Hybrid Cross-Linked Lipase Aggregates with Magnetic Nanoparticles

Sep 6, 2016 - Highly stable and easily recyclable hybrid magnetic cross-linked lipase aggregates (HM-CSL-CLEAs) were prepared by coaggregation of ...
2 downloads 0 Views 785KB Size
Subscriber access provided by CORNELL UNIVERSITY LIBRARY

Article

Hybrid cross-linked lipase aggregates with magnetic nanoparticles: a robust and recyclable biocatalysis for epoxidation of oleic acid Jiandong Cui, Lili Cui, Shiru Jia, Zhiguo Su, and Songping Zhang J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b01939 • Publication Date (Web): 06 Sep 2016 Downloaded from http://pubs.acs.org on September 6, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 37

Journal of Agricultural and Food Chemistry

1

Hybrid cross-linked lipase aggregates with magnetic nanoparticles:

2

a robust and recyclable biocatalysis for epoxidation of oleic acid

3 4

Jiandong Cui 1, 3* Lili Cui3 Shiru Jia1

Zhiguo Su2 Songping Zhang2*

5

1

6

of Science and Technology, No 29, 13th, Avenue, Tianjin Economic and Technological

7

Area (TEDA), Tianjin 300457, P R China

8

2

9

Academy of Sciences, Beijing, P R China

Key Laboratory of Industrial Fermentation Microbiology, Ministry of Education, Tianjin University Development

National Key Laboratory of Biochemical Engineering, Institute of Process Engineering, Chinese

10

3

11

Hebei University of Science and Technology, Shijiazhang, P R China

Research Center for Fermentation Engineering of Hebei, College of Bioscience and Bioengineering,

12 13 14

* Corresponding author:

15

Jiandong Cui, E-mail:[email protected], Tel: +86-311-81668486

16

Address: Key Laboratory of Industrial Fermentation Microbiology, Ministry of Education, Tianjin

17

University of Science and Technology, No 29, 13th, Avenue, Tianjin Economic and Technological

18

Development Area (TEDA), Tianjin 300457, P R China

19

Songping Zhang, E-mail: [email protected], Tel: +86-10-82544958

20

Address: National Key Laboratory of Biochemical Engineering, Institute of Process Engineering,

21

Chinese Academy of Sciences, Beijing, 100190, P R China

22 23 24 25

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

26

Abstract

27

The highly stable and easy recyclable hybrid magnetic crosslinked lipase aggregates

28

(HM-CSL-CLEAs) were prepared by co-aggregation of lipase aggregates with nonfunctionalized

29

magnetic nanoparticles and subsequent chemical cross-linking with glutaraldehyde. Analysis by SEM

30

and CLSM indicated that the CLEAs were embedded in nanoparticle aggregates instead of covalent

31

immobilization. The resulting HM-CSL-CLEAs exhibited higher thermostability, storage stability,

32

and reusability than standard CLEAs. For example, HM-CSL-CLEAs maintained more than 60% of

33

their initial activity after 40 min incubation time at 60 oC, whereas standard CLEAs lost most of their

34

activities. The HM-CSL-CLEAs can be easily recovered from the reaction mixture by an external

35

magnetic field. Moreover, the H2O2-tolerance of the lipase in HM-CSL-CLEAs was also enhanced,

36

which could relieve the inhibitory effect on lipase activity. High conversion yield (55%) for the

37

epoxidation of oleic acid using H2O2 as oxidizing agent was achieved by HM-CSL-CLEAs.

38

Keywords: Magnetic nanoparticles; Cross-linked enzyme aggregates; Lipase; Epoxidation; Enzyme

39

Immobilization

40 41 42 43 44 45 46

2

ACS Paragon Plus Environment

Page 2 of 37

Page 3 of 37

Journal of Agricultural and Food Chemistry

47

1. Introduction

48

Epoxides are used widely in pharmaceuticals, cosmetics, and food additives1. Furthermore,

49

epoxidized vegetable oils and their derivatives have been used as additives for production of poly

50

(vinyl chloride) (PVC) 2. Furthermore, as a higher quantity of additive, PVC has been considered to

51

be used in food packaging 3. Over the past twenty years, the enzymatic epoxidation has shown to be

52

effective for the epoxidation of vegetable oils with aqueous hydrogen peroxide. Lipases (E.C.3.1.1.3)

53

are the most broadly employed biocatalysts due to their variety of reactions such as hydrolysis 4, lipid

54

synthesis

55

epoxidation of alkenes with hydrogen peroxide (H2O2) 9. Generally, it is necessary to consume H2O2

56

during epoxidation of alkenes. However, H2O2 is able to modify the peptide core of the enzymes,

57

which results in the oxidation of some cofactors and prostetic groups 8, 9. As a result, the inactivation

58

of the lipases occur 10. Moreover, most free lipases display low stability and difficulties in recovery

59

and reuse, leading to high cost and the contamination of the products

60

enzymes is a promising technology that can overcome these limitations

61

dissociation of enzymes subunit can be prevented by multisubunit immobilization 15, and the enzyme

62

structure can be rigidified by multipoint covalent attachment 16. Among the various methodologies of

63

enzyme immobilization, cross-linked enzyme aggregates (CLEAs) may be considered to be a

64

promising method for enzyme immobilization

65

volumetric productivities can be simply prepared by cross-linking the physical enzyme aggregates 19,

66

20

67

treatments, which would cause internal mass-transfer limitations. Especially, it is difficult to handle

5, 6

, transesterification

7

and epoxidation 8. Especially, lipases could be used for the direct

17, 18

11, 12

. Immobilization of

13, 14

. For example,

. CLEAs with superior stability and high

. Unfortunately, the small and fragile CLEAs can form clumps during centrifugation and filtration

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 37

68

and fully recover the CLEA particle 21-23. Therefore, development of a feasible and efficient approach

69

for improving the catalytic properties of traditional CLEAs is highly desirable. In recent years,

70

supported CLEAs strategies were developed to overcome these disadvantages. For example, CLEAs

71

of papain were prepared in commercial macroporous silica gel. The resulting CLEAs showed an

72

impressive stability

73

Rhodotorula glutinis into macroporous silica gel exhibited good reusability

74

noncatalytic mass of the carrier causes decrease of the volumetric activity and production of the

75

biocatalyst. In order to improve these drawbacks, recently, magnetic nanoparticles have attracted

76

more and more attention due to their high specific surface area magnetic properties, which can

77

efficiently improve enzyme loading and be easily separated using a magnetic field

78

was immobilized as CLEAs on amino functionalized magnetite nanoparticles. The magnetic CLEAs

79

exhibited excellent thermostability, storage stability and reusability

80

were bonded on amino functionalized magnetite nanoparticles. The magnetic CLEAs were recycled

81

without appreciable loss of activity

82

CLEAs was obtained by cross linking aggregated glucoamylase using dialdehyde pectin with amino

83

functionalized magnetite nanoparticles. The magnetic macromolecular CLEAs retained 84% activity

84

after 10th cycle without leaching of enzyme which was 22% higher than traditional CLEAs 31. Laccase

85

was immobilized as CLEAs (MAC-CLEAs) on amine functionalized magnetic nanoparticles using

86

chitosan/1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide hydrochloride as the cross-linking system.

87

The MAC-CLEAs exhibited enhanced storage stability over one year at 4℃ and showed better

88

temperature resistance compared to its soluble counterpart 32.

24

. The formation of CLEAs of phenylalanine ammonia lyase (PAL) from

30

29

25

. However, the

26-28

. α-amylase

. Magnetic CLEAs of laccase

. More recently, the magnetic macromolecular glucoamylase

4

ACS Paragon Plus Environment

Page 5 of 37

Journal of Agricultural and Food Chemistry

89

Moreover, CLEAs of Candida antarctica lipase B (CALB) were immobilized on magnetic

90

nanoparticles by covalently bound. The magnetic CLEAs can be easily recovered and reused for 10

91

cycles of 24 h without apparent loss of activity

92

magnetic CLEAs could contribute to prepare stable CLEAs. Moreover, the preparation strategy of the

93

magnetic CLEAs in these reports is standard. Namely, the iron oxide nanoparticles were first reacted

94

with activation agent 3-aminopropyl triethoxysilane (APTES) to synthesize stable amino

95

functionalized particles. Afterwords, enzymes were precipitated onto amino functionalized particles.

96

Lastly, a strong chemical bond was formed by cross-linker (glutaraldehyde) between primary amine

97

group which reacted with the carboxyl terminals (–COOH) on the surface of the precipitated enzymes

98

on the magnetic support

99

magnetite nanoparticles results in conformational changes in the enzyme. Moreover, this standard

100

process is time-consuming and high cost. Considering these drawbacks to the industrial application of

101

CLEAs, an innovative approach is required. In this study, a novel strategy was developed to prepare a

102

hybrid magnetic CLEA (HM-CSL-CLEAs) by co-aggregated lipase with unfunctionalized magnetite

103

nanoparticles and subsequently cross-linked with glutaraldehyde. The differences in this strategy with

104

standard magnetic CLEAs are that magnetite nanoparticles are not modified by amino

105

functionalization. Moreover, the CLEAs were embedded in nanoparticle aggregates instead of

106

covalent immobilization. Compared to covalent immobilization of CLEAs on functionalized

107

magnetite nanoparticles, the natural structure of lipase in this strategy can be maintained due to

108

prevent the formation of covalent bonds between lipase and magnetite nanoparticles. Generally, the

109

activities of the CLEAs are affected by many factors such as the precipitan, the cross linker

33-35

33

. The above reports indicated that the strategy of

. Unfortunately, the formation of covalent bonds between enzymes and

5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 37

110

(glutaraldehyde) and the cross-linking time, etc. Therefore, it is necessary to optimize different

111

parameters for the prepararion of CLEAs. Response surface methodology (RSM) is a powerful

112

statistical and mathematical tool. This method has been successfully applied in many areas of

113

biotechnology. Effect of main factors and their possible interactions on biological process can be

114

determined accurately by RSM

115

preparation of CLEAs, such as CLEA- hydroxynitrile lyase 36, CLEAs-lipase

116

38

117

activity of the HM-CSL-CLEAs were first investigated by RSM. Secondly, the catalytic properties of

118

HM-CSL-CLEAs were investigated, including temperature, pH, and operational stability. Lastly, the

119

catalytic efficiency of HM-CSL-CLEAs and standard lipase-CLEAs for the epoxidation of oleic acid

120

using H2O2 as oxidizing agent were evaluated.

34, 35

. Recently, RSM was successfully applied to optimize the 37

and CLEA-protease

. Therefore, in this study, the main immobilization parameters which affect the final volumetric

121 122

2 Materials and Methods

123

2.1 Materials

124

Candida sp. lipase (30 U/mg) was purchased from Sigma (St. Louis, USA). Olive oil was purchased

125

from Sinopharm Chemical Reagent Co., Ltd. (Beijing, China). Bis (2-ethylhexyl) sulfosuccinate

126

sodium (AOT), FeCl2·4H2O, and FeCl3·6H2O were purchased from Xilong Chemical Co., Ltd.

127

(Shantou, China). Other reagents used were of analytical grade .The water used in this work was

128

deionized water.

129

2.2 Preparation of magnetite nanoparticles (MNPs), standard lipase-CLEAs (CSL-CLEAs), and

130

HM-CSL-CLEAs

6

ACS Paragon Plus Environment

Page 7 of 37

Journal of Agricultural and Food Chemistry

131

MNPs were prepared by the modified method of Ranjbakhsh et al 39. Briefly, 1.25 g of FeCl2·4H2O

132

and 3.40 g of FeCl3·6H2O were dissolved in 100 mL deionized water under nitrogen gas; then, 6 mL

133

of 25% NH3·H2O were added and vigorously stirred for 30–40 min at room temperature. The MNPs

134

were recovered and washed several times with the deionized water.

135

The CSL-CLEAs were prepared as previously reported

136

phosphate buffer (pH 7.5) were precipitated using 80% ammonium sulfate saturation under gentle

137

stirring at 4 °C for 1 h. A certain amount of glutaraldehyde (10% v/v in water) were slowly added in

138

the mixture for the cross-linking of the aggregates, and stirred at 4 °C for a specified time. Then the

139

suspension was centrifuged at 10000×g for 10 min at 4 °C. The precipitate was washed three times by

140

50 mM sodium phosphate buffer (pH 7.5) and finally stored in 50 mM sodium phosphate buffer (pH

141

7.5) at 4 °C prior to use. For HM-CSL-CLEAs, MNPs (20 mg/mL) and commercial lipase solution

142

(1.75 mg/mL) were mixed, and ammonium sulfate was added to a final 80% ammonium sulfate

143

saturation under gentle stirring at 4 °C for 1 h. Afterthat, glutaraldehyde (25 % v/v) was added to

144

cross-link the precipitate for 2 h under shaking at 200 rpm. Then the immobilized lipases were

145

separated by magnet. The biocatalysts were washed three times with distilled water, and finally the

146

prepared HM-CSL-CLEAs were stored in 50 mM sodium phosphate buffer (pH 7.5) at 4 oC.

147

2.3 The optimization of preparing HM-CSL-CLEAs

148

The optimization of preparing HM-CSL-CLEAs

149

one-factor-at-a-time experiments and RSM experiments based experimental design. First, the

150

influence of main immobilization parameters on the preparation of HM-CSL-CLEAs was evaluated

151

by traditional non-statistical technology; they are enzyme concentration, (NH4)2SO4 concentration

40

was

. Lipases (0.1 g) in 50 mM sodium

carried out by combination of

7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 37

152

and glutaraldehyde (GA) concentration. Central composite design (CCD) was used to find the

153

optimal concentrations of these three factors and to understand the relationship between the factors

154

and lipase activity. The used CCD was generated by “Design-Expert” software (version 7.0, Stat-Ease,

155

Inc., Minneapolis, USA)

156

glutaraldehyde concentration (X3) were optimized by the experimental plan. Each factor in the design

157

was studied at five different levels (Table 1).

158

2.4 Protein and activity assay

159

The quantities of protein were measured by the method of ultraviolet absorption method at 280 nm 41.

160

Lipase activity was determined by measuring the fatty acids produced in reversed micellar solution.

161

Fatty acids produced were determined by modified copper soap colorimetry

162

used to prepare cupric acetate reagent

163

reversed micellar solution was pre-incubated at 37 oC with magnetic stirring (500 rpm). Olive oil

164

(10%, v/v) was added to the reversed micellar solution for the hydrolysis. The reactant was collected

165

at certain time and diluted with benzene. afterthat, 1 mL of the cupric acetate reagent was added and

166

mixed vigorously at 180 rpm for 10 min at 37 oC. After centrifugation, the rate of formation of free

167

acids was determined by measuring the absorbance of isooctane solution at 710 nm against the

168

control which contains no free fatty acids. One unit of lipase activity was defined as the amount of

169

lipase releasing 1 mmol fatty acids from olive oil per minute. The activity recovery in

170

HM-CSL-CLEAs and CSL-CLEAs were calculated as given in Eq.1:

171

Activity recovery (%)

172

2.5 Morphology of HM-CSL-CLEAs

35

. The enzyme concentration (X1), (NH4)2SO4 concentration (X2) and

43

42

. Lowry method was

. The bis (2-ethylhexyl) sulfosuccinate sodium/isooctane

activity of CLEAs (U) = Total free lipaseTotal activity used for CLEAs production (U)

8

ACS Paragon Plus Environment

1)

Page 9 of 37

Journal of Agricultural and Food Chemistry

173

Morphology of the MNPs, CSL-CLEAs and HM-CSL-CLEAs was examined by scanning electron

174

micrographs (SEMs) (JEOL JSM6700, Japan). The distribution of lipase within CLEAs was observed

175

by Confocal Laser Scanning Microscopy (CLSM) (Leica Camera AG, Germany). Before observation,

176

modified FITC labeled lipase was first prepared by adding 50 mg/mL FITC solution (FITC in acetone)

177

into lipase solution for 3 min, and then was immobilized 44. The samples were excited at 390 nm and

178

the emitted fluorescent light was detected between 460 and 480 nm. Magnetisation measurements

179

were carried out using a Magnetic Property Measure System (MPMS, Quantum design) under

180

magnetic fields at 300 K.

181

2.6 Measurement of apparent kinetic parameters

182

The apparent kinetic parameters for free lipase and CLEA were obtained by the Lineweaver-Burk

183

double-reciprocal plot method of Michaelis-Menten Equation between 0.12 mM and 4.9 mM olive oil

184

concentrations at a constant enzyme concentration (0.5 mg/mL) under assay conditions.

185

2.7 Optimal temperature and pH for CLEA activity

186

The optimum temperature of free lipase, CSL-CLEAs, and HM-CSL-CLEAs on the activity were

187

determined by measuring activity at different temperature (35-55 oC). Meanwhile, the optimal pH

188

free lipase, CSL-CLEAs, and HM-CSL-CLEAs on the activity was determined by measuring activity

189

at various pHs (3-8). The activity of the free lipase, CSL-CLEAs and HM-CSL-CLEAs was

190

determined under assay conditions.

191

2.8 The stability of free lipase, CSL-CLEAs and HM-CSL-CLEAs

192

The thermal stability of the free lipase, CSL-CLEAs and HM-CSL-CLEAs was determined by

193

incubating them in 50 mM sodium phosphate buffer (pH 7.5) without substrate at 60 oC for 10-40

9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

194

min. At different incubation time, the residual lipase activity was determined. The storage stability of

195

free lipase, CSL-CLEAs and HM-CSL-CLEAs were determined by incubating enzyme samples in 50

196

mM sodium phosphate buffer (pH 7.5) at 4 oC. At different storage time, the residual activities of free

197

lipase and immobilized lipase were determined. To analyze the reusability of the CSL-CLEAs and

198

HM-CSL-CLEAs, olive oil was hydrolyzed as described above. Upon completion of one cycle, the

199

HM-CSL-CLEAs were then separated by a magnet (CAL-CLEAs were recovered by centrifugation).

200

The recovered CLEAs were washed three times with 50 mM sodium phosphate buffer (pH 7.5) and

201

then suspended again in a fresh reaction mixture.The residual lipase activity at each cycle was

202

calculated by taking the enzyme activity of the first cycle as 100%.

203

2.9 The epoxidation of oleic acid

204

The formation of epoxystearic acid from the epoxidation of oleic acid using H2O2 as oxidizing agent

205

was evaluated by HM-CSL-CLEAs and CSL-CLEAs. The epoxidation reaction was carried out in a

206

50 mL serum bottle. The reaction mixture containing 20 mg/mL oleic acid, 25 mmol/L phosphate

207

buffer (pH 7.0), 200 µL H2O2 (30 wt%), and 20 mg immobilized enzymes (10 U/mg) was incubated

208

at 40 °C for 96 h at 120 rpm. Every 48 h, the formation of epoxystearic acid was analyzed by

209

HPLC-ELSD. HPLC conditions were as follows: injection volume of 10 µL; column temperature of

210

40 oC; flow rate of 1 mL/min; gradient elution was applied in chromatographic separation, acetonitrile

211

concentration was increasing from 85% to 100%. Substrate and product were detected by ELSD at 40

212

o

C, the gas flow rate was 1.5 mL/min, Gain was 1.

213 214

3 Result and discussion

10

ACS Paragon Plus Environment

Page 10 of 37

Page 11 of 37

Journal of Agricultural and Food Chemistry

215

3.1 Structural characterization and magnetic response of HM-CSL-CLEAs

216

General scheme of preparation of HM-CSL-CLEAs was illustrated in Fig. 1. The synthesis of

217

HM-CSL-CLEAs involved three main steps. First, some nanoparticles were encapsulated into lipase

218

aggregates when lipases and non-functionalized magnetic nanoparticles were mixed and then

219

precipitated with ammonium sulfate. Second, enzyme aggregates with nanoparticles were

220

cross-linked by glutaraldehyde to form HM-CSL-CLEAs. Lastly, HM-CSL-CLEAs were further

221

embeded into the nanoparticles due to aggregation of magnetic nanoparticles. Here, we firstly

222

analyzed the size and shape of HM-CSL-CLEAs. SEM images (Fig. 2a) of standard CSL-CLEAs

223

revealed no defined morphologies and large size particles. Moreover, the surface of standard

224

CSL-CLEAs was relatively smooth, except for a few tiny pores. However, it is obviously observed

225

that lipase CLEAs were inlayed within nanoparticles (The CLEAs embedded in the nanoparticles was

226

indicated by the arrow). Furthermore, the porous and loose structures of HM-CSL-CLEAs were

227

found (Fig. 2b). We estimated that HM-CSL-CLEAs with porous structures decreased mass transfer

228

resistance compared to standard CSL-CLEAs, leading to superior catalytic efficiency. To further

229

confirm the embedding of CLEAs within the HM-CSL-CLEAs aggregates, we used a confocal laser

230

scanning microscopy (LCSM) to observe the distribution of CLEAs in nanoparticles. As shown in Fig.

231

2(c) and (d), no fluorescence was observed for nanoparticles without CLEAs (Fig. 2(c)), whereas

232

nanoparticles with CLEAs had an apparent fluorescence (Fig. 2(d)), indicating that the CLEAs were

233

embedded within the nanoparticles. In addition, the field-dependent magnetisation curve of

234

HM-CSL-CLEAs was displayed in Fig. 3. There was no hysteresis observed in low fields, indicating

235

that the superparamagnetic behavior of HM-CSL-CLEAs is intact. The HM-CSL-CLEAs with

11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

236

superparamagnetic characteristics could be easily recovered from reaction mixture by an external

237

magnetic field (Fig. 3b), and re-dispersed easily by a slight shake while removing the magnetic field

238

(Fig. 3a), indicating efficient recyclability in industrial application.

239

3.2 The optimization of preparing HM-CSL-CLEAs

240

In the CLEA formation, the amount of ammonium sulfate, glutaraldehyde concentration, enzyme

241

concentration, and cross-linking time play a significant role in CLEA activity 45, 46. For example, too

242

little cross-linker is not enough to form the cross-linked aggregates that causes leaching of enzyme

243

molecules, while high glutaraldehyde concentration can result in decrease of enzyme activity due to

244

loss of the minimum flexibility 47, 48. In our study, to find the critical factors which have affected on

245

activity recovery, critical factors for preparation of HM-CSL-CLEA were first chosen based on the

246

one-factor-at-a-time method. The results showed that enzyme concentration, (NH4)2SO4 concentration

247

and glutaraldehyde concentration have significant effects on the activity recovery of HM-CSL-CLEA.

248

To further find the optimal concentrations of these three factors and relationship between the factors

249

and activity recovery, the CCD experiments for studying the effects of enzyme concentration,

250

(NH4)2SO4 concentration, and glutaraldehyde concentration on activity recovery was carried out. The

251

results were presented in the Table 2 and Table 3. The model F-value of 7.78 implied that the model

252

was significant. The regression equation obtained from analysis of variance (ANOVA) indicated that

253

the multiple correlation coefficient of R2 is 0.8752. The value of the determination coefficent

254

indicates that the model can explain 87.52% variation in the response. The high R2 value indicated a

255

high significance of the model 49. The above results showed that the models were adequate to predict

256

the activity recovery within the range of variables studied.

12

ACS Paragon Plus Environment

Page 12 of 37

Page 13 of 37

Journal of Agricultural and Food Chemistry

257

The analysis of variance indicated that the model terms of X1, X12, X32, and X1X3 were significant

258

(“probe>F” less than 0.05), and the interactive effect of X1X3 was significant (“probe>F” less than

259

0.05). However, the interactive effects of X1X2, and X2X3 were not significant. It indicates that

260

glutaraldehyde concentration and enzyme concentration play a key role on activity recovery. A

261

second-order polynomial equation was obtained by multiple regression analysis of the experimental

262

data:

263

Y=41.6 + 4.26X1 + 2.53X2 +1.48X1X3 – 6.9X12 – 0.56X1X2 + 6.6X1X3 – 2.51X22 – 1.34X2X3 – 7.7X32.

264

where Y is the response, that is activity recovery (%), and X1, X2 and X3 are the coded values of the

265

test variables lipase, (NH4)2SO4 and glutaraldehyde, respectively.

266

The 2D contour plots and 3D response surface are generally the graphical representations of the

267

regression equation, and 2D contour plots and 3D response surface are presented in Fig. 4. As

268

mentioned by Muralidhar et al 50, the interactions are negligible between the corresponding variables

269

when the contour plots of response surfaces is circular. In contrast, the interactions between the

270

corresponding variables are significance while the contour plots exhibit an elliptical shape. The

271

contour plots in Fig. 4 (c, d) showed that there was significant interaction between enzyme

272

concentration and glutaraldehyde concentration. The optimal conditions were extracted by Design

273

Expert Software. They were enzyme concentration at 1.95 mg/mL, (NH4)2SO4 concentration at 2.56

274

M, and glutaraldehyde concentration at 0.23%, respectively. Moreover, the model predicted that the

275

maximum activity recovery that can be obtained under the above optimum conditions was 38..67%.

276

The maximum activity recovery obtained experimentally was found to be 40.86%. This is obviously

277

in close agreement with the model prediction, which thus proved the validity. In a word, an 19.32%

13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 37

278

increase in activity recovery of HM-CSL-CLEAs was acquired under optimized conditions.

279

3.3 Measurement of apparent kinetic parameters

280

Apparent kinetic properties of free lipase, CSL-CLEAs, and HM-CSL-CLEAs showed that Km was

281

increased when lipase was immobilized (Table 4). The increase of Km value of CLEA could be due to

282

the mass transfer limit for the substrate as well as the rigid conformation of lipase caused by

283

cross-linking 48, 51 . However, Km of HM-CSL-CLEAs was lower than that of standard CSL-CLEAs.

284

Moreover, the catalytic efficiency (Kcat/Km) of HM-CSL-CLEA was higher compared to standard

285

CSL-CLEAs, indicating that HM-CSL-CLEAs were an overall better catalyst. The porous and loose

286

structures of HM-CSL-CLEAs are critical for ensuring high catalytic efficiency, due to an increase in

287

mass substrate transfer for interaction with the interior enzymes of CLEAs. This phenomenon was

288

also observed in other reports 52, 53.

289

3.4 Optimal pH and temperature on enzyme activity

290

Fig. 5 showed the effect of pH on the activity of free lipase, CSL-CLEAs, and HM-CSL-CLEAs. The

291

optimum pH of the free lipase was at pH 7.0, which shifted to pH 6.0 for CSL-CLEAs and

292

HM-CSL-CLEAs. Moreover, compared to free lipase and CSL-CLEAs, HM-CSL-CLEAs appeared

293

a much broader pH range of 4 to 8 with higher activity. The shift in pH could be due to the change in

294

acidic and basic amino acid side chain ionization in the microenvironment around the active site, as

295

well as the interaction between the cross-linking agent and the enzyme

296

temperature on the activity of free lipase and all immobilized lipase samples was compared. As

297

shown in Fig. 6, the optimum reaction temperature for free lipase was at 45 oC. However, the optimal

298

temperature on the activity of both CSL-CLEAs and HM-CSL-CLEAs were at 50 oC. It indicates that

14

ACS Paragon Plus Environment

54

. In addition, optimal

Page 15 of 37

Journal of Agricultural and Food Chemistry

299

immobilization improves the thermostability of the enzyme. The high thermostability could be due to

300

the increased rigidity of enzymes by the cross-linking among enzyme molecules 30.

301

3.5 Stability of HM-CSL-CLEAs

302

The thermal stability of free lipase, CSL-CLEAs and HM-CSL-CLEAs at 60 oC was shown in Fig. 7.

303

All CLEAs exhibited the higher thermal stability than that of free lipase. Moreover, HM-CSL-CLEAs

304

showed the highest stability. For example, free lipase and CSL-CLEAs lost most of their activities

305

after 40 min incubation time at 60 oC, whereas HM-CSL-CLEAs retained more than 60% of their

306

initial activity in the same conditions. These results indicated that the thermal stability of

307

HM-CSL-CLEAs was significantly improved. This enhancement of HM-CSL-CLEAs in thermal

308

stability could be due to the following reasons: (1) magnetic nanoparticles provide a suitable

309

protection to retard heat transfer and collapse of the CLEAs structure at high temperatures, (2) the

310

rigidity of the active conformation was improved by cross-linking among enzyme aggregates. In

311

addition, storage stability of the free lipase, CSL-CLEAs and HM-CSL-CLEAs was presented in Fig.

312

8. Most activity of both CLEAs forms was retained after 28 days storage. However, free lipase

313

retained only 28% of its initial activity at the same storage time. HM-CSL-CLEAs still retained more

314

than 75% of its initial activity after 38 d. However, the CSL-CLEAs only retained 54% of its initial

315

activity. The reusability of CSL-CLEAs and HM-CSL-CLEAs was shown in Fig. 9, although the

316

activities of both CSL-CLEAs and HM-CSL-CLEAs decrease with increasing number of recycles,

317

the HM-CSL-CLEAs retained more than 65% of its initial activity until 10 cycles. However, the

318

CSL-CLEAs almost lost activity. The result showed that the HM-CSL-CLEAs presented good

319

reusability. This reusability property of HM-CSL-CLEAs is necessary for cost-effective use of the

15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 37

320

enzyme in industrial applications. In the past few decades, lipases have been immobilized in a great

321

variety of supports, such as, mesoporous silica

322

calcium carbonate microcapsule

323

activity and reusability of the biocatalysts are undesirability.

324

3.6 The epoxidation of oleic acid

325

In this section, the epoxidation of oleic acid with H2O2 was tested using the HM-CSL-CLEAs and

326

CSL-CLEAs. The results showed that the final conversion yield was 55% after 96 h when

327

HM-CSL-CLEAs were used. However, only 36% of conversion yield was obtained at the same

328

reaction time when standard CSL-CLEAs were used (Fig.10). The results indicated that

329

HM-CSL-CLEAs had excellent stability of the HM-CSL-CLEAs against high concentrations of H2O2.

330

Therefore, the study not only further improves the catalytic properties of traditional CLEAs, but also

331

provides a robust and recyclable biocatalysis for the epoxidation of alkenes with low cost.

57

55

, Poly (lactic acid)/chitosan fiber mats

56

, and

. However, compared with magnetic CLEAs, the volumetric

332 333

Acknowledgments

334

The authors thank support from the National Natural Science Foundation of China (Grant Nos.

335

21676069, 21376249, 21336010), 973 Program (2013CB733604). National Major Scientific

336

Equipment Development Project (2013YQ14040508), Natural Science Foundation of Hebei Province,

337

China (project no. B2014208054), and the Key Foundation of Hebei University of Science and

338

Technology (project no. SW5).

339 340

16

ACS Paragon Plus Environment

Page 17 of 37

Journal of Agricultural and Food Chemistry

341

References

342

(1) Ho, K. P.; Wong, W. L.; Lam, K. M.; Lai, C. P.; Chan, T. H.; Wong, K. Y. A simple and effective

343

catalytic system for epoxidation of aliphatic terminal alkenes with manganese (II) as the catalyst,

344

Chem. Eur. J. 2008, 14, 7988-7996.

345

(2) Das, G.; Bordoloi, N.K.; Rai, S.K.; Mukherjee, A.K.; Karak, N. Biodegradable and biocompatible

346

epoxidized vegetable oil modified thermostable poly(vinyl chloride): Thermal and performance

347

characteristics post biodegradation with Pseudomonas aeruginosa and Achromobacter sp. J. Hazard.

348

Mater. 2012, 209, 434- 442.

349

(3) Coltro, L.; Pitta, J.B.; da Costa, R.A.; Perez, M.Â.F.; de Araǘjo, V.A.; Rodrigues, R. Migration of

350

conventional and new plasticizers from PVC films into food simulants: A comparative study. Food

351

Control. 2014, 44, 118-129.

352

(4) Ramachandra Murty, V.; Jayadev, B.; Muniswaran, P. K. A. Hydrolysis of oils by using

353

immobilized lipase enzyme: a review. Biotechnol. Bioprocess. Eng. 2002, 7, 57-66.

354

(5) Ahmed, E. H.; Raghavendra, T.; Madamwar, D. An alkaline lipase from organic solvent tolerant

355

Acinetobacter sp. EH28: Application for ethyl caprylate synthesis. Bioresour.Technol. 2010, 101,

356

3628-3634.

357

(6) Kilic, I.; Yesiloglu, Y. Lipase-catalyzed esterification of glycerol and oleic acid. J. Am. Oil Chem.

358

Soc. 2004, 81, 281-284.

359

(7) Gupta, M. N.; Sharma, S.; Shah, S. Biodiesl prepararation by lipase-catalyzed transesterification

360

of Jatropha oil. Energy Fuel, 2004, 18, 154-159.

17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

361

(8) Hernandez, K.; Berenguer-Murcia, A.; Rodrigues, R.C.; Fernandez-Lafuente, R. Hydrogen

362

peroxide in biocatalysis. A dangerous liaison. Curr. Org. Chem, 2012, 16, 2652-2672.

363

(9) Rocha-Martin, J.; Acosta, A.; Guisan, J. M.; Lopez-Gallego, F. Immobilizing systems biocatalysis

364

for the selective oxidation of glycerol coupled to in situ cofactor recycling and hydrogen peroxide

365

elimination. Chemcatchem. 2015, 7, 1939-1947.

366

(10) Ulrika, T.; Center, O. C.; Rajni, H. K.; Dietlind, A. Stability of immobilized Candida Antarctica

367

lipase B during chemo-enzymatic epoxidation of fatty acids. Enzyme Microb. Technol. 2007, 40,

368

447-451.

369

(11) Svedendahl, M.; Carlqvist, P.; Branneby, C.; Allner, O.; Frise, A.; Hult, K.; Berglund, P.; Brinck,

370

T. Direct epoxidation in Candida antarctica lipase B studied by experiment and theory, Chembiochem.

371

2008, 9, 2443-2451.

372

(12) Ranganathan, S.; Gartner, T.; Wiemann, L.; Sieber, V. A one pot reaction cascade of in situ

373

hydrogen peroxide production and lipase mediated in situ production of peracids for the epoxidation

374

of monoterpenes, J. Mol. Catal. B: Enzym, 2015, 114, 72-76.

375

(13) Rodrigues, R. C.; Ortiz, C.; Berenguer-Murcia, Á.; Torres, R.; Fernández-Lafuente, R.

376

Modifying enzyme activity and selectivity by immobilization. Chem. Soc. Rev. 2013, 42, 6290-6307.

377

(14) Quilles, J. C. J.; Brito, R. R.; Borges, J. P.; Aragon, C. C.; Fernadez-Lorente, G.;

378

Bocchini-Martins, D. A.; Gomes, E.; da. Silva, R.; Boscolo, M.; Guisan, J. M. Modulation of the

379

activity and selectivity of the immobilized lipases by surfactants and solvents. Biochem. Eng. J. 2015,

380

93, 274-280.

381

(15) Adlercreutz, P. Immobilisation and application of lipases in organic media. Chem. Soc. Rev, 2013,

18

ACS Paragon Plus Environment

Page 18 of 37

Page 19 of 37

Journal of Agricultural and Food Chemistry

382

42, 6406-6436.

383

(16) Ansorge-Schumacher, M. B.; Thum, O. Immobilised lipases in the cosmetics industry. Chem.

384

Soc. Rev. 2013, 42, 6475-6490.

385

(17) Ranjbakhsh, E.; Bordbar, A. K. Enhancement of stability and catalytic activity of immobilized

386

lipase on silia-coated modified magnetite nanoparticles. Chem. Eng. J. 2012, 179, 272-276.

387

(18) Mateo, C.; Palomo, J. M.; Fernandez-Lorente, G.; Guisan, J. M.; Fernandez-Lafuente, R.

388

Improvement of enzyme activity, stability and selectivity via immobilization techniques. Enzyme

389

Microb. Tech. 2007, 40, 1451-1463.

390

(19) Cao, L. Q. Immobilised enzymes: science or art? Curr. Opin. Chem. Biol. 2005, 9, 217-226 .

391

(20) Cui, J. D.; Jia, S. R. Optimization protocols and improved strategies of cross-linked enzyme

392

aggregates technology: current development and future challenges. Crit. Rev. Biotechnol. 2015, 35,

393

15-28.

394

(21) Torres, M. P. G.; Foresti, M. L.; Ferreira, M. L. Effect of different parameters on the hydrolytic

395

activity of cross-linked enzyme aggregates (CLEAs) of lipase from Thermomyces lanuginose.

396

Biochem. Eng. J. 2013, 72, 18-23.

397

(22) Cui, J. D.; Sun, L. M.; Li, L. L. A simple technique of preparing stable CLEAs of

398

phenylalanine ammonia lyase using co-aggregation with starch and bovine serum albumin. Appl.

399

Biochem. Biotechol. 2013, 170, 1827-1837.

400

(23) Garcia-Galan, C.; Berenguer-Murcia, A.; Fernandez-Lafuente, R.; Rodrigues, R.C. Potential of

401

different enzyme immobilization strategies to improve enzyme performance. Adv. Syn. Catal. 2011,

402

353, 2885-2904.

19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

403

(24) Wang, M. F.; Qi, W.; Yu, Q. X.; Su, R. X.; He, Z. M. Cross-linking enzyme aggregates in the

404

macropores of silica gel: a practical and efficient method for enzyme stabilization. Biochem. Eng. J.

405

2010, 52, 168-174.

406

(25) Cui, J. D.; Li L. L.; Zhao Y. M. Simple technique for preparing stable and recyclable

407

crosslLinked enzyme aggregates with crude-pored microspherical silica core, Ind. Eng. Chem. Res.

408

2014, 53, 16176-16182.

409

(26) Reshmi, R.; Sugunan, S. Improved biochemical characteristics of crosslinked β-glucosidase on

410

nanoporous silica foams. J. Mol. Catal. B: Enzym, 2013, 85,111-118.

411

(27) Cui, J. D.; Cui, L. L.; Zhang, S. P.; Zhang, Y. F.; Su, Z. G.; Ma, G. H. Hybrid magnetic

412

cross-linked enzyme aggregates of phenylalanine ammonia lyase from Rhodotorula glutinis. PLoS

413

ONE, 2014, 9, e97221.

414

(28) Bhattacharya, A.; Pletschke, B. I. Magnetic cross-linked enzyme aggregates (CLEAs): A novel

415

concept towards carrier free immobilization of lignocellulolytic enzymes. Enzyme. Microb. Tech.

416

2014, 61, 17-27.

417

(29) Talekar, S.; Ghodake, V.; Ghotage, T.; Rathod, P.; Deshmukh, P.; Nadar, S.; Mulla, M.; Ladole,

418

M. Novel magnetic cross-linked enzyme aggregates (magnetic CLEAs) of alpha amylase. Bioresour.

419

Technol. 2012, 123, 542-547.

420

(30) Kumar, V. V.; Sivanesan, S.; Cabana, H.; Magnetic cross-linked laccase aggregates

421

-Bioremediation tool for decolorization of distinct classes of recalcitrant dyes, Sci. Total Environ.

422

2014, 487, 830-839.

423

(31) Nadar, S. S.; Rathod, V. K.; Magnetic macromolecular cross linked enzyme aggregates (CLEAs)

20

ACS Paragon Plus Environment

Page 20 of 37

Page 21 of 37

Journal of Agricultural and Food Chemistry

424

of glucoamylase, Enzyme Microb. Tech. 2016, 83,78-87.

425

(32) Arca-Ramos, A.; Kumar, V.V.; Eibes, G.; Moreira, M.T.; Cabana, H. Recyclable cross-linked

426

laccase aggregates coupled to magnetic silica microbeads for elimination of pharmaceuticals from

427

municipal wastewater. Environ. Sci. Pollut. Res. 2016, 23, 8929-8939.

428

(33) Cruz-Izquierdo, Á..; Picó, E.A..; López, C..; Serra, J. L.; Llama, M. J. Magnetic Cross-Linked

429

Enzyme Aggregates (mCLEAs) of Candida antarctica lipase: An efficient and stable biocatalyst for

430

biodiesel synthesis. PLoS ONE. 2014, 9, e115202.

431

(34) Avishek, M.; Arun, G. Enhanced production of exocellular glucansucrase from Leuconostoc

432

dextranicum NRRL B-1146 using response surface method. Bioresour.Technol. 2008, 99, 3685-3691.

433

(35) Zhang, B. Zh.; Cui, J. D.; Zhao, G. X.; Jia, Sh. R. Modeling and optimization of phenylalanine

434

ammonia lyase Stabilization in recombinant E. coli for the continuous synthesis of L-phenylalanine

435

on the statistics-based experimental designs. J. Agric. Food. Chem. 2010. 58, 2795-2800.

436

(36) Yildirim, D.; Tükel, S. S.; Alagöz, D. Crosslinked enzyme aggregates of hydroxynitrilelyase

437

partially purified from Prunusdulcis seeds and its application for the synthesis of enantiopure

438

cyanohydrins. Biotechnol. Prog. 2014, 30, 818–827.

439

(37) Khanahmadi, S.; Yusof, F.; Amid, A.; Mahmod, S. S.; Mahat, M. K. Optimized preparation and

440

characterization of CLEA-lipase from cocoa pod husk. J. Biotech. 2015, 202,153-161.

441

(38) Mahmod, S .S.; Yusof, F.; Jami, M. S.; Khanahmadi, S. Optimizing the preparation conditions

442

and characterization of a stable and recyclable cross-linked enzyme aggregate (CLEA)-protease.

443

Bioresour. Bioprocess. 2016, 3, 3.

444

(39) Ranjbakhsh, E.; Bordbar, A. K. Enhancement of stability and catalytic activity of immobilized

21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

445

lipase on silia-coated modified magnetite nanoparticles. Chem. Eng. J. 2012, 179, 272-276.

446

(40) Galvis, M.; Barbosa, O.; Ruiz, M.; Cruz, J.; Ortiz, C.; Torres, R.; Fernandez-Lafuente, R.

447

Chemical amination of lipase B from Candida antarctica is an efficient solution for the preparation of

448

crosslinked enzyme aggregates. Process. Biochem. 2012, 47, 2373-2378.

449

(41) Noble, J. E.; Bailey, M. J. A. Quantitation of protein. Methods Enzymol. 1990, 182, 50-69.

450

(42) Kwon, Ch. W.; Park, K. M.; Choi, S. J.; Chang, P. Sh. A reliable and reproducible method for the

451

lipase assay in an AOT/isooctane reversed micellar system: Modification of the copper-soap

452

colorimetric method. Food Chem, 2015, 182, 236-241.

453

(43) Lowry, R. R.; Tinsley, I. J. Rapid colorimetric determination of free fatty acids. J. Am. Oil Chem.

454

Soc. 1976, 53, 470-472.

455

(44) Li, Y.; Fei, G.; Wei, W.; Jian, B. Q.; Guang, H. M.; Wei, Q. Zh. Pore size of macroporous

456

polystyrene microspheres affects lipase immobilization. J. Mol. Catal. B-Enzym. 2010, 66, 182-189.

457

(45) Yu, H.; Chen, H.; Wang, X.; Yang, Y.; Ching, C. Cross-linked enzyme aggregates (CLEAs) with

458

controlled particles: application to Candida rugosa lipase. J. Mol. Catal. B: Enzym. 2006, 43,

459

124–127.

460

(46) Torres, M.d.P.G., Foresti, M.L., Ferreira, M.L., 2013. Cross-linked enzyme aggregates (CLEAs)

461

of selected lipases: a procedure for the proper calculation of their recovered activity. AMB Express.

462

2013, 3, 1–11.

463

(47) Sheldon, R.A. Characteristic features and biotechnological applications of cross-linked enzyme

464

aggregates (CLEAs). Appl. Microbiol. Biotechnol. 2011, 92, 467–477.

465

(48) Dong, T.; Zhao, L.; Huang, Y.; Tan, X. Preparation of cross-linked aggregates of aminoacylase

22

ACS Paragon Plus Environment

Page 22 of 37

Page 23 of 37

Journal of Agricultural and Food Chemistry

466

from Aspergillusmelleus by using bovine serum albumin as an inert additive. Bioresour. Technol.

467

2010, 101, 6569–6571.

468

(49) Reddy, L.; Wee, Y. J.; Yun, J. S.; Ryu, H. W. Optimization of alkaline protease production by

469

batch culture of Bacillus sp. Rky3 through plackett-burman and response surface methodological

470

approaches. Bioresour. Technol. 2008, 99, 2242–2249.

471

(50) Muralidhar, R.; Chirumamila, R.; Marchant, R.; Nigam, P. A. response surface approach for the

472

comparison of lipase production by Candida cylindracea using two different carbon sources.

473

Biochem. Eng. J. 2001, 9, 17–23.

474

(51) Xu, D. Y.; Yang, Y.; Yang, Z. Activity and stability of cross-linked tyrosinase aggregates in

475

aqueous and nonaqueous media. J. Biotechnol. 2011, 152, 30–36.

476

(52) Zhang, W. W.; Yang, X. L.; Jia, J. Q.; Wang, N.; Hu, Ch. L.; Yu, x. Q. Surfactant-activated

477

magnetic cross-linked enzyme aggregates (magnetic CLEAs) of Thermomyces lanuginosus lipase for

478

biodiesel production. J. Mol. Catal. B-Enzym. 2015, 115, 83-89.

479

(53) Cui, J. D.; Liu, R. L.; Li, L. B.

480

bovine pancreatic lipase using bovine serum albumin as an additive. Korean J. Chem. Eng. 2016. 33,

481

610-615.

482

(54) Lai, J. Q.; Hu, Zh. L.; Sheldon, R. A.; Yang, Zh. Catalytic performance of cross-linked enzyme

483

aggregates of Penicillium expansum lipase and their use as catalyst for biodiesel production. Process.

484

Biochem. 2012, 47, 2058-2063.

485

(55) Yu, W. H.; Tong, D. Sh.; Fang, M.; Shao, P.; Zhou, Ch. H. Immobilization of Candida rugosa

486

lipase on MSU-H type mesoporous silica for selective esterification of conjugated linoleic acid

A facile technique to prepare cross-linked enzyme aggregates of

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

487

isomers with ethanol. J. Mol. Catal. B-Enzym. 2015, 111, 43-50.

488

(56) Siqueira, N. M.; Garcia, K. C.; Bussamara, R.; Both, F. S.; Vainstein, M. H.; Soares, R. M. D.

489

Poly (lactic acid)/chitosan fiber mats: Investigation of effects of the support on lipase immobilization.

490

Int. J. Bio.Macromol. 2015, 72, 998-1004.

491

(57) Fujiwara, M.; Shiokawa, K.; Yotsuya, K.; Matsumoto, K. Immobilization of lipase from

492

Burkholderia cepacia into calcium carbonate microcapsule and its use for enzymatic reactions in

493

organic and aqueous media. J. Mol. Catal. B-Enzym. 2014, 109, 94-100.

494 495 496 497 498 499 500 501 502 503 504 505 506 507

24

ACS Paragon Plus Environment

Page 24 of 37

Page 25 of 37

Journal of Agricultural and Food Chemistry

508

Caption of Figures

509

Figure 1 General scheme for hybrid magnetic CLEAs.

510

Figure 2 SEM images of (a) CSL-CLEAs magnified 30000×; (b) HM-CSL-CLEAs

magnified

511

30000×; as well as confocal laser scanning microscopy (LCSM) images of (c) magnetite

512

nanoparticles with

513

treatment.

fluorescamine treatment and (d) HM-CSL-CLEAs with fluorescamine

514

Figure 3 Magnetisation curves of immobilized lipase and the separation process of immobilized

515

lipase from solution by magnet. (a) Immobilized lipase was dispersed in solution; (b)

516

immobilized lipase was collected by an external magnet.

517

Figure 4 2D and 3D contour plots at center point. 2D and 3D plot of (NH4)2SO4 concentration and

518

glutaraldehyde concentration (A, B); 2D and 3D plot of (NH4)2SO4 concentration and

519

glutaraldehyde concentration (C, D); 2D and 3D plot of glutaraldehyde concentration and

520

enzyme concentration (E, F).

521

Figure 5 The effect of pH on the activity of free enzyme, CSL-CLEAs, and HM-CSL-CLEAs.

522

Figure 6 Influence of temperature on activity of free enzyme, CSL-CLEAs, and HM-CSL-CLEAs.

523

Figure 7 The thermal stability of free enzyme, CSL-CLEAs, and HM-CSL-CLEAs at 60 oC.

524

Figure 8 Storage stability of the free enzyme, CSL-CLEAs, and HM-CSL-CLEAs.

525

Figure 9 Reusability of the CSL-CLEAs and HM-CSL-CLEAs.

526

Figure 10 The conversion rates of CSL-CLEAs and HM-CSL-CLEAs for the epoxidation of oleic

527

acid with H2O2

528

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 37

529 530 531

Table 1 Experimental range levels of enzyme concentration, (NH4)2SO4 concentration,

532

and glutaraldehyde concentration in terms of actual factors

Variable

Symbols

Coded-variable levels -1.68

-1

0

1

1.68

Enzyme concentration(mg/mL)

X1

1.33

1.50

1.75

2.00

2.17

(NH4)2SO4 concentration (%)

X2

1.16

1.50

2.00

2.50

2.84

Glutaraldehyde concentration (%)

X3

0.12

0.15

0.20

0.25

0.28

533 534 535 536 537 538 539 540

26

ACS Paragon Plus Environment

Page 27 of 37

Journal of Agricultural and Food Chemistry

+ Fe3O4

Aggregation Cross-linking

Enzyme

Cross-linker

Figure 1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 2

ACS Paragon Plus Environment

Page 28 of 37

Page 29 of 37

Journal of Agricultural and Food Chemistry

Figure 3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 4

ACS Paragon Plus Environment

Page 30 of 37

Page 31 of 37

Journal of Agricultural and Food Chemistry

Figure 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 6

ACS Paragon Plus Environment

Page 32 of 37

Page 33 of 37

Journal of Agricultural and Food Chemistry

Figure 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

120

Free enzyme CSL-CLEAs HM-CSL-CLEAs

100

Residual activity(%)

Page 34 of 37

80 60 40 20 -5

0

5

10

15

20

25

30

35

Time(d) Figure 8

ACS Paragon Plus Environment

40

45

Page 35 of 37

Journal of Agricultural and Food Chemistry

Figure 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 10

ACS Paragon Plus Environment

Page 36 of 37

Page 37 of 37

Journal of Agricultural and Food Chemistry

Graphic Abstract

ACS Paragon Plus Environment