Gelatin Nanofibers Containing Zinc Oxide

Aug 23, 2018 - The ethylcellulose/gelatin solutions containing various concentrations of zinc oxide (ZnO) nanoparticles were electrospun, and the resu...
0 downloads 0 Views 3MB Size
Subscriber access provided by Kaohsiung Medical University

Food and Beverage Chemistry/Biochemistry

Hydrophobic Ethylcellulose/Gelatin Nanofibers Containing Zinc Oxide Nanoparticles for Antimicrobial Packaging Yuyu Liu, Yang Li, LingLi Deng, Lin Zou, Fengqin Feng, and Hui Zhang J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b03267 • Publication Date (Web): 23 Aug 2018 Downloaded from http://pubs.acs.org on August 25, 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 37

Journal of Agricultural and Food Chemistry

1

Hydrophobic Ethylcellulose/Gelatin Nanofibers Containing Zinc Oxide

2

Nanoparticles for Antimicrobial Packaging

3 4

Yuyu Liu, Yang Li, Lingli Deng, Lin Zou, Fengqin Feng, Hui Zhang *

5 6

College of Biosystems Engineering and Food Science, Fuli Institute of Food Science, Zhejiang Key

7

Laboratory for Agro-Food Processing, Zhejiang R&D Center for Food Technology and Equipment,

8

Zhejiang University, Hangzhou 310058, China

9 10

* E-mail: [email protected]. Phone: +86-571-88982981. Fax: +86-571-88982981.

11

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

12

Abstract

13

The ethylcellulose/gelatin solutions containing various concentrations of zinc oxide (ZnO)

14

nanoparticles were electrospun, and the resultant nanofibers were characterized by scanning

15

electron microscopy, energy dispersive X-ray, X-ray photoelectron spectrometer, X-ray diffraction,

16

Fourier transform infrared spectroscopy, mechanical testing, water contact angle, and water stability.

17

Results indicated that ZnO nanoparticles acting as fillers interacted with polymers, resulting in the

18

enhanced surface hydrophobicity and water stability of nanofibers. The antibacterial assay showed a

19

concentration-dependent effect of ZnO on the viabilities of Escherichia coli and Staphylococcus

20

aureus. Notably, the antimicrobial efficiency of the 1.5 wt% ZnO-containing fibers against S.

21

aureus was 43.7%, but increased to 62.5% after UV irradiation at 364 nm, possibly due to the

22

significantly increased amounts of intracellular reactive oxygen species. These results suggested

23

that the ZnO-containing nanofibers with excellent surface hydrophobicity, water stability and

24

antimicrobial activity exhibited potential uses in food packaging.

25

2

ACS Paragon Plus Environment

Page 2 of 37

Page 3 of 37

26

Journal of Agricultural and Food Chemistry

TOC Graphic

27 28

Keywords

29

electrospinning, ethylcellulose, gelatin, ZnO, antibacterial activity, UV irradiation

30

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

31

Introduction

32

Packaging has been widely used in food and beverages to maintain their quality and safety, and

33

to reduce food waste.1 Except for the basic role in acting as inert barrier to external conditions, the

34

ability to protect products from microbial contamination is preferred in food packaging, since

35

microbial spoilage is responsible for shortened shelf life and degraded food quality.2-4 Antimicrobial

36

packaging is reported as one of the most effective approaches to restrain multiplication of

37

microorganisms by incorporating antimicrobial substances in the packaging systems.3 Due to the

38

extensive antimicrobial activity, high specific surface area, and ability to improve mechanical

39

properties of matrix, metal oxide nanoparticles are promising antibacterial agents in food

40

packaging.5, 6 However, the potential toxicity of nanoparticles is an important obstacle in food

41

applications.7 Zinic oxide (ZnO) nanoparticles as GRAS approved food additive have attracted an

42

increasing interest in food packaging not only due to the excellent antimicrobial ability but also for

43

the desired mechanical performance and the minimum impact on human cells.5, 8 Notably, as a

44

photocatalytic material, UV illumination can induce the generation of reactive oxygen species (ROS)

45

on the surface of ZnO nanoparticles, leading to enhanced bioactivity due to the fact that the ROS

46

may penetrate into the cells and cause damage to bacteria.9 Li et al. detected the generation of ROS

47

on surface of ZnO nanoparticles under UV (365 nm) irradiation, and indicated that the higher the

48

total ROS concentration, the more significant the antibacterial activity.10

49

Encapsulation of metal oxide nanoparticles into polymers is a promising approach to enhance

50

the antimicrobial activity of nanoparticles due to the improved stability and sustained release.11

51

Electrospinning has been thought to be of great suitability in encapsulating active compounds in

52

food polymers and biopolymers, since it is a versatile and simple technique to fabricate functional 4

ACS Paragon Plus Environment

Page 4 of 37

Page 5 of 37

Journal of Agricultural and Food Chemistry

53

fibrous membranes with high porosity and surface-to-volume ratio, and improved stability and

54

mechanical strength.2 The enhanced dispersion of metal oxide nanoparticles was reported by

55

Mayorga

56

(3-hydroxybutyrate-co-3-hydroxyvalerate), and the resultant electrospun fiber mats showed the

57

enhanced antimicrobial activity against Salmonella enterica and Listeria monocytogenes.11

58

Augustine et al. observed that the addition of 5 wt% ZnO nanoparticles in the electrospun

59

polycaprolactone nanofibers enhanced the mechanical stability as well as the antibacterial

60

properties against Escherichia coli and Staphylococcus aureus.12

et

al.

through

incorporating

CuO

nanoparticles

into

poly

61

Additionally, the encapsulation matrices for food packaging should be regarded as safe. The

62

active loading based on gelatin nanofibers has been widely fabricated through electrospinning

63

owing to its desirable biodegradability and biocompatibility.13 Ultrafine gelatin/nisin fiber mats with

64

antimicrobial activity against S. aureus and L. monocytogenes were reported by Dheraprasart et al.13

65

The gelatin-based electrospun fibrous membranes displaying remarkable antibacterial properties

66

against E. coli and S. aureus were produced by Padrao et al. with fish gelatin as structural matrix

67

and bovine lactoferrin as the active antimicrobial agent.14 On the other hand, elements such as

68

excellent hydrophobicity to block the entry of water and moisture are necessary for packaging.7 Our

69

recent work suggested that due to the advantageous properties of safety, electrospinability and

70

hydrophobicity, blending ethylcellulose with gelatin is an ideal approach to overcome the natural

71

hydrophilicity of gelatin based fibers, rendering the improved water stability.15

72

In this study, before electrospinning the ethylcellulose/gelatin solutions containing various

73

levels of ZnO nanoparticles were evaluated by viscosity and zeta potential measurements. The

74

resultant fibers were characterized by scanning electron microscopy (SEM), energy dispersive 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

75

X-ray (EDX), X-ray photoelectron spectrometer (XPS), X-ray diffraction (XRD), Fourier transform

76

infrared spectroscopy (FTIR), mechanical testing, water contact angle (WCA), and water stability.

77

The antibacterial activities of the nanofibers were tested against E. coli and S. aureus based on disc

78

diffusion technique. Subsequently, the fiber mat containing 1.5 wt% ZnO nanoparticles was selected

79

for analyzing the impact of UV irradiation on the antimicrobial viability using culture turbidity

80

method.

81

MATERIALS AND METHODS

82

Materials. Ethylcellulose (3-7 cP, MW, 20-30 kDa), gelatin of type B (Bloom 250, MW, 100 kDa)

83

and zinc oxide (ZnO) nanoparticles with particle size of 30 nm were supplied by Aladdin, Inc.

84

(Shanghai, China). Absolute ethanol (Sinopharm), acetic acid (99.8%, Sinopharm), and deionized

85

water were used for solution preparations. All reagents were utilized as received.

86

Solution preparation. Solutions were prepared by dissolving 15 wt% (w/v) ethylcellulose and 15

87

wt% (w/v) gelatin into water/ethanol/acetic acid at a volume ratio of 2/2/6. Subsequently, ZnO

88

nanoparticles were added at concentrations of 0 wt%, 1 wt%, 1.5 wt%, and 2 wt% (w/v), which

89

were denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG, respectively. Thus, the theoretical mass

90

fraction of ZnO regarded to polymers was 0%, 3.2%, 4.8% and 6.3%, respectively. The solutions

91

were gently stirred for 1 h at 60 °C and then vigorously stirred for 2 h at 25 °C prior to

92

electrospinning.

93

Rheology and zeta potential. Rheology was characterized using an Anton Paar MCR302

94

rheometer (Anton Paar, Austria, QC) at 25 °C with a scanning frequency from 1 to 100 rad/s using 1%

95

strain. The complex viscosity at a shear rate of 1 rad/s was taken as the average solution viscosity.

96

Zeta potential was measured using a dynamic light scattering instrument (Malvern, Zetasizer Nano 6

ACS Paragon Plus Environment

Page 6 of 37

Page 7 of 37

Journal of Agricultural and Food Chemistry

97

ZS90, UK) at 25 °C. Triplicate measurements were performed.

98

Electrospinning. The fabrication of nanofibers by electrospinning was according to our previous

99

studies.15 4 mL solution was pumped into a 5 mL syringe with a 20-gauge needle. A high-voltage

100

power supply (Gamma High Voltage) was employed to generate a voltage up to 15 kV. A pump

101

(LSP02-1B, Baoding Longer Precision Pump Co., Ltd., China) provided a constant flow rate of 1.0

102

mL/h. The resultant fibers were collected on a grounded cylindrical aluminum foil at a distance of

103

100 mm away from the needle at 25 °C with a humidity of 50%. The thickness of the fibers was in

104

the range of 200-300 µm measured by a thickness meter (C640, Labthink, China).

105

Morphology and element mapping. The morphology observation was carried out using a field

106

emission scanning electron microscope (SEM, SU8010, Hitachi, Japan). Nano Measure software

107

was applied to measure the fiber diameter of 120 fibers at random from SEM images. To observe

108

the distribution of Zn element, the fibers were analyzed by energy dispersive X-ray (EDX, SU8010,

109

Hitachi, Japan) at a working depth of 5 µm.

110

X-ray photoelectron spectrometer. The surface atomic composition of fibers was examined by

111

X-ray photoelectron spectrometer (XPS, AXIS SUPRA, Kratos) using Al Kα X-ray source (1486.6

112

eV) with a standard working depth of 10 nm. Wide scan spectra were recorded from 1000 to 0 eV.

113

The scan steps of wide scan spectra and high-resolution spectra were 1.0 eV and 0.1 eV,

114

respectively. The atom percentage was calculated using Speclab software, and the curve fitting of

115

high-resolution C 1s spectra was analyzed by Origin pro9 software. The spectra calibration was

116

based on setting the C-C/C-H contribution in the C 1s emission at 284.8 eV.16

117

X-ray diffraction. The X-ray diffraction (XRD, Bruker D8 ADVACNCE, Germany) was

118

performed using Cu Kα (1.5406 Å) radiation at a scanning rate of 1 °/min ranging from 4° to 80°. 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

119

The average spacing of the peak was calculated as follow:17

d =

0.9λ (1) β cos

120

where λ, β, and θ represent X-ray wavelength (1.541), full width of half maximum (FWHM) and

121

the Bragg diffraction angle, respectively.

122

Fourier transform infrared spectrometer. The Fourier transform infrared spectrometer (FTIR,

123

AVA TAR 370) spectra were obtained by mixing nanofiber mats with KBr. The resolution of

124

spectra was 4 cm−1 with 16 scans, and the measurement range was from 4000 to 400 cm−1.

125

Mechanical testing. The tensile properties were measured at 25 °C using an electron testing

126

machine (Zwick GmbH & Co.KG, China). The fiber mats were cut into 5 mm × 30 mm stripes,

127

with thickness of 200 ± 50 µm measured by a thickness meter (C640, Labthink, China). The

128

stress-stain curves of the mats were recorded to determine the elastic modulus, tensile strength, and

129

elongation at break. Five tests were performed for each sample.

130

Water contact angle. The water contact angle (WCA) was measured at 25 °C using a surface

131

tension meter (DCAT-21, Delta Phase, Germany). Milli-Q water was dropped on the fiber surface,

132

and the images were captured by a camera system. Three random spots were measured for each

133

sample.

134

Water stability. The water stability measurements of the fibers were conducted by following the

135

procedure of our previous method.15 The disc-shaped fiber mats (D0=11 mm) were first determined

136

(W0) before immersed in water at 25 °C with a shaking speed at 150 rpm for 24 h. After immersion,

137

the fiber diameter (DT) was measured, followed by weighing the mats before (Ww) and after (WD)

138

drying at 60 °C for 1 h, respectively. The surface swelling (SW), water-uptake capacity (WT), and

139

weight loss (WL) were calculated by the equations: 8

ACS Paragon Plus Environment

Page 8 of 37

Page 9 of 37

Journal of Agricultural and Food Chemistry

D −D (2) D W − W W = (3) W W − W W = × 100% (4) W S =

140

Inhibitory zone evaluation. The antibacterial activities of the fibers with different ZnO levels

141

against Escherichia coli (ATCC 25922) and Staphylococcus aureus (CMCC 26003) were measured

142

using a disc diffusion method.18 The cells (~105 CFU/mL) of each pathogen were prepared by the

143

serial dilution method. 0.1 mL diluent was spread on agar media using a triangular metal rod. The

144

disc-shaped fibers with a diameter of 11 mm (2-3 mg) were sterilized by UV for 1 h before being

145

adhered to the medium. After incubation for 24 h at 37 °C, the inhibitory zone was measured to

146

estimate the sensitivity of bacteria to fibers. Three replications were performed for each sample.

147

Bactericidal kinetics after UV irradiation. The Z1.5-EG fibrous mat was selected to analyze the

148

antibacterial activity after UV irradiation (364 nm, 250 W), using a culture turbidity method.10, 18 11

149

mg of the sterilized mat was added into tubes containing 20 mL culture medium. One set of tubes

150

was exposed to UV for 1 h, and the other set was protected from light for 1 h. After treatment, 0.2

151

mL of each bacterial suspension (106 CFU/mL) was added, and then incubated in the shaker

152

platform (150 rpm, 37 °C). The growth of bacteria was monitored at an interval of 2 h for 24 h by a

153

microplate reader (MDS ANALYTICAL, America). Medium with only inoculum or the fibrous mat

154

without ZnO (Z0-EG) was used as negative control or positive control, respectively. All

155

experiments were performed in triplicate. The inhibition efficiency (IE) was calculated as follows:

IE =

OD#" − OD$" OD#"

× 100% (5)

156

where OD#" and OD$" were the OD600 values of culture medium without and with the fiber mat

157

at time t, respectively. 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

158

In order to estimate the generation of intracellular reactive oxygen species (ROS) caused by

159

ZnO nanoparticles under UV irradiation, the above E. coli suspensions after 24 h of exposure to

160

fibers were used to quantify the ROS level within the cells by the method provided by Najim et al.19

161

After incubation, the cells were collected by centrifugation and suspended with PBS. The resultant

162

solutions were mixed with 2,7-dichlorofluorescein diacetate (DCFH-DA, sigma, USA) at a volume

163

ratio of 2000:1 for 45 min at 37 °C under dark conditions. The fluorescence intensity of DCFH was

164

measured by fluorescence spectrophotometer (Molecular Devices, China) at an excitation and

165

emission wavelength of 488 nm and 535 nm, respectively.

166

Statistical analysis. The experimental results were presented as mean plus or minus standard

167

deviation. The statistical evaluation was carried out using Origin pro9 software via one-way

168

ANOVA with analysis of variance at the 0.05 significance level.

169

Results and Discussion

170

Solution rheology and zeta potential. As listed in Table 1, the apparent viscosities of the

171

ethylcellulose/gelatin solutions without (Z0-EG) and with 1 wt% (Z1-EG), 1.5 wt% (Z1.5-EG), and

172

2 wt% (Z2-EG) ZnO nanoparticles were 2.05, 1.77, 1.30 and 1.78 Pa s, respectively. Obviously, the

173

solutions became less viscous after the addition of ZnO nanoparticles, while the viscosity of the

174

Z2-EG solution gave a significant (p < 0.05) rise compared with that of the Z1.5-EG solution. A

175

viscosity decrease after adding AgNO3 into poly (vinyl pyrrolidone)/cellulose nanocrystals systems

176

was observed by Huang et al., who owed the changes to the competition between polymers and Ag+,

177

leading to destruction of some intramolecular or intermolecular hydrogen bonds in polymers.20 An

178

interpenetrating network among ethylcellulose and gelatin through interactions such as hydrogen

179

bonds has been proposed previously.15 Therefore, it could be assumed that the ZnO nanoparticles 10

ACS Paragon Plus Environment

Page 10 of 37

Page 11 of 37

Journal of Agricultural and Food Chemistry

180

probably acting as fillers were dispersed in the network by blocking interactions among polymers,

181

resulting in the decreased viscosity. However, the increased viscosity of the Z2-EG solution could

182

be an indicator of aggregation of ZnO nanoparticles at a higher concentration. Augustine et al.

183

suggested that with the concentration of ZnO nanoparticles up to 2 wt% in polycaprolactone

184

solution, particle agglomeration encountered, leading to high solution viscosity and apparently

185

increased fiber diameter.12

186

The solutions of 0%, 1%, 1.5%, and 2% ZnO nanoparticles without polymers displayed zeta

187

potential values of 3.89, 6.98, 5.73, and 3.94 mV, respectively. The lower zeta potential value of 2%

188

ZnO solution might be due to the formation of ZnO aggregates. Meanwhile, the zeta potential

189

values of the Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG solutions were 1.47, 1.08, 0.87, and 1.00 mV,

190

respectively. Matusiak et al. reported that in the case of polymer and oxide systems, adsorption of

191

the polymer on oxide surface could change the zeta potential of the bulk.21 Yu et al. ascribed the

192

zeta potential changes of ZnO-containing matrix to the interaction effect between nanoparticles and

193

polymers.22 Thus, the observed decrease in zeta potential values of ZnO nanoparticles after

194

introduced into the ethylcellulose/gelatin solutions could be a result of interactions between ZnO

195

and the absorbed polymers. Additionally, the zeta potential of Z2-EG solution was slightly higher

196

than that of the Z1.5-EG solution, probably resulting from the reduced interactions caused by

197

aggregates of ZnO nanoparticles at a higher concentration.

198

Nanofiber morphology and element mapping. SEM images in Figure 1 (A-D) showed that the

199

fibers without ZnO nanoparticles (Z0-EG) were smooth with an average diameter of 500 nm. The

200

diameters of fibers with ZnO nanoparticles (Z1-EG, Z1.5-EG and Z2-EG) were 588, 640 and 804

201

nm, respectively, without any visually observed morphology changes, indicating the homogenous 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

202

distribution of ZnO nanoparticles within the fibers.23 It is universally accepted that the high zeta

203

potential value is responsible for strong electrostatic force.20 During electrospinning, the

204

electrostatic force acting as a strong drag force fights on the jet, leading to strong elongation of jet

205

and generation of uniform and thin fibers.15 Thus, the reduced zeta potential values of solutions

206

after the ZnO introduction played a major role in the formation of thicker fibers. Figure 1 (E-J)

207

presented the EDX analysis and Zn element mapping of the Z0-EG, Z1-EG and Z2-EG fibers,

208

respectively. Obviously, all the fibers showed the presence of C, N, O, while the signals of Zn

209

element were detected in the ZnO-containing fibers. Figure 1 (H) exhibited a highly Zn dispersion

210

in the Z1-EG fibers, indicating good system compatibility.24 On the other hand, the increased Zn

211

distribution density and slight aggregates were observed for the Z2-EG fibers in Figure 1 (J),

212

validating the interpretation of viscosity changes proposed in solution rheology characterization.

213

X-ray photoelectron spectrometer. To evaluate the chemical composition of the fiber surface, the

214

XPS analysis was conducted and the results were displayed in Figure 2 and Table 2. The wide scan

215

spectra showed the presence of C, N and O atoms in each fiber mat, and the typical peak of Zn was

216

observed except the Z0-EG fibers. The high resolution scan of Zn element exhibited two peaks at

217

1020 and 1043 eV assigned to Zn (2p3/2) and Zn (2p3/2),25 confirming the existence of Zn element in

218

the Z1-EG, Z1.5-EG and Z2-EG fibers, respectively. The high resolution XPS spectra for C 1s

219

revealed that two peaks at 284.8 eV and 286.4 eV corresponding to C-C and C-O existed in all the

220

fiber samples, but another small peak at 288.1 eV related to C=O was only observed for the Z2-EG

221

fibers.26

222

As summarized in Table 2, the N atom percentage values for the Z0-EG, Z1-EG, Z1.5-EG and

223

Z2-EG fibers were 7.3%, 6.9%, 5.8% and 5.2%, respectively, indicating the decreased N content on 12

ACS Paragon Plus Environment

Page 12 of 37

Page 13 of 37

Journal of Agricultural and Food Chemistry

224

the fiber surface with the increasing ZnO concentration. The calculated ratio of surface Zn regarded

225

to theoretical content was 47.6% for the Z1-EG fibers, 46.9% for the Z1.5-EG fibers but 71.4% for

226

the Z2-EG fibers, implying a higher ZnO content on the Z2-EG fiber surface. The decreased N but

227

increased Zn content on the fiber surface suggested the replacement of hydrophilic groups by

228

hydrophobic ZnO nanoparticles, contributing to the improved surface hydrophobicity.27

229

X-ray diffraction. The XRD spectra of the fabricated nanofibers as well as the ZnO nanoparticles

230

were shown in Figure 3 and Table 3. It was evident that there was an extra peak at 31.82° in the

231

case of the fibers containing 2 wt% ZnO nanoparticles (Z2-EG), consistent with the typical

232

crystalline peak of ZnO nanoparticles assigned to the (100) plane.28 This implied the presence of

233

ZnO aggregates in the Z2-EG fibers, which was in agreement with our EDX analysis. Generally,

234

ethylcellulose possesses two crystalline reflections at 2θ = 7.8° and 2θ = 20.3°, and gelatin has a

235

broad diffraction peak in the 2θ range around 20°.29, 30 Figure 3 indicated that the second reflection

236

at around 20° became less prominent after the addition of ZnO nanoparticles, and the calculated d

237

spacing value of the fibers without ZnO nanoparticles (Z0-EG) in Table 3 was increased from 0.222

238

to 0.289, 0.292 and 0.336 after the addition of 1 wt% (Z1-EG), 1.5 wt% (Z1.5-EG), and 2 wt%

239

(Z2-EG) ZnO nanoparticles, respectively. Li et al. demonstrated that the decreased peak intensity

240

and increased d value implied a less compact structure of the composite system.31 Thus, these

241

results suggested that the higher amounts of ZnO nanoparticles, the less compact structure of the

242

polymer network, which might be due to some destroyed intramolecular or intermolecular

243

interactions of polymers caused by the ZnO addition.

244

Fourier transform infrared spectrometer. It can be seen from Figure 4 that some characteristic

245

absorption peaks of the fibers without ZnO nanoparticles (Z0-EG) were observed at 3385 cm−1, 13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

246

1651 cm−1, 1538 cm−1, and 1446 cm−1, related to O-H bending, C=O stretching (amide I), N-H/C-N

247

stretching (amide II), respectively.29 In contrast, the ZnO-containing fibers (Z1-EG, Z1.5-EG and

248

Z2-EG) showed a new peak at 667 cm−1 assigned to the metal-oxygen (M-O) stretching mode,28

249

indicating the existence of ZnO nanoparticles. Additionally, the fibers containing ZnO nanoparticles

250

exhibited the decreased intensities of the amide II bands, especially an almost nonexistent peak at

251

1446 cm−1. Generally, the higher the intensity of amide band, the more number of the triple helix

252

structures gelatin chains have.32 Therefore, the observed changes indicated the loss of triple helix

253

state linked by hydrogen bonds, which evidenced the destruction of gelatin molecular interactions

254

after introduction of ZnO nanoparticles.

255

Mechanical testing. The tensile properties of the fiber mats including average elastic modulus (E),

256

tensile strength (σy) and elongation at break (εb) were listed in Table 4. The fiber mats without ZnO

257

(Z0-EG) possessed E, σy and εb of 148.10 MPa, 1.70 MPa and 1.30%, respectively. After the ZnO

258

addition ranging from 1 - 1.5 wt%, these values increased with the highest E (203.24 MPa), σy (2.42

259

MPa) and εb (1.57%) observed for the Z1.5-EG fibers. It was reported that dispersion of fillers and

260

filler-matrix interactions were the dominant factors in determining mechanical performance of

261

polymer composites. The enhancement in tensile properties was indicators of good nanoparticle

262

dispersion and strong interfacial interactions.4 However, the fibers containing 2 wt% ZnO presented

263

a slight decrease in E (134.28 MPa), σy (1.65 MPa) and εb (0.98%) compared with the Z0-EG fibers.

264

In the case of ZnO loaded cellulose films reported by Zhao et al., the 7.4 wt% ZnO addition led to

265

decreased εb values compared with pure cellulose films, which was attributed to the poor dispersion

266

of ZnO in cellulose matrix caused by formation of agglomerates at higher concentrations.8 Thus,

267

our results indicated that the addition of ZnO at a concentration below 1.5 wt% might improve the 14

ACS Paragon Plus Environment

Page 14 of 37

Page 15 of 37

Journal of Agricultural and Food Chemistry

268

mechanical properties of fibers, while a tendency of formation of agglomerates at a higher

269

concentration (2 wt%) could lead to a poor mechanical performance.

270

Water contact angle. The results of water contact angle (WCA) recorded at 1 s, 5 s and 15 s were

271

displayed in Figure 5. The WCA of the fibers without ZnO nanoparticles (Z0-EG) was 119.3° at 1 s,

272

lower than the values of the Z1-EG (129.7°), Z1.5-EG (134.0°) and Z2-EG (128.5°) fibers. Anitha

273

et al. observed that the wetting property of the cellulose acetate fibers changed from hydrophilic

274

(47°) to hydrophobic (124°) after impregnating ZnO.33 Hydrophobicity is largely dependent on the

275

surface morphology and chemical structure.34 Considering the reduced N atom percentage on

276

nanofiber surface by introduction of ZnO nanoparticles from the XPS analysis, the improved

277

surface hydrophobicity could be a result of the decreased number of gelatin chains distributed on

278

surface, since N element was the characteristic element of gelatin in the fiber matrix. Additionally,

279

the presence of the hydrophilic group (C=O) of the Z2-EG fibers observed in the XPS spectrum for

280

C 1s might give an explanation to the decreased WCA value, compared to the other ZnO-containing

281

nanofibers. The Z0-EG, Z1-EG, Z1.5-EG and Z2-EG fibers gave the WCA values of 109.5°, 117.5°,

282

130.0° and 122.1° at 15 s showing a 9.8°, 12.2°, 4.0° and 5.4° reduction compared to the initial

283

values, suggesting the higher water stability at higher ZnO levels. Thus, the introduction of ZnO

284

nanoparticles into fibers effectively strengthened the surface hydrophobic property, which was

285

suitable for packaging materials to inhibit water penetration.7

286

Water stability. As presented in Table 5, the fibers without ZnO nanoparticles (Z0-EG) gave the

287

highest water-uptake capacity (WT) value of 536.41%, compared with the Z1-EG (297.9%),

288

Z1.5-EG (209.31%) and Z2-EG (203.61%) fibers, respectively. The lower gelatin content on the

289

fiber surface supported the remarkably decreased WT values of the ZnO-containing fibers due to the 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

290

fact that the lower the gelatin content, the less the hydrophilic groups to absorb water molecules.35 It

291

was obvious that the surface swelling (SW) and weight loss (WL) displayed a similar trend with the

292

WT value, the higher the ZnO concentration the lower the values. Notably, the SW and WL values of

293

the Z0-EG and Z1-EG fibers were much higher than those of the Z1.5-EG and Z2-EG fibers. By

294

comparing the surface dimensional stability of pure chitosan or gelatin fibers with its binary

295

scaffolds containing hydrophobic polycaprolactone, Gomes et al. reported much lower SW values

296

for binary scaffolds due to the increased surface hydrophobicity caused by polycaprolactone.36 Our

297

results indicated that the enhanced hydrophobicity of the ZnO-containing fibers provided by WCA

298

analysis contributed to the improved water stability. These parameters were related to dislocations

299

or blistering and sensitivity to water, which was necessary for using as food package and coating

300

films.35, 37

301

Inhibitory zone evaluation. Table 6 summarized the inhibitory diameters of the nanofibers against

302

E. coli and S. aureus. The fibers without ZnO (Z0-EG) as control showed absence of antimicrobial

303

activity, while the fibers containing 1 wt%, 1.5 wt% and 2 wt% ZnO nanoparticles showed the

304

inhibitory diameters of 0.69, 1.30, 1.61 mm/mg against E. coli and 0.75, 1.17, 1.33 mm/mg against

305

S. aureus, respectively. These results indicated that the bacteria inhibition displayed a dose-effect

306

manner, and E. coli was observed to be more susceptible than S. aureus. Xu et al. revealed that the

307

addition of 2 wt% AgNO3 endowed the gelatin fibers with an inhibition zone of 1.6 mm against S.

308

aureus and 2.5 mm against Pseudomonas aeruginosa, and attributed the antimicrobial activity

309

difference to the thinner cell wall of P. aeruginosa.18 The similar results were also reported by

310

Kanmani et al., who suggested that the thicker and complex peptidoglycan structure of Gram

311

positive bacterial obstructed the nanoparticles penetration, while a negatively charged and thinner 16

ACS Paragon Plus Environment

Page 16 of 37

Page 17 of 37

Journal of Agricultural and Food Chemistry

312

peptidoglycan layer facilitated the nanoparticles transport.38

313

Bactericidal kinetics after UV irradiation. The growth curves of E. coli and S. aureus in Figure 6

314

(A) consisted of three phases: lag phase, exponential phase, and stabilization phase. Under exposure

315

to the different fibers, no reduction in cell viability was observed between the positive (Z0-EG

316

fibers) and negative control, while substantial decrease at exponential phase was obvious for the

317

fibers containing 1.5 wt% ZnO nanoparticles (Z1.5-EG). Notably, the Z1.5-EG fibers after UV

318

irradiation for 1 h (denoted as UV Z1.5-EG) was found to exhibit a higher antimicrobial activity

319

than the light protection for 1 h (denoted as Dark Z1.5-EG), as indicated by the lower OD600 value

320

at each time interval. From Figure 6 (B), the inhibition efficiency of the Z1.5-EG fibers against

321

either E. coli or S. aureus firstly increased, and then decreased significantly as the incubation time

322

prolonged. Finally, no significant difference of the inhibition efficiency between the Z1.5-EG fibers

323

and negative control was observed at stationary phase. The highest inhibition efficiency values of

324

the UV Z1.5-EG and Dark Z1.5-EG were 98.4% and 98.0% against E. coli, and 62.5% and 43.7%

325

against S. aureus, respectively, suggesting a more efficient antibacterial activity after UV irradiation.

326

This suggested that UV irradiation was an effective way to enhance the bacteriostatic performance

327

of the prepared ZnO-containing nanofibers.

328

It was proposed that the increased reactive oxygen species (ROS) production from ZnO under

329

UV irradiation resulted in the enhanced antibacterial activity due to the high reactivity and

330

oxidizing property of ROS.9 Thus, the E. coli suspensions after exposure to fibers for 24 h were

331

selected to detect the ROS concentration, and the results expressed as percentages of the value

332

against negative control were presented in Figure 6 (C). A significant difference of fluorescence

333

intensity was observed after exposure with the highest value (166%) for the UV Z1.5-EG fibers 17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

334

while the lowest value (130%) for the Z0-EG fibers, indicating the higher intracellular ROS levels

335

of the ZnO-containing fibers after exposure to UV light. The similar results were also reported by Li

336

et al., who demonstrated that under UV irradiation (365 nm), ZnO nanoparticles generated three

337

types of ROS, and the more the total ROS concentration, the higher the antibacterial activity against

338

E. coli.10

339

In summary, the present work demonstrated that ZnO nanoparticles were successfully

340

encapsulated in the ethylcellulose/gelatin nanofibers through electrospinning. Generally, ZnO

341

nanoparticles were homogeneously dispersed and did not change the fiber morphology, due to the

342

fact that ZnO nanoparticles acting as fillers destroyed intramolecular or intermolecular interactions

343

among polymers and instead formed new interactions with polymers. It was confirmed that the

344

ZnO-containing fibers exhibited the enhanced surface hydrophobicity and water stability owing to

345

the decreased gelatin distribution on fiber surface caused by the ZnO addition. The introduction of

346

ZnO nanoparticles endowed the ethylcellulose/gelatin nanofibers with the improved mechanical

347

properties and concentration-dependent antibacterial properties against E. coli and S. aureus.

348

Additionally, UV irradiation enabled the increased antibacterial efficiency of the fibers containing

349

1.5% ZnO nanoparticles from 43.7% to 62.5% against S. aureus. These results suggested that the

350

ZnO-containing nanofibers with excellent hydrophobicity, water stability and antibacterial

351

performance may have promising applications as food packaging materials.

352

Acknowledgement

353

This work was supported by the National Natural Science Foundation of China (Grant No.

354

31772013).

355

Conflict of interest 18

ACS Paragon Plus Environment

Page 18 of 37

Page 19 of 37

Journal of Agricultural and Food Chemistry

The authors declare no competing financial interest.

356 357

References

358

(1)

359

with a Consideration of Opportunities for Developing Countries. Trends Food Sci. Technol. 2011,

360

22, 604-610.

361

(2)

362

Potential Food Based Applications. Trends Food Sci. Technol. 2014, 38, 21-33.

363

(3)

364

Antimicrobial Films in Food Applications. Food Chem. 2009, 114, 1173-1182.

365

(4)

366

with Improved Mechanical, Barrier and Antibacterial Properties. Int. J. Mol. Sci. 2014, 15,

367

10950-10973.

368

(5)

369

Poly(Ethylene Oxide) and Zno Nanoparticles on the Morphology, Tensile and Thermal Properties of

370

Cellulose Acetate Nanocomposite Fibrous Film. Polym. J. 2011, 43, 978-986.

371

(6)

372

Innovative Packaging Material for Meat and Meat-Products. J. Food Sci. Technol.-Mysore 2015, 52,

373

4600-4606.

374

(7)

375

Characterization of Phb/Pbat-Based Biodegradable Antibacterial Hydrophobic Nanofibrous

376

Membranes. Polym. Adv. Technol. 2018, 29, 481-489.

377

(8)

Bradley, E. L.; Castle, L.; Chaudhry, Q., Applications of Nanomaterials in Food Packaging

Bhushani, J. A.; Anandharamakrishnan, C., Electrospinning and Electrospraying Techniques:

Dutta, P. K.; Tripathi, S.; Mehrotra, G. K.; Dutta, J., Perspectives for Chitosan Based

Diez-Pascual, A. M.; Diez-Vicente, A. L., Poly(3-Hydroxybutyrate)/Zno Bionanocomposites

Pittarate, C.; Yoovidhya, T.; Srichumpuang, W.; Intasanta, N.; Wongsasulak, S., Effects of

Amna, T.; Yang, J.; Ryu, K. S.; Hwang, I. H., Electrospun Antimicrobial Hybrid Mats:

Lin, X. H.; Fan, X. Y.; Li, R.; Li, Z. G.; Ren, T.; Ren, X. H.; Huang, T. S., Preparation and

Zhao, S. W.; Guo, C. R.; Hu, Y. Z.; Guo, Y. R.; Pan, Q. J., The Preparation and Antibacterial 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 37

378

Activity of Cellulose/Zno Composite: A Review. Open Chem. 2018, 16, 12.

379

(9) Sirelkhatim, A.; Mahmud, S.; Seeni, A.; Kaus, N. H. M.; Ann, L. C.; Bakhori, S. K. M.; Hasan,

380

H.; Mohamad, D., Review on Zinc Oxide Nanoparticles: Antibacterial Activity and Toxicity

381

Mechanism. Nano-Micro Lett. 2015, 7, 219-242.

382

(10) Li, Y.; Zhang, W.; Niu, J. F.; Chen, Y. S., Mechanism of Photogenerated Reactive Oxygen

383

Species and Correlation with the Antibacterial Properties of Engineered Metal-Oxide Nanoparticles.

384

ACS Nano 2012, 6, 5164-5173.

385

(11) Mayorga, J. L. C.; Rovira, M. J. F.; Mas, L. C.; Moragas, G. S.; Cabello, J. M. L.,

386

Antimicrobial

387

Poly(3-Hydroxybutyrate-Co-3-Hydroxyvalerate) and Copper Oxide Nanoparticles for Active

388

Packaging and Coating Applications. J. Appl. Polym Sci. 2018, 135, 11.

389

(12) Augustine, R.; Malik, H. N.; Singhal, D. K.; Mukherjee, A.; Malakar, D.; Kalarikkal, N.;

390

Thomas, S., Electrospun Polycaprolactone/Zno Nanocomposite Membranes as Biomaterials with

391

Antibacterial and Cell Adhesion Properties. J. Polym. Res. 2014, 21, 17.

392

(13) Dheraprasart, C.; Rengpipat, S.; Supaphol, P.; Tattiyakul, J., Morphology, Release

393

Characteristics, and Antimicrobial Effect of Nisin-Loaded Electrospun Gelatin Fiber Mat. J. Food

394

Prot. 2009, 72, 2293-2300.

395

(14) Padrao, J.; Machado, R.; Casal, M.; Lanceros-Mendez, S.; Rodrigues, L. R.; Dourado, F.;

396

Sencadas, V., Antibacterial Performance of Bovine Lactoferrin-Fish Gelatine Electrospun

397

Membranes. Int. J. Biol. Macromol. 2015, 81, 608-614.

398

(15) Liu, Y. Y.; Deng, L. L.; Zhang, C.; Feng, F. Q.; Zhang, H., Tunable Physical Properties of

399

Ethylcellulose/Gelatin Composite Nanofibers by Electrospinning. J. Agric. Food Chem. 2018, 66,

Nanocomposites

and

Electrospun

20

ACS Paragon Plus Environment

Coatings

Based

on

Page 21 of 37

Journal of Agricultural and Food Chemistry

400

1907-1915.

401

(16) Li, M. C.; Mei, C. T.; Xu, X. W.; Lee, S.; Wu, Q. L., Cationic Surface Modification of

402

Cellulose Nanocrystals: Toward Tailoring Dispersion and Interface in Carboxymethyl Cellulose

403

Films. Polymer. 2016, 107, 200-210.

404

(17) Agarwal, D. C.; Chauhan, R. S.; Kumar, A.; Kabiraj, D.; Singh, F.; Khan, S. A.; Avasthi, D. K.;

405

Pivin, J. C.; Kumar, M.; Ghatak, J.; Satyam, P. V., Synthesis and Characterization of Zno Thin Film

406

Grown by Electron Beam Evaporation. J. Appl. Phys. 2006, 99, 6.

407

(18) Xu, X. L.; Zhou, M. H., Antimicrobial Gelatin Nanofibers Containing Silver Nanoparticles.

408

Fiber. Polym. 2008, 9, 685-690.

409

(19) Najim, N. A., Comparative Study of Different Methods to Determine the Role of Reactive

410

Oxygen Species Induced by Zinc Oxide Nanoparticles. Aro 2015, 3, 30-34.

411

(20) Huang, S. W.; Zhou, L.; Li, M. C.; Wu, Q. L.; Kojima, Y.; Zhou, D. G., Preparation and

412

Properties of Electrospun Poly (Vinyl Pyrrolidone)/Cellulose Nanocrystal/Silver Nanoparticle

413

Composite Fibers. Mater. 2016, 9, 14.

414

(21) Matusiak, J.; Grzadka, E.; Bastrzyk, A., Stability, Adsorption and Electrokinetic Properties of

415

the Chitosan/Silica System. Colloid Surf. A-Physicochem. Eng. Asp. 2018, 554, 245-252.

416

(22) Yu, J.; Kim, H. J.; Go, M. R.; Bae, S. H.; Choi, S. J., Zno Interactions with Biomatrices: Effect

417

of Particle Size on Zno-Protein Corona. Nanomaterials 2017, 7, 14.

418

(23) Esfahani, H.; Prabhakaran, M. P.; Salahi, E.; Tayebifard, A.; Keyanpour-Rad, M.; Rahimipour,

419

M. R.; Ramakrishna, S., Protein Adsorption on Electrospun Zinc Doped Hydroxyapatite Containing

420

Nylon 6 Membrane: Kinetics and Isotherm. J. Colloid Interface Sci. 2015, 443, 143-152.

421

(24) Lu, H. Y.; Wang, Q. Q.; Li, G. H.; Qiu, Y. Y.; Wei, Q. F., Electrospun Water-Stable Zein/Ethyl 21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

422

Cellulose Composite Nanofiber and Its Drug Release Properties. Mater. Sci. Eng. C. 2017, 74,

423

86-93.

424

(25) Kim, J.; Mousa, H. M.; Park, C. H.; Kim, C. S., Enhanced Corrosion Resistance and

425

Biocompatibility of Az31 Mg Alloy Using Pcl/Zno Nps Via Electrospinning. Appl. Surf. Sci. 2017,

426

396, 249-258.

427

(26) Zhou, Z.; Liu, Y. G.; Liu, S. B.; Liu, H. Y.; Zeng, G. M.; Tan, X. F.; Yang, C. P.; Ding, Y.; Yan,

428

Z. L.; Cai, X. X., Sorption Performance and Mechanisms of Arsenic(V) Removal by Magnetic

429

Gelatin-Modified Biochar. Chem. Eng. J. 2017, 314, 223-231.

430

(27) Ding, B.; Ogawa, T.; Kim, J.; Fujimoto, K.; Shiratori, S., Fabrication of a Super-Hydrophobic

431

Nanofibrous Zinc Oxide Film Surface by Electrospinning. Thin Solid Films 2008, 516, 2495-2501.

432

(28) Ghaedi, M.; Ansari, A.; Habibi, M. H.; Asghari, A. R., Removal of Malachite Green from

433

Aqueous Solution by Zinc Oxide Nanoparticle Loaded on Activated Carbon: Kinetics and Isotherm

434

Study. J. Ind. Eng. Chem. 2014, 20, 17-28.

435

(29) Chitprasert, P.; Sutaphanit, P., Holy Basil (Ocimum Sanctum Linn.) Essential Oil Delivery to

436

Swine Gastrointestinal Tract Using Gelatin Microcapsules Coated with Aluminum Carboxymethyl

437

Cellulose and Beeswax. J. Agric. Food Chem. 2014, 62, 12641-12648.

438

(30) Lomakin, S. M.; Rogovina, S. Z.; Grachev, A. V.; Prut, E. V.; Alexanyan, C. V., Thermal

439

Degradation of Biodegradable Blends of Polyethylene with Cellulose and Ethylcellulose.

440

Thermochim. Acta 2011, 521, 66-73.

441

(31) Li, X. G.; Kresse, I.; Springer, J.; Nissen, J.; Yang, Y. L., Morphology and Gas Permselectivity

442

of Blend Membranes of Polyvinylpyridine with Ethylcellulose. Polymer. 2001, 42, 6859-6869.

443

(32) Muyonga, J. H.; Cole, C. G. B.; Duodu, K. G., Fourier Transform Infrared (Ftir) 22

ACS Paragon Plus Environment

Page 22 of 37

Page 23 of 37

Journal of Agricultural and Food Chemistry

444

Spectroscopic Study of Acid Soluble Collagen and Gelatin from Skins and Bones of Young and

445

Adult Nile Perch (Lates Niloticus). Food Chem. 2004, 86, 325-332.

446

(33) Anitha, S.; Brabu, B.; Thiruvadigal, D. J.; Gopalakrishnan, C.; Natarajan, T. S., Optical,

447

Bactericidal and Water Repellent Properties of Electrospun Nano-Composite Membranes of

448

Cellulose Acetate and Zno. Carbohydr. Polym. 2012, 87, 1065-1072.

449

(34) Natu, M. V.; de Sousa, H. C.; Gil, M. H., Effects of Drug Solubility, State and Loading on

450

Controlled Release in Bicomponent Electrospun Fibers. Int. J. Pharm. 2010, 397, 50-58.

451

(35) Soradech, S.; Nunthanid, J.; Limmatvapirat, S.; Luangtanaanan, M., An Approach for the

452

Enhancement of the Mechanical Properties and Film Coating Efficiency of Shellac by the

453

Formation of Composite Films Based on Shellac and Gelatin. J. Food Eng. 2012, 108, 94-102.

454

(36) Gomes, S.; Rodrigues, G.; Martins, G.; Henriques, C.; Silva, J. C., Evaluation of Nanofibrous

455

Scaffolds Obtained from Blends of Chitosan, Gelatin and Polycaprolactone for Skin Tissue

456

Engineering. Int. J. Biol. Macromol. 2017, 102, 1174-1185.

457

(37) Vatankhah, E.; Prabhakaran, M. P.; Jin, G.; Mobarakeh, L. G.; Ramakrishna, S., Development

458

of Nanofibrous Cellulose Acetate/Gelatin Skin Substitutes for Variety Wound Treatment

459

Applications. J. Biomater. Appl. 2014, 28, 909-921.

460

(38) Kanmani, P.; Rhim, J. W., Physicochemical Properties of Gelatin/Silver Nanoparticle

461

Antimicrobial Composite Films. Food Chem. 2014, 148, 162-169.

462 463

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

464

Figure captions

465

Figure 1 SEM images (A-D) of the ethylcellulose/gelatin fibers with ZnO levels of 0 wt%, 1 wt%,

466

1.5 wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG, respectively), and EDX

467

analysis (E, G, I) and Zn mapping (F, H, J) of the Z0-EG, Z1-EG and Z2-EG fibers.

468 469

Figure 2 XPS spectra of wide scan (A), high resolution of Zn scan (B) and C1s scan (C-F) of the

470

ethylcellulose/gelatin fibers with ZnO levels of 0 wt%, 1 wt%, 1.5 wt%, and 2 wt% (denoted as

471

Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG, respectively).

472 473

Figure 3 XRD patterns of ZnO nanoparticles and the ethylcellulose/gelatin fibers with ZnO levels

474

of 0 wt%, 1 wt%, 1.5 wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG,

475

respectively).

476 477

Figure 4 FTIR spectra of the ethylcellulose/gelatin fibers with ZnO levels of 0 wt%, 1 wt%, 1.5

478

wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG, respectively).

479 480

Figure 5 Water contact angles recorded at 1 s, 5 s and 15 s of the ethylcellulose/gelatin fibers with

481

ZnO levels of 0 wt%, 1 wt%, 1.5 wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG,

482

respectively).

483 484

Figure 6 Growth curves (A), inhibition efficiency (B) of E. coli and S. aureus, and DCFH

485

fluorescence intensities in E. coli (C) exposed to the ethylcellulose/gelatin fibers containing 1.5 wt% 24

ACS Paragon Plus Environment

Page 24 of 37

Page 25 of 37

Journal of Agricultural and Food Chemistry

486

ZnO with and without UV light (denoted as UV Z1.5-EG and dark Z1.5-EG, respectively).

487

Lowercases indicated statistical significance (p < 0.05).

488

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

489

Figure 1

26

ACS Paragon Plus Environment

Page 26 of 37

Page 27 of 37

Journal of Agricultural and Food Chemistry

Figure 2

490

27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 3

491

28

ACS Paragon Plus Environment

Page 28 of 37

Page 29 of 37

Journal of Agricultural and Food Chemistry

Figure 4

492

29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 5

493

30

ACS Paragon Plus Environment

Page 30 of 37

Page 31 of 37

Journal of Agricultural and Food Chemistry

Figure 6 (A)

(B)

494 495

(C)

496 497 498

31

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 32 of 37

499

Tables

500

Table 1. The apparent viscosities and zeta potential values of the ethylcellulose/gelatin solutions

501

with ZnO levels of 0 wt%, 1 wt%, 1.5 wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and

502

Z2-EG, respectively), and zeta potentials of the ZnO solutions at levels of 0 wt%, 1 wt%, 1.5 wt%,

503

and 2 wt% (denoted as Z0, Z1, Z1.5, and Z2, respectively).

504

Type

Apparent viscosity (Pa·s)

Zeta potential (mV)

Type

Zeta potential (mV)

Z0-EG

2.05 ± 0.06 a a

1.47 ± 0.21 a

Z0

3.89 ± 0.64 c

Z1-EG

1.77 ± 0.07 a

1.08 ± 0.16 ab

Z1

6.98 ± 0.37 a

Z1.5-EG

1.30 ± 0.12 b

0.87 ± 0.10 b

Z1.5

5.73 ± 0.26 b

Z2-EG

1.78 ± 0.03 a

1.00 ± 0.12 ab

Z2

3.94 ± 0.25 c

a

Lowercases indicated statistical significance (p < 0.05).

505

32

ACS Paragon Plus Environment

Page 33 of 37

Journal of Agricultural and Food Chemistry

506

Table 2. Atomic concentration percentages of C, O, N and ZnO examined by XPS of the

507

ethylcellulose/gelatin fibers with ZnO levels of 0 wt%, 1 wt%, 1.5 wt%, and 2 wt% (denoted as

508

Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG, respectively). Surface element composition Type

509 510

a

Zn (Theoretical) (%)

Zn ratio (%)a

C (%)

O (%)

N (%)

Zn (%)

Z0-EG

65.7

27.0

7.3

-

0

0

Z1-EG

66.9

25.2

6.9

1.0

2.1

47.6

Z1.5-EG

65.8

27.0

5.8

1.5

3.2

46.9

Z2-EG

66.5

25.2

5.2

3.0

4.2

71.4

Zn ratio = Zn (surface)/Zn (Theoretical)×100%.

33

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 34 of 37

511

Table 3. The 2θ value, FWHM value and crystal area of the ethylcellulose/gelatin fibers with ZnO

512

levels of 0 wt%, 1 wt%, 1.5 wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG,

513

respectively). Type

2θ (°)

Area

FWHM (β) (°)a

d (Å)b

Z0-EG

19.08

1376.46

6.32

0.222

Z1-EG

19.94

411.73

4.87

0.289

Z1.5-EG

20.04

538.50

4.82

0.292

Z2-EG

20.09

251.00

4.19

0.336

514

a

FWHM is the angular peak width at half maximum.

515

b

d = 0.9λ/(β cos(θ)), where λ is X-ray wavelength (1.541).

516

34

ACS Paragon Plus Environment

Page 35 of 37

Journal of Agricultural and Food Chemistry

517

Table 4. The tensile parameters of the ethylcellulose/gelatin fibers with ZnO levels of 0 wt%, 1 wt%,

518

1.5 wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG, respectively).

519

Type

Elastic modulus (E, MPa)

Tensile strength (σy, MPa)

Elongation at break (εb, %)

Z0-EG

148.10 ± 14.33 a a

1.70 ± 0.16 b

1.30 ± 0.11 ab

Z1-EG

188.77 ± 32.67 a

2.39 ± 0.17 a

1.39 ± 0.22 ab

Z1.5-EG

203.24 ± 43.06 a

2.42 ± 0.24 a

1.57 ± 0.31 a

Z2-EG

134.28 ± 45.17 a

1.65 ± 0.18 b

0.98 ± 0.17 b

a

Lowercases indicated statistical significance (p < 0.05).

520 521

35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 36 of 37

522

Table 5. The surface swelling, water uptake and weight loss of the ethylcellulose/gelatin fibers with

523

ZnO levels of 0 wt%, 1 wt%, 1.5 wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG,

524

respectively) after immersed in water for 24 h.

525

a

Type

Surface swelling (SW, %)

Water uptake (WT, %)

Weight loss (WL, %)

Z0-EG

-13.98 ± 0.63 a a

536.41 ± 21.46 a

13.78 ± 0.67 a

Z1-EG

-13.22 ± 0.33 a

297.90 ± 23.19 b

12.44 ± 0.97 ab

Z1.5-EG

-12.55 ± 0.13 b

209.31 ± 14.38 c

9.64 ± 0.86 b

Z2-EG

-12.08 ± 0.43 b

203.61 ± 20.71 c

9.33 ± 0.97 b

Lowercases indicated statistical significance (p < 0.05).

526

36

ACS Paragon Plus Environment

Page 37 of 37

Journal of Agricultural and Food Chemistry

527

Table 6. The inhibition zone of the ethylcellulose/gelatin fibers with ZnO levels of 0 wt%, 1 wt%,

528

1.5 wt%, and 2 wt% (denoted as Z0-EG, Z1-EG, Z1.5-EG, and Z2-EG, respectively).

529

a

Type

E. coli (mm/mg)

S. aureus (mm/mg)

Z0-EG

0 ± 0 da

0±0d

Z1-EG

0.69 ± 0.03 c

0.75 ± 0.020 c

Z1.5-EG

1.30 ± 0.04 b

1.17 ± 0.04 b

Z2-EG

1.61 ± 0.05 a

1.33 ± 0.06 a

Lowercases indicated statistical significance (p < 0.05).

37

ACS Paragon Plus Environment