α-Amylase@magnetite: magnetic nanocomposites with enhanced

Jul 5, 2018 - Alpha-amylase was entrapped within magnetite derived xerogel matrix in a course of a room-temperature sol-gel transition, leading to enz...
0 downloads 0 Views 5MB Size
Subscriber access provided by Stony Brook University | University Libraries

Biotechnology and Biological Transformations

#-Amylase@magnetite: magnetic nanocomposites with enhanced thermal stability for starch hydrolysis Bazhena V. Astafyeva, Olga E. Shapovalova, Andrey S. Drozdov, and Vladimir V. Vinogradov J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b01298 • Publication Date (Web): 05 Jul 2018 Downloaded from http://pubs.acs.org on July 6, 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 19

Journal of Agricultural and Food Chemistry

α-Amylase@ferria: magnetic nanocomposites with enhanced thermal stability for starch hydrolysis Bazhena V. Astafyeva, Olga E. Shapovalova, Andrey S. Drozdov,∗ and Vladimir V. Vinogradov Laboratory of Solution Chemistry of Advanced Materials and Technologies, ITMO University, St. Petersburg, 191002, Russian Federation E-mail: [email protected]

1

Abstract

2

The present study is devoted to the development of a new class recyclable magnetic

3

catalytic nanocomposites for starch hydrolysis. Alpha-amylase was entrapped within

4

magnetite derived xerogel matrix in a course of a room-temperature sol-gel transition,

5

leading to enzyme immobilization within the pores of a rigid magnetic matrix. For

6

hybrid organo-inorganic composites with enzyme mass fractions less than 10 wt. %

7

no enzyme leaching was observed. At 80◦ C amylase@ferria composite demonstrates

8

catalytic activity on the level of 10 U/mg and starch hydrolysis rate comparable to

9

free enzyme, while at 90◦ C the activity of amylase@ferria at least twice higher that

10

of free amylase due to higher thermal stability of the composite. Entrapped amylase

11

showed excellent stability and lost only 9 % of its activity after 21 days of storage in a

12

buffer solution, while free enzyme was totally inactivated after 17 days. The material

13

can be used either a magnetically-separable reusable catalyst or can be applied as a

14

catalytic ceramic coating with at least 10 cycles of use.

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

15

Introduction

16

Advance in nanomaterials synthesis and nanotechnology led to active implementation of

17

nanostructured materials in different areas of industry as adsorbents, catalysts or construc-

18

tion materials. 1–4 When it comes to food industry, the main attention is paid to enzyme

19

immobilization, as it can significantly reduce costs of enzymatically catalyzed reactions by

20

thermal and chemical stabilization of enzymes, simplifying separation of enzymes from reac-

21

tion mixtures and offering the possibility to make recyclable enzymatic catalysts. 5–8 Selection

22

of a support material can have a critical effect on the stability of enzymes and the efficiency

23

of enzyme immobilization, although it is difficult to predict in advance which support will be

24

the most suitable for a particular enzyme. The support must be insoluble in water, should

25

have high enzyme loading capacity, be mechanically stable and must not have deleterious

26

effect on the catalytic activity of the enzyme. The variety of nanostructured materials for

27

immobilization of enzymes includes both organic and inorganic polymers and materials, such

28

as: porous polymeric matrixes, 9,10 polystyrene particles, 8,11,12 phospholipid liposomes, 13–15

29

zeolites, 16,17 metal oxides, 18,19 silica matrixes 20–22 and ect. Among all enzymatic nanocom-

30

posites significant attention is paid to ones with magnetic properties. 23,24 Magnetic field

31

susceptibility revealed a mechanism for efficient recovery of the enzyme complex thereby

32

preventing the contamination of the final product by enzymatic catalyst. Magnetic enzy-

33

matic catalysts mainly presented by either core-shell type structures or polymeric matrixes

34

with encapsulated magnetic nanoparticles and are known to be effective systems with good

35

application results. 25–30 Despite the effectiveness, magnetic composites shares one common

36

disadvantage of their complex structure that involves a magnetic core, an enzyme and a

37

linking agent. Such multicomponent structure requires multi-step synthesis procedures with

38

intermediate purification steps resulting in higher production costs that can negate the prac-

39

tical advantage of such systems. In this article, we describe magnetic nanocomposite with

40

amylase activity consisting of only two components: amylase and magnetite ceramic matrix.

41

The material is produced by direct entrapment of the enzyme within magnetite xerogel in a 2

ACS Paragon Plus Environment

Page 2 of 19

Page 3 of 19

Journal of Agricultural and Food Chemistry

42

course of the room-temperature sol-gel transition. Entrapped amylase not only demonstrated

43

the ability to perform catalytic hydrolysis of starch solutions with the rate comparable to that

44

of the free enzyme, but also showed excellent thermal stability and extended temperature

45

of activity window and prolonged period of storage. Composites demonstrated leach-proof

46

behavior and can be magnetically separated and reused up to 10 times with only 5% decrease

47

of its initial activity. Alternatively the composite can be applied in a form of ceramic coating

48

to ensure effective application in batch or flow reactors.

49

Materials and methods

50

Chemicals: Iron (II) chloride tetrahydrate, iron (III) chloride hexahydrate, ammonia 25%

51

solution, α-amylase from porcine pancreas (100 units/mg) and soluble starch, all obtained

52

from Sigma-Aldrich.

53

Preparation of a ferria hydrosol: Pure ferria hydrosol was prepared from iron (II) chloride

54

tetrahydrate and iron (III) chloride hexahydrate as described in the Ref. 31. Briefly, 2.5

55

g FeCl2 *4H2 O and 5 g FeCl3 *6H2 O were dissolved in 100 mL of deionized water under

56

constant stirring (500 rpm). Then, 12 mL of aqueous ammonia solution was added at room

57

temperature. Using a magnet, the formed magnetite precipitate was collected and washed

58

with deionized water until neutral pH. The washed black precipitate was mixed with 100 mL

59

of deionized water and subjected to ultrasonic treatment (37 kHz, 110 W) under constant

60

stirring (300 rpm) for 120 minutes. The mass fraction of magnetite nanoparticles in the

61

resulting sol was 2 wt. %.

62

Entrapment of α-amylase in a ferria matrix: 20 mL of freshly prepared ferria sol was

63

mixed with 0 to 1250 µL of α-amylase solution (4000 U/mL). After stirring, the contents

64

were dried in a vacuum desiccator to yield composites with enzyme mass fractions ranging

65

from 0 to 12.5 wt. % (0-11 U/mg). The composite materials were milled in an agate mortar

66

and used in experiments. To evaluate the release profile of the enzyme, 20 mg of the crushed

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

67

composite in a quartz cuvette was treated with 2 mL of the saline solution and the absorption

68

spectrum at 37◦ was measured over time at 210 nm.

69

Amylase@ferria coatings: In order to prepare the catalytic coating mixture of 5 mL of

70

freshly prepared ferria sol with 1 mL of α-amylase solution (4000 U/mL) was applied on

71

the walls of a glass and condensed on air. The glass was continuously rotated to ensure the

72

homogeneity of the coating.

73

Preparation of the starch solution: 2 g of soluble starch were placed in 98 mL of water

74

preheated to 55◦ and thoroughly stirred. The temperature was gradually elevated to 80◦ with

75

continuous stirring using a magnetic stirrer for an hour. After incubation the temperature

76

of the solution was decreased to 60◦ , it was stirred for another 30 minutes and used in the

77

experiments.

78

Evaluation of amylase@ferria particles enzymatic activity: Activity of the amylase@ferria

79

composite was evaluated by measuring the viscosity of the starch solution. 7 mL of the starch

80

solution was placed into a thermostated rotary viscometer sleeve and, after adding 100 mg

81

of the 10 wt. % amylase@ferria composite (enzymatic activity 10 U/mg), the viscosity

82

was measured in the kinetic mode at different temperatures ranging from 60 to 90◦ . For

83

comparison, 7 mL of the starch solution was incubated with 25 µL of amylase solution (4000

84

U/mL) and subsequently tested in the same manner.

85

Evaluation of amylase@ferria particles reusability: In order to test the reusability of the

86

composite magnetic catalyst 100 mg of the 10 wt. % amylase@ferria composite (enzymatic

87

activity 10 U/mg) was added to 7 mL of the starch solution and stirred for 20 minutes while

88

monitoring the process by measuring the viscosity of the solution. After that, the composite

89

material was separated by applying magnetic field and added to a new portion of the starch

90

solution for another catalytic cycle.

91

Catalytic activity of amylase@ferria coatings: Into a model reactor with the total volume

92

1 L, coated with amylase@ferria nanocomposite, 100 mL of the starch solution was added

93

and mixed at 80◦ for 15 minutes. The process was monitored by measuring the viscosity.

4

ACS Paragon Plus Environment

Page 4 of 19

Page 5 of 19

Journal of Agricultural and Food Chemistry

94

After the incubation, the solution was changed with a new portion of the starch solution and

95

another catalytic cycle was performed.

96

Characterization methods: The crystalline phase of the samples was measured by X-

97

ray diffraction (Bruker D8 Advance) using Cu-Kα radiation (λ = 1.54 ˚ A); the samples

98

were scanned at a rate of 0.5 degrees per minute. To analyze the samples with a high-

99

resolution scanning electron microscopy (SEM), the obtained ground xerogel was deposited

100

on a metal tip and investigated without additional spraying using a Magellan 400L ultra-high

101

resolution electron microscope. The samples for a transmission electron microscopy (TEM)

102

were obtained by dispersing a small probe in ethanol to form a homogeneous suspension.

103

Then, a suspension drop was coated on a copper mesh covered with carbon for a TEM

104

analysis (FEI TECNAI G2 F20, at an operating voltage of 200 kV). Specific surface area,

105

pore volume and pore size distribution were investigated using Quantachrome Nova 1200e by

106

nitrogen adsorption at 77 K and analyzed by the BET and BJH equations. Prior to analysis,

107

all samples were degassed at room temperature for 48 hours. Hydrodynamic diameter was

108

measured by the DLS technique on Photocor Compact Z. The Raman spectra were recorded

109

using the 633 nm He-Ne laser line on a Horiba Jobin Yvon Micro Raman 300 spectrometer.

110

The laser power on the samples employed was 0.03 mW and 0.344 mW, with 300 s and 120

111

s exposition per diffraction window, respectively. In all the measurements, 50x Olympus

112

lens, hole of 500mm, slit of 100mm and a diffraction grid with 1800 grooves/mm were

113

employed. The IR measurements were performed in transmission and ATR modes. Viscosity

114

was determined by a Fungilab rotational viscometer.

115

Results and discussion

116

Synthesis of amylase@ferria composites

117

In order to produce amylase@magnetite composites, the stable magnetite hydrosol (desig-

118

nated as ferria) was used that was described in details in our previous works. 31 The hydrosol 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

119

was synthesized by ultrasound-assisted coprecipitation procedure using non-stoichiometric

120

ratios of iron(II) and iron(III) salts with the subsequent ultrasonic treatment of the resulting

121

sol in deionized water. The final hydrosol consists of magnetite nanoparticles with an aver-

122

age hydrodynamic diameter of 30 nm dispersed in water at neutral pH level (pH 7). Surface

123

of the particles is densely covered with the hydroxyl groups both of magnetite and Fe(OH)2

124

nature as proven by Raman spectroscopy (Fig. 1a). Due to this fact, the isoelectric point

125

of the magnetite NPs is shifted from pH = 6.8, which is common for magnetite, to a higher

126

pH value of 8.3 (Fig. 1b). As a result, the magnetite NPs showed high zeta-potential values

127

(+36 mV at neutral pH level) and excellent colloidal stability, forming stable hydrosols which

128

are behaves as typical magnetic fluids (Fig. 1c).

Figure 1: Ferria hydrosol. The ATR spectra demonstrates the presence of the hydroxyl groups of both Fe(OH)2 and Fe3 O4 nature (a); in contrast to magnetite, the isoelectric point of the ferria nanoparticles is shifted to pH = 8.3 (b); due to the high zeta potential values the ferria nanoparticles form a stable hydrosol with a magnetic fluid-like behavior at a neutral pH level (c).

129

Amylase@ferria composites were produced by the condensation of the ferria hydrosol

130

with α-amylase solution under reduced pressure at room temperature. Upon removal of

131

the solvent, a sol-gel transition occured leading to the formation of the mesoporous xerogel

132

matrix. The matrix was composed of the 10 nm magnetite nanoparticles with a truncated

133

octahedron morphology according to SEM and TEM images (Fig. 2a and 2b). The XRD

134

pattern of the material demonstrated the lines typical for magnetite (compared to JCPDS

135

file no. 19-0629) (Fig. 2c). The main peak at 35.45◦ attributed to the crystalline plane with 6

ACS Paragon Plus Environment

Page 6 of 19

Page 7 of 19

Journal of Agricultural and Food Chemistry

136

Miller indices of (3 1 1) is clearly seen. Other distinctive peaks at 18.52◦ - (0 2 0), 30.10◦ -

137

(2 2 0), 43.08◦ - (4 0 0), 53.45◦ - (4 2 2), 56.98◦ - (5 1 1), 62.57◦ - (4 4 0), 70.99◦ - (5 3 3),

138

and 74.02◦ - (4 4 4) (Fig. 2f) matched those of the standard magnetite diffraction pattern. 32

139

Another proof that the ferria xerogel matrix consisted of the magnetite nanoparticles came

140

from Raman spectroscopy. The Raman spectra of the sample showed the characteristic

141

broad band of magnetite at 667 cm−1 (Fig. 2d, line a) when the spectra was recorded with

142

a 633 nm He-Ne laser (0.030 mW laser beam power and an acquisition time of 300 sec per

143

diffraction window). 33,34 Further confirmation came from observations of the known phase

144

transformations of magnetite under the Raman-laser measurement conditions: when the

145

power was increased to 0.344 and 1.35 mW (acquisition times of 90 and 30 sec, respectively),

146

then for the former the characteristic maghemite bands at 711, 492, and 372 cm−1 are seen

147

(Fig. 2d, line b), and for the latter full oxidation to α-hematite is observed (Fig. 2d, line c)

148

by the characteristic bands at 591, 385, 273, and 213 cm−1 . 33,34

149

The formed matrix possesses well-developed mesoporous nanoarchitecture (Fig. 2a,b)

150

with a high specific surface area (110 m2 /g for 10 wt. % amylase@ferria and 119 m2 /g

151

for pure ferria xerogel) and an average pore diameter of 8 nm (Fig. 1S(a,b)). While the

152

molecule of amylase has the size of 7x6x5 nm, 35 it is complementary to the pores diameter

153

of ferria matrix, and it is entrapped in the course of a sol-gel transition. The mechanism

154

of the entrapment can be briefly described as follows: at the first step, negatively charged

155

amylase molecules (isoelectric point of amylase is at pH = 5.5) 35 electrostatically interacts

156

with positively charged magnetite nanoparticles (isoelectric point of magnetite sol is at pH

157

= 8, Fig. 1b). Upon drying of the resulting mixture, the rigid magnetite xerogel matrix is

158

formed (Fig. 2a,b), in which enzyme molecules may be buried inside of the matrix or located

159

in the surface layer. According to our previous observations, such structural organization

160

of a material allows the entrapped enzyme to interact with macromolecular substrates, such

161

as polysaccharides and to perform enzymatic transformations without being released from

162

a xerogel matrix. 36,37 Due to the formation of interparticle Fe-O-Fe bonds in the process of

7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 2: The ferria xerogel matrix. The ferria matrix consists of nanoparticles with a truncated tetrahedron morphology. Interplanar spacing is shown in the insert (a); SEM image of the ferria xerogel demonstrates porous structure of the material (b); XRD patternt of the ferria xerogel shows peaks typical for the magnetite crystal phase (c); Raman spectra recorded at a 0.03 mW laser beam power demonstrates spectra typical for magnetite (line a), after increasing the laser power to 0.344 mW magnetite partially converts into maghemite (line b), at 1.35 mW complete conversion of magnetite into hematite is observed (line c). All lines are obtained from a single sample (d); the magnetization curve of ferria and amylase@ferria. Absence of hysteresis can be observed in the insert (e); the material is readily attracts to the magnet (f). 163

sol-gel transition, the formed composite matrix is not prone to spontaneous re-suspendeition

164

in water media and can be used as a magnetically-sensitive catalytic composite. 38 The mate-

165

rial demonstrated the typical superparamagnetic behavior: its magnetization curve showed

166

absence of the hysteresis loop and no remnant magnetization in absence of magnetic field

167

(Fig. 2e). The magnetic moment of the composite at 6000 Oe amounted to 61 emu/g for

168

10% amylase@ferria, that is 25% lower than for the pure ferria matrix. Nevertheless, a high

169

magnetic susceptibility allowed to separate the amylase@ferria composite using an external

170

magnet (Fig. 2f). The entrapment of amylase may be either full or there may be partial

171

release from the composite depending on the mass fraction of the enzyme. According to 8

ACS Paragon Plus Environment

Page 8 of 19

Page 9 of 19

Journal of Agricultural and Food Chemistry

172

release studies, at mass fractions of entrapped amylase below 10 wt. %, only a minor release

173

of approximately 1% of the total entrapped enzyme amount occurred, while at higher mass

174

fractions release was increased up to 21% of the total amount due to matrix enzyme loading

175

capacity overload (Fig. 2S). 36

176

Catalytic activity of amylase@ferria

177

To evaluate the enzymatic activity of the synthetized biocomposites, the enzymatic starch

178

cleavage reaction was selected. Amylase acts on α-1,4-glycosidic bonds of starch, cleav-

179

ing polysaccharides inside the chain and leading to formation of low-molecular hydrolysis

180

products, normal α-dextrins. Further hydrolysis gives maltose, maltotriose, and glucose. 39

181

Since the hydrolysis process results in biopolymer cleaving to smaller monomers, the viscos-

182

ity of the system decreases, and this phenomenon serves as a convenient analytical signal

183

to describe the catalytic process. Amylase is a thermally stable enzyme and has an opti-

184

mal temperature (Topt ) valued 70◦ C. To evaluate the catalytic activity of the composites,

185

amylase@ferra with 10 wt. % was selected and incubated with the starch solution at tem-

186

peratures varied from 60 to 90◦ C. For comparison, the process was performed using equal

187

amounts of free enzyme (Fig. 3). Evaluation of the process at 60◦ C demonstrated that the

188

amylase@ferria composite demonstrates catalytic activity and is able to hydrolyze the starch

189

solution. Compared to free amylase, the composite material showed approximately 1.5-fold

190

lower activity under these conditions, hydrolyzing starch in 34 minutes, while free enzyme

191

performed this process in 22 minutes (Fig. 3a). With elevation of the temperature to 70◦ C

192

accelerated the reaction rate of the both systems, but also makes the difference between free

193

enzyme and amylase@ferria even more pronounced: while free enzyme catalyzes the process

194

in 5 minutes, for the amylase@ferria composite the rate of the reaction was 2.5-folds lower

195

and required 13 minutes for hydrolysis of the same amount of the substrate (fig. 3b). The

196

difference between the reaction rates can be easily explained by the fact that in case of the

197

free amylase the reaction takes place in a homogeneous solution, while for the immobilized 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

198

enzyme the system is heterogeneous and diffusion processes limits the reaction. While amy-

199

lase has an optimal temperature of 70◦ C, further elevation of the temperature to 80◦ C led to

200

its denaturation and inactivation processes. This is true for the free enzyme, and a notable

201

decrease in the reaction rate is observed, with prolongation of the hydrolysis time from 5

202

min at 70◦ C to 14 min at 80◦ C. In contrast, entrapped amylase showed the catalytic activity

203

almost identical to that at 70◦ C and close to the activity of the free enzyme (Fig. 3c). Fur-

204

ther increase of the reaction temperature to 90◦ C made the difference in the behavior of the

205

catalysts more obvious. While in the case of free amylase the starch hydrolysis was halted

206

due to the denaturation, the composite material demonstrated a high degree of catalytic

207

activity compared to the activity measured at 70 and 80◦ C (Fig. 3d). Preservation of the

208

catalytic activity by entrapped amylase at elevated temperatures is in good correlation with

209

our previous observations. 40 This phenomenon can be explained by tight interactions of the

210

enzyme’s tertiary structure with the inorganic matrix pore walls, resulting in stabilization

211

of the protein structure and its stabilization against harsh conditions.

212

One of the features of the amylase@ferria composite material is its magnetic responsive-

213

ness. The composite material can be separated from the reaction mixture by an external

214

magnetic field and be reused for another cycle of the catalytic process. Experiments showed

215

that the activity of the composite material remains essentially at the same level at least after

216

ten cycles of use (Fig. 3e), decreasing by only 4% in total and revealing a high stability of

217

the composite material and a prolonged profile of action, characteristic for the entrapped

218

enzymatic systems, when the activity is not related to the release of the enzyme. Another

219

important parameter of enzymatic formulations and composite materials is their storage sta-

220

bility. By their nature, enzymes are unstable in solutions and deteriorate over time. One of

221

the goals that can be achieved though enzyme immobilization is an increase of its stability

222

and prolongation of the storage period. 22,41 In order to evaluate the effect of entrapment

223

within sol-gel magnetite in amylase long-term stability the material was stored in phosphate

224

buffer solution (pH = 7.4) at 4◦ C for a period of 21 days and the activity was compared

10

ACS Paragon Plus Environment

Page 10 of 19

Page 11 of 19

Journal of Agricultural and Food Chemistry

225

to one measured for free enzyme (Fig. 3f). Entrapped amylase demonstrated significantly

226

higher stability at storage conditions compared to free enzyme. While free amylase lost all

227

its activity within 17 days, composite material demonstrates 91% of its initial activity after

228

21 days of storage with only 0.5%/day degradation rate. When compared to results reported

229

earlier it can be seen that immobilization of amylase by entrapment results in higher stability

230

compared to covalent bonding or adsorption. 22,42–45 This fact can be explained by tighter

231

three dimensional interaction of the enzyme with inorganic matrix and absence of covalent

232

bonds between the supporting material and the enzyme that can affect the conformation of

233

enzyme. 40

234

Amylase@magnetite ceramic catalytic coatings

235

As an alternative to application as a magnetically separable catalyst, amylase@magnetite

236

composite can be applied as a ceramic coating on a surface of a reactor. For this purpose,

237

mixture of amylase and stable magnetite hydrosol was applied on the vessel surface before

238

solvent removal by spraying, dip-coating or by using the Mayer rod. Upon condensation,

239

irreversible sol-gel transition occurred resulted in the formation of the enzymatically-doped

240

ceramic magnetite coating on the surface of the substrate. The thickness of the formed

241

ceramic layer depends on the amount of the applied colloid solution by the factor of 24,

242

(that is, application of the 5 micrometer layer of hydrosol mixture resulted in the formation

243

of the composite ceramic coating with the thickness of ≈210 nm) and can be precisely

244

regulated in that way. The formed coating showed a high degree of the homogeneity (Fig.

245

4a).

246

The composite bioceramic coating demonstrated high thermal stability similar to the

247

one observed on the magnetic particles (Fig. 4b). The relative activity of the composite

248

coating was stable at elevated temperatures without notable degradation at 70◦ C, Topt of

249

α-amylase, and only slight degradation of the material for 4% after incubation at 80◦ C after

250

90 min of incubation was observed. At 90◦ the enzymatic activity linearly decreased over 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

251

the time, loosing 0.37% of its relative activity per minute, but still demonstrated higher

252

thermal stability than free enzyme, which was fully inactivated in 5 minutes at this tem-

253

perature. The performance of the model reactor was tested by evaluating the rate of starch

254

hydrolysis in successive catalytic cycles. For this purpose, the starch solution was subjected

255

to hydrolysis in batch experiments in the reactor at 80◦ C, the process was monitored by

256

measuring the viscosity of the solution every 5 minutes (Fig. 4c). When the viscosity of

257

solution reached the constant value, the starch solution was changed with a new portion and

258

the hydrolysis dynamic was measured again (for details see Experimental section). Viscosity

259

measurements in the set of experiments demonstrated the excellent stability of the model re-

260

actor performance in time. Starch solution hydrolysis demonstrated almost equivalent rates

261

in 10 successive experiments and took 20 minutes at 80◦ C with only minor degradation by

262

4%, which demonstrated the high stability of the hybrid amylase@ferria coating catalytic

263

properties. It can be concluded that sych and makes it promising material for application

264

in design of new types of advanced batch and flow reactors.

265

Acknowledgement

266

This work was supported by Russian Foundation for Basic Research, grant 18-33-01170

267

mol a. The authors declare no conflict of interests.

268

References

269

270

271

272

273

(1) Sanchez, C.; Belleville, P.; Popall, M.; Nicole, L. Chemical Society Reviews 2011, 40, 696–753. (2) Piccinno, F.; Gottschalk, F.; Seeger, S.; Nowack, B. Journal of Nanoparticle Research 2012, 14, 1109. (3) Lee, J.; Mahendra, S.; Alvarez, P. J. ACS nano 2010, 4, 3580–3590. 12

ACS Paragon Plus Environment

Page 12 of 19

Page 13 of 19

Journal of Agricultural and Food Chemistry

274

(4) Pansini, M.; Dell’Agli, G.; Marocco, A.; Netti, P. A.; Battista, E.; Lettera, V.; Ver-

275

gara, P.; Allia, P.; Bonelli, B.; Tiberto, P. Journal of Biomedical Nanotechnology 2017,

276

13, 337–348.

277

278

279

280

(5) Cushen, M.; Kerry, J.; Morris, M.; Cruz-Romero, M.; Cummins, E. Trends in Food Science & Technology 2012, 24, 30–46. (6) Rashidi, L.; Khosravi-Darani, K. Critical reviews in food science and nutrition 2011, 51, 723–730.

281

(7) Lin, Y.; Ren, J.; Qu, X. Accounts of chemical research 2014, 47, 1097–1105.

282

(8) Verma, M. L.; Barrow, C. J.; Puri, M. Applied microbiology and biotechnology 2013,

283

284

285

286

287

97, 23–39. (9) Wang, Z.-G.; Wan, L.-S.; Liu, Z.-M.; Huang, X.-J.; Xu, Z.-K. Journal of Molecular Catalysis B: Enzymatic 2009, 56, 189–195. (10) Jia, H.; Zhu, G.; Vugrinovich, B.; Kataphinan, W.; Reneker, D. H.; Wang, P. Biotechnology progress 2002, 18, 1027–1032.

288

(11) Ge, J.; Lu, D.; Liu, Z.; Liu, Z. Biochemical Engineering Journal 2009, 44, 53–59.

289

(12) Filippusson, H.; Hornby, W. Biochemical Journal 1970, 120, 215–219.

290

(13) Yoshimoto, M. Enzyme Stabilization and Immobilization: Methods and Protocols 2017,

291

292

293

9–18. (14) Liu, W.; Ye, A.; Singh, H. Microencapsulation and Microspheres for Food Applications. Sagis LMC, ed. Academic Press, New York, NY, USA. p 2015, 151–170.

294

(15) Koshel, E. I.; Chelushkin, P. S.; Melnikov, A. S.; Serdobintsev, P. Y.; Stolbovaia, A. Y.;

295

Saifitdinova, A. F.; Shcheslavskiy, V. I.; Chernyavskiy, O.; Gaginskaya, E. R.; Ko-

13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

296

shevoy, I. O. Journal of Photochemistry and Photobiology A: Chemistry 2017, 332,

297

122–130.

298

299

300

301

302

303

304

305

306

307

308

309

310

311

(16) Awasthi, M. K.; Pandey, A. K.; Bundela, P. S.; Wong, J. W.; Li, R.; Zhang, Z. Bioresource technology 2016, 213, 181–189. (17) Liang, K.; Coghlan, C. J.; Bell, S. G.; Doonan, C.; Falcaro, P. Chemical Communications 2016, 52, 473–476. (18) Shi, X.; Gu, W.; Li, B.; Chen, N.; Zhao, K.; Xian, Y. Microchimica Acta 2014, 181, 1–22. (19) Chapurina, Y. E.; Drozdov, A. S.; Popov, I.; Vinogradov, V. V.; Dudanov, I. P.; Vinogradov, V. V. Journal of Materials Chemistry B 2016, 4, 5921–5928. (20) Cui, J.; Feng, Y.; Yue, S.; Zhao, Y.; Li, L.; Liu, R.; Lin, T. Journal of Chemical Technology and Biotechnology 2016, 91, 1905–1913. (21) Li, J.; Jin, X.; Liu, Y.; Li, F.; Zhang, L.; Zhu, X.; Lu, Y. Chemical Communications 2015, 51, 9628–9631. (22) Kahraman, M. V.; Bayramo˘glu, G.; Kayaman-Apohan, N.; G¨ ung¨or, A. Food Chemistry 2007, 104, 1385–1392.

312

(23) Govan, J.; Gun’ko, Y. K. Nanomaterials 2014, 4, 222–241.

313

(24) Xu, J.; Sun, J.; Wang, Y.; Sheng, J.; Wang, F.; Sun, M. Molecules 2014, 19, 11465–

314

11486.

315

(25) Zhou, L.; Yuan, J.; Wei, Y. Journal of Materials Chemistry 2011, 21, 2823–2840.

316

(26) Dyal, A.; Loos, K.; Noto, M.; Chang, S. W.; Spagnoli, C.; Shafi, K. V.; Ulman, A.;

317

Cowman, M.; Gross, R. A. Journal of the American Chemical Society 2003, 125, 1684–

318

1685. 14

ACS Paragon Plus Environment

Page 14 of 19

Page 15 of 19

Journal of Agricultural and Food Chemistry

319

(27) Wu, Y.; Wang, Y.; Luo, G.; Dai, Y. Bioresource technology 2009, 100, 3459–3464.

320

(28) Safarik, I.; Baldikova, E.; Prochazkova, J.; Safarikova, M.; Pospiskova, K. 2018,

321

(29) Hosseini, S. H.; Hosseini, S. A.; Zohreh, N.; Yaghoubi, M.; Pourjavadi, A. Journal of

322

323

324

325

326

327

328

329

330

331

332

333

334

335

336

337

338

339

340

341

agricultural and food chemistry 2018, (30) Zahra, Z.; Waseem, N.; Zahra, R.; Lee, H.; Badshah, M. A.; Mehmood, A.; Choi, H.-K.; Arshad, M. Journal of agricultural and food chemistry 2017, 65, 5598–5606. (31) Shapovalova, O. E.; Drozdov, A. S.; Brushkova, E. A.; Morozov, M. I.; Vinogradov, V. V. Arabian Journal of Chemistry 2018, (32) Cornell, R. M.; Schwertmann, U. The iron oxides: structure, properties, reactions, occurrences and uses; John Wiley & Sons, 2003. (33) Li, Y.-S.; Church, J. S.; Woodhead, A. L. Journal of Magnetism and Magnetic Materials 2012, 324, 1543–1550. (34) Drozdov, A. S.; Ivanovski, V.; Avnir, D.; Vinogradov, V. V. Journal of colloid and interface science 2016, 468, 307–312. (35) Larson, S. B.; Greenwood, A.; Cascio, D.; Day, J.; McPherson, A. Journal of molecular biology 1994, 235, 1560–1584. (36) Shabanova, E. M.; Drozdov, A. S.; Ivanovski, V.; Suvorova, I. I.; Vinogradov, V. V. RSC Advances 2016, 6, 84354–84362. (37) Drozdov, A. S.; Vinogradov, V. V.; Dudanov, I. P.; Vinogradov, V. V. Scientific reports 2016, 6 . (38) Anastasova, E. I.; Ivanovski, V.; Fakhardo, A. F.; Lepeshkin, A. I.; Omar, S.; Drozdov, A. S.; Vinogradov, V. V. Soft matter 2017, 13, 8651–8660. (39) Layer, P.; Zinsmeister, A. R.; DiMagno, E. P. Gastroenterology 1986, 91, 41–48. 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

342

343

(40) Drozdov, A. S.; Shapovalova, O. E.; Ivanovski, V.; Avnir, D.; Vinogradov, V. V. Chemistry of Materials 2016, 28, 2248–2253.

344

(41) Singh, V.; Singh, D. Process Biochemistry 2013, 48, 96–102.

345

(42) Sohrabi, N.; Rasouli, N.; Torkzadeh, M. Chemical Engineering Journal 2014, 240,

346

347

348

349

350

351

352

426–433. (43) Hasirci, N.; Aksoy, S.; Tumturk, H. Reactive and Functional Polymers 2006, 66, 1546– 1551. (44) Mukherjee, A. K.; Kumar, T. S.; Rai, S. K.; Roy, J. K. Biotechnology and Bioprocess Engineering 2010, 15, 984–992. (45) Ertan, F.; Yagar, H.; Balkan, B. Preparative biochemistry & biotechnology 2007, 37, 195–204.

16

ACS Paragon Plus Environment

Page 16 of 19

Page 17 of 19

Journal of Agricultural and Food Chemistry

Figure 3: Catalytic hydrolysis of starch solution by free amylase and amylase@ferria kinetic curves. Hydrolysis process kinetics at 60◦ C (a), 70◦ C (b), 80◦ C (c), and 90◦ C (d); relative activity of the amylase@ferria composite reused in ten successive experiments (e); long term stability of amylase@ferria and free amylase stored at 4◦ C in phosphate buffer solution (f). The results presented is represented by average values of 3 measurements for each of the curves.

17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 4: Enzymatically-doped ceramic coating amylase@ferria on the surface of the glass vessel. SEM cross section of the coating is presented (a); the dynamic of starch hydrolysis in the model reactor in successive cycles. Catalytic coating in model experiment have demonstrated a good stability against thermal inactivation and peeling, and showed high degree of reusability of the system with only 6% degradation after 10 cycles of starch hydrolysis.

18

ACS Paragon Plus Environment

Page 18 of 19

Page 19 of 19

Journal of Agricultural and Food Chemistry

199x268mm (300 x 300 DPI)

ACS Paragon Plus Environment