33913 for Enzymatic Synthesis of Ampicillin. J. Mol. Catal. B

548

Enzym. 2010, 67 (1–2), 21–28.

549

(17)

Tanaka, T.; Takagi, K.; Saddam, H. M.; Takeda, Y.; Wakayama, M. Purification and

550

Characterization of Elizabethkingia L-Amino Acid Esterase: An Enzyme Useful for

551

Enzymatic Synthesis of the Dipeptide, Valyl-Glycine. Appl. Biochem. Biotechnol.

552

2017, 183 (1), 362-373.

553

(18)

Hossain, M. S.; Tanaka, T.; Takagi, K.; Hayashi, J.; Wakayama, M. Purification and

554

Characterization of Stenotrophomonas maltophilia-Derived l-Amino Acid Ester

555

Hydrolase for Synthesizing Dipeptide, Isoleucyl-Tryptophan. 3 Biotech 2018, 8 (3),

556

173.

557

(19)

Blum, J. K.; Ricketts, M. D.; Bommarius, A. S. Improved Thermostability of AEH by

558

Combining B-FIT Analysis and Structure-Guided Consensus Method. J. Biotechnol.

559

2012, 160 (3–4), 214–221.

560

(20)

Sant’Anna, V.; Utpott, M.; Cladera-Olivera, F.; Brandelli, A. Kinetic Modeling of the

561

Thermal Inactivation of Bacteriocin-like Inhibitory Substance P34. J. Agric. Food

562

Chem. 2010, 58 (5), 3147–3152.

563

(21)

Daroit, D. J.; Sant’Anna, V.; Brandelli, A. Kinetic Stability Modelling of Keratinolytic

564

Protease P45: Influence of Temperature and Metal Ions. Appl. Biochem. Biotechnol.

565

2011, 165 (7–8), 1740–1753.

566

(22)

Knockaert, G.; Pulissery, S. K.; Lemmens, L.; Van Buggenhout, S.; Hendrickx, M.;

567

Van Loey, A. Carrot β-Carotene Degradation and Isomerization Kinetics during

568

Thermal Processing in the Presence of Oil. J. Agric. Food Chem. 2012, 60 (41),

569

10312–10319.

ACS Paragon Plus Environment

Page 24 of 41

Page 25 of 41

Journal of Agricultural and Food Chemistry

25 570

(23)

Asthana, N. S.; Kolah, A. K.; Vu, D. T.; Lira, C. T.; Miller, D. J. A Kinetic Model for

571

the Esterification of Lactic Acid and Its Oligomers. Ind. Eng. Chem. Res. 2006, 45 (15),

572

5251–5257.

573

(24)

1951, 18 (3), 293–297.

574 575

(25)

Corradini, M. G.; Peleg, M. A Model of Non-Isothermal Degradation of Nutrients, Pigments and Enzymes. J. Sci. Food Agric. 2004, 84 (3), 217–226.

576 577

Weibull, W. A Statistical Distribution Function of Wide Applicability. J. Appl. Mech.

(26)

Da Silva Malheiros, P.; Sant, V.; Brandelli, A.; de Melo Franco, B. D. G. Kinetic

578

Modeling of Thermal Inactivation of Antimicrobial Peptides Produced by

579

Lactobacillus sakei Subsp. sakei 2a. Thermochim. Acta 2015, 605, 95–99.

580

(27)

Mint Leaves. J. Food Eng. 2008, 85 (1), 147–153.

581 582

Shalini, G. R.; Shivhare, U. S.; Basu, S. Thermal Inactivation Kinetics of Peroxidase in

(28)

Altschul, S. F.; Madden, T. L.; Schäffer, A. A.; Zhang, J.; Zhang, Z.; Miller, W.;

583

Lipman, D. J. Gapped BLAST and PSI-BLAST: A New Generation of Protein

584

Database Search Programs. Nucleic Acids Res. 1997, 25 (17), 3389–3402.

585

(29)

Bornscheuer, U. T. Microbial Carboxyl Esterases: Classification, Properties and Application in Biocatalysis. FEMS Microbiol. Rev. 2002, 26 (1), 73–81.

586 587

(30) Abe, I.; Hara, S.; Yokozeki, K. Gene Cloning and Characterization of α-Amino Acid

588

Ester Acyl Transferase in Empedobacter brevis ATCC14234 and Sphingobacterium

589

siyangensis AJ2458. Biosci. Biotechnol. Biochem. 2011, 75 (11), 2087–2092.

590

(31)

Polderman-Tijmes, J. J.; Jekel, P. A.; Jeronimus-Stratingh, C. M.; Bruins, A. P.; van

591

der Laan, J.-M.; Sonke, T.; Janssen, D. B. Identification of the Catalytic Residues of α-

592

Amino Acid Ester Hydrolase from Acetobacter turbidans by Labeling and Site-

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

26

Directed Mutagenesis. J. Biol. Chem. 2002, 277 (32), 28474–28482.

593 594

(32)

Guex, N.; Peitsch, M. C. SWISS-MODEL and the Swiss-Pdb Viewer: An

595

Environment for Comparative Protein Modeling. Electrophoresis 1997, 18 (15), 2714–

596

2723.

597

(33)

Hossain, M. S.; Azad, A. K.; Chowdhury, P. A.; Wakayama, M. Computational

598

Identification and Characterization of a Promiscuous T-Cell Epitope on the

599

Extracellular Protein 85B of Mycobacterium Spp. for Peptide-Based Subunit Vaccine

600

Design. Biomed Res. Int. 2017, 2017, 1-14

601

(34)

Biotechnol. 2003, 61 (5–6), 385–392.

602 603

Elander, R. P. Industrial Production of β-Lactam Antibiotics. Appl. Microbiol.

(35)

Sant’Anna, V.; Corrêa, A. P. F.; Daroit, D. J.; Brandelli, A. Kinetic Modeling of

604

Thermal Inactivation of the Bacillus Sp. Protease P7. Bioprocess Biosyst. Eng. 2013,

605

36 (7), 993–998.

606

(36)

Schokker, E. P.; van Boekel, A. J. S. Kinetic Modeling of Enzyme Inactivation:

607

Kinetics of Heat Inactivation at 90−110° C of Extracellular Proteinase from

608

Pseudomonas fluorescens 22F. J. Agric. Food Chem. 1997, 45 (12), 4740–4747.

609

(37)

Anthon, G. E.; Sekine, Y.; Watanabe, N.; Barrett, D. M. Thermal Inactivation of

610

Pectin Methylesterase, Polygalacturonase, and Peroxidase in Tomato Juice. J. Agric.

611

Food Chem. 2002, 50 (21), 6153–6159.

612

(38)

Sant’Anna, V.; Cladera-Olivera, F.; Brandelli, A. Kinetic and Thermodynamic Study

613

of Thermal Inactivation of the Antimicrobial Peptide P34 in Milk. Food Chem. 2012,

614

130 (1), 84–89.

615

(39)

Geiger, T.; Clarke, S. Deamidation, Isomerization, and Racemization at Asparaginyl

ACS Paragon Plus Environment

Page 26 of 41

Page 27 of 41

Journal of Agricultural and Food Chemistry

27 616

and Aspartyl Residues in Peptides. Succinimide-Linked Reactions That Contribute to

617

Protein Degradation. J. Biol. Chem. 1987, 262 (2), 785–794.

618

(40)

Peptidoglutaminase. J. Agric. Food Chem. 1992, 40 (5), 719–723.

619 620

Hamada, J. S. Effects of Heat and Proteolysis on Deamidation of Food Proteins Using

(41)

Stratton, L. P.; Kelly, R. M.; Rowe, J.; Shively, J. E.; Smith, D. D.; Carpenter, J. F.;

621

Manning, M. C. Controlling Deamidation Rates in a Model Peptide: Effects of

622

Temperature, Peptide Concentration, and Additives. J. Pharm. Sci. 2001, 90 (12),

623

2141–2148.

624

(42)

Arcus, V. L.; Prentice, E. J.; Hobbs, J. K.; Mulholland, A. J.; Van der Kamp, M. W.;

625

Pudney, C. R.; Parker, E. J.; Schipper, L. A. On the Temperature Dependence of

626

Enzyme-Catalyzed Rates. Biochemistry 2016, 55 (12), 1681–1688.

627

(43)

Raabe, I.; Himmel, D.; Krossing, I. Computational Study of the Enthalpies of

628

Formation, ∆f H°, and Mean Bond Enthalpies, MBEs, of H4-n EX n 0/-and H3-n EX

629

N+/0 (E= C, B; X= F− I). J. Phys. Chem. A 2007, 111 (50), 13209–13217.

630

(44) Alberty, R. A. Biochemical Thermodynamics and Rapid-Equilibrium Enzyme Kinetics. J. Phys. Chem. B 2010, 114 (51), 17003–17012.

631 632

(45)

Synthesis on Solid Support. J. Am. Chem. Soc. 2002, 124 (37), 10988–10989.

633 634

(46)

Edgar, R. C. MUSCLE: Multiple Sequence Alignment with High Accuracy and High Throughput. Nucleic Acids Res. 2004, 32 (5), 1792–1797.

635 636

Ulijn, R. V; Baragaña, B.; Halling, P. J.; Flitsch, S. L. Protease-Catalyzed Peptide

(47)

Totrov, M.; Abagyan, R. Flexible Protein–ligand Docking by Global Energy

637

Optimization in Internal Coordinates. Proteins Struct. Funct. Bioinforma. 1997, 29

638

(S1), 215–220.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

28 639 640

(48)

Chen, C. S.; Wu, M. C. Kinetic Models for Thermal Inactivation of Multiple Pectinesterases in Citrus Juices. J. Food Sci. 1998, 63 (5), 747–750.

641

642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658

ACS Paragon Plus Environment

Page 28 of 41

Page 29 of 41

Journal of Agricultural and Food Chemistry

29

659

Figure legends

660

Figure 1. SDS–PAGE of recombinant AEH protein from Stenotrophomonas maltophilia.

661

SDS–PAGE (10% separating and 3% stacking) gel stained with Coomassie brilliant blue.

662

Lane 1, pETSM-expressed protein lysate; lane 2, DEAE-Cellufine; and lane 3, DEAE-

663

Sepharose (2 µg protein). The band corresponding to SmAEH is indicated. Molecular mass

664

markers were loaded in the lane labeled M, and their masses are shown on the left in kDa.

665

Figure 2. (a) Partial alignment of SmAEH with experimentally characterized AEHs. The

666

sequences shown are AEHs from Stenotrophomonas maltophilia HS1 (SmAEH, in this

667

study), Elizabethkingia sp. (EsAEH, data from our lab), Acetobacter turbidans (AtAEH,

668

protein ID no. AAL60195), and Xanthomonas citri (XcAEH, protein ID no. AAO24758).

669

The alignment was performed using the online “MUSCLE” multiple sequence alignment

670

server46. Catalytic triad residues within sequences are shaded black, and carboxylate cluster

671

residues are shaded red. (b and c) Superposition of the catalytic triad and carboxyl cluster

672

residues of SmAEH with XcAEH (b) and AtAEH (c). SmAEH is shown as red sticks, and

673

XcAEH and AtAEH are in green. Residues are labeled black and are numbered. The figure

674

was prepared using Molscript47.

675

Figure 3. Thermal inactivation SmAEH. Data are averages of two independent experiments.

676

The standard deviations are always lower than 6%. Residual activity was determined after

677

incubation at 35°C (closed square), 40°C (closed rectangle), 45°C (closed angle), and 50°C

678

(closed circle). Data were fitted to the first-order model (Eq. 1).

679

Figure 4. Arrhenius plot for inactivation of SmAEH. The graph was obtained by plotting the

680

logarithm of the rate constant ( k ) versus the inverse temperature (1/T). The regression

681

equation was determined to be − ln ( k ) = 24552T −1 − 75.142 ( r 2 = 0.984 ) . Data are presented as

682

the mean ± standard errors of the mean (SEM) of three independent experiments.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 30 of 41

30 683

Figure 5. Correlation between log(D) and temperature for the thermal inactivation of SmAEH.

684

The graph was obtained by plotting the D values on a log scale against the corresponding

685

temperatures.

686

log ( D ) = −0.124T + 6.7357 r 2 = 0.987 . Data were presented as the mean ± standard errors of

687

the mean (SEM) of three independent experiments.

The

regression

(

equation

was

)

ACS Paragon Plus Environment

determined

to

be

Page 31 of 41

Journal of Agricultural and Food Chemistry

31

Table 1. Kinetic equations used to analyze inactivation of L-Amino acid ester hydrolase Equation no. Model 1

Equations

First-order20

A A0

= exp ( −kt )

2

Two-fraction48

A = aAL exp ( −kLt ) + (1 − a ) exp ( −kRt ) A0

3

Fractional conversion22

A = Ar + ( A0 − Ar ) exp ( −kt ) A0

4

nth order23

5

Weibull distribution24

1/ (1− n ) A = { A01−n + ( n − 1) kt} A0

A = exp ( −bt n ) A0

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 32 of 41

32

Table 2. Purification of SmAEH from E. coli Rosetta-gami B (DE3) expressing pETSM plasmid

Steps

Total protein (mg)

Total activity (U)

Specific activity (U/mg)

Yield (%)

Purification fold

Cell free extract

150.0

250

1.7

100

1

DEAE-Cellufine

25.0

175

7.0

70.0

4.21

DEAE-Sepharose

4.5

51

11.3

20.4

6.80

ACS Paragon Plus Environment

Page 33 of 41

Journal of Agricultural and Food Chemistry

33

Table 3. Kinetics parameter of SmAEH

Substrates

Mean ± SE of:

Acyl donor Acyl Acceptor

Km (mM) acyl donor

Km (mM) acyl acceptor

kcat (s-1) acyl donor

kcat (s-1) acyl acceptor

kcat/Km acyl donor

kcat/Km acyl acceptor

L-Ile-OMe

L-Trp

10.4±0.8

35±1.5

17.26±0.8

17.5±1.0

1.65±0.15

0.50±0.05

L-Val-OMe

L -Gly

22.0±1.5

65±5.0

10.16±0.8

10.5±0.5

0.37±0.02

0.16±0.01

L-Trp-OMe

L -His

27.0±2.0

55±3.0

7.05±0.5

7.0±0.6

0.26±0.04

0.13±0.01

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 34 of 41

34

Table 4. Statistical parameters of the kinetic models describing thermal inactivation of SmAEH

r2

χ2

SEM

Remarks

First-order (1)

0.989; 0.998

0.022; 0.00027

0.026;0.00077

Accepted: higher r2and low x ଶ and SEM

Two-fraction (2)

0.987; 0.998

0.0886; 0.0012

0.0511;.0026

Rejected: equal parameter estimates

Fractional conversion (3)

0.995; 0.998

0.0894; 0.02651

0.013;.0009

Rejected: negative parameter estimates

nth order (4)

0.945; 0.988

0.094; 0.026

0.0936;0.025

Rejected: low r2 and high x ଶ and SEM

Weibull distribution (5)

0.980; 0.997

0.0482; 0.00137

0.027; 0.00224

Rejected: low r2 and high x ଶ and SEM

Model (Eq.)

ACS Paragon Plus Environment

Page 35 of 41

Journal of Agricultural and Food Chemistry

35

Table 5. Kinetics and thermodynamics data for thermal inactivation of SmAEH

Temperature (°C)

k (min-1)

35

0.011

40

D (min)

∆H#

∆G#

∆S#

63.0

212.80

201.61

87.08

371.85

0.029

23.90

81.30

201.56

86.01

369.17

45

0.16

4.30

12.10

201.52

80.56

380.40

50

0.39

1.80

3.40

201.48

81.87

370.34

t 1/2 (min)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 36 of 41

36

(kDa)

M

1

2

3

204 114 84 72 kDa

62 46

27

14

Figure 1

ACS Paragon Plus Environment

Page 37 of 41

Journal of Agricultural and Food Chemistry

37

(a) SmAEH EsAEH AtAEH XcAEH

151 126 174 143

RVDHSTDAWDTIDWLVKNVPESNGKVGMLGSSYKGFTVVMALTDPHPALKVAAPQSPMVDGWMSDDWLNYGA-FRPGQLQ GIDESTDTFDTLEWLSKNLKNYNQKAGVYGISYPGFYSTTTLVNSHPTLKAVSPQAPVTNWYLGDDFHHKGAMFLNDAFM KTDETTDAWDTVDWLVHNVPESNGRVGMTGSSYEGFTVVMALLDPHPALKVAAPESPMVDGWMGDDWFHYGA-FRQGAFD EVDHATDAWDTIDWLVKNVSESNGKVGMIGSSYEGFTVVMALTNPHPALKVAVPESPMIDGWMGDDWFNYGA-FRQVNFD

SmAEH EsAEH AtAEH XcAEH

325 288 328 297

LKVPTMWLQGLWDQEDMWGANHAYQAMEGRDSGNNRNYLVMGPWRHSQ-VNYSGSELGALKFDGDTALQFRRDVLKPFFD PAV--MVVGGFFDAEDAYGTFETYKAIEKQNPKAN-NILVAGPWFHGGWVRGDGKQFGDIKFDHPTSIDYQQNLELPFFN PTVPMLWEQGLWDQEDMWGAIHAWQALKDADVKAP-NTLVMGPWRHSG-VNYNGSTLGPLEFEGDTAHQYRRDVFRPFFD LKVPTMWLQGLWDQEDMWGAIHSYAAMEPRDKRNTLNYLVMGPWRHSQ-VNYDGSALGALNFEGDTARQFRHDVLRPFFD

(b)

(c)

Figure 2

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 38 of 41

38

1 0.9 0.8 0.7

A/A0

0.6 0.5 0.4 0.3 0.2 0.1 0 0

10

20

30

40

Time (min) Figure 3

ACS Paragon Plus Environment

50

60

70

80

Page 39 of 41

Journal of Agricultural and Food Chemistry

39

5 4.5 4

-ln(k)

3.5 3 2.5 2 1.5 1 0.5 0 0.00308 0.00311 0.00314 0.00317 0.0032 0.00323 0.00326 1/T (k-1)

Figure 4

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 40 of 41

40

3 2.5

Log (D)

2 1.5 1 0.5 0 30

35

40 45 Temperature ( °C)

Figure 5

ACS Paragon Plus Environment

50

55

Page 41 of 41

Journal of Agricultural and Food Chemistry

41

TOC graphics (width, 3.30 inch; height 1.85 inch)

ACS Paragon Plus Environment