Emulsifying Properties of Oxidatively Stressed Myofibrillar Protein

Mar 7, 2017 - Lab of Meat Processing and Quality Control of EDU, College of Food Science and Technology, Synergetic Innovation Center of Food Safety ...
0 downloads 0 Views 5MB Size
Subscriber access provided by University of Newcastle, Australia

Article

Emulsifying Properties of Oxidatively Stressed Myo#brillar Protein Emulsion Gels Prepared with (-)-Epigallocatechin-3-gallate and NaCl Lin Chen, Na Lei, Shuangxi Wang, Xing-lian Xu, Guanghong Zhou, Zhixi Li, and Xianchao Feng J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b05517 • Publication Date (Web): 07 Mar 2017 Downloaded from http://pubs.acs.org on March 8, 2017

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 35

Journal of Agricultural and Food Chemistry

1

Emulsifying Properties of Oxidatively Stressed Myofibrillar Protein Emulsion Gels

2

Prepared with (-)-Epigallocatechin-3-gallate and NaCl

3

Lin Chen,‡, Na Lei, Shuangxi Wang, Xinglian Xu‡, Guanghong Zhou‡, Zhixi Li, Xianchao

4

Feng *,

5



6

Yangling, Shaanxi 712100, China

7



8

Synergetic Innovation Center of Food Safety and Nutrition, Nanjing Agricultural University,

9

Nanjing, Jiangsu 210095, China

College of Food Science and Engineering, Northwest A&F University, No. 22 Xinong Road,

Lab of Meat Processing and Quality Control of EDU, College of Food Science and Technology,

10 11 12

*Corresponding author-Xianchao Feng

13

Associate Professor, College of Food Science and Engineering, Northwest A&F University, No.

14

22 Xinong Road, Yangling, Shaanxi, China 712100. Email address: [email protected]

15

Tel/Fax: 86029-87092486.

16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 1 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

32

ABSTRACT: The dose-dependent effects of (-)-epigallocatechin-3-gallate (EGCG; 0, 100 or

33

1000 ppm) on the textural properties and stability of a myofibrillar protein (MP) emulsion gel

34

were investigated. EGCG addition significantly inhibited formation of carbonyl but promoted the

35

loss of both thiol and free amine groups. Addition of EGCG, particularly at 1000 ppm, initiated

36

irreversible protein modifications, as evidenced by surface hydrophobicity changes, patterns in

37

SDS-PAGE, and differential scanning calorimetry (DSC). These results indicated that MP was

38

modified by additive reactions between the quinone of EGCG and thiols and free amines of

39

proteins. These adducts increased cooking loss and destabilized the texture, especially at a high

40

dose of EGCG. Confocal laser scanning microscopy (CLSM) and scanning electron microscopy

41

(SEM) images clearly indicated the damage to the emulsifying properties and the collapse of the

42

internal structure when the MP emulsion gel was treated with a high dose of EGCG. A high

43

concentration of NaCl (0.6 M) improved modification of MP and increased deterioration of

44

internal structure, especially at the high dose of EGCG (1000 ppm), resulting in extremely

45

unstable emulsifying properties of MP emulsion gel.

46

KEYWORDS: emulsion gel; myofibrillar protein; EGCG; chemical properties; microstructural

47

properties; confocal laser scanning microscopy

48 49 50 51 52 53 2 / 35

ACS Paragon Plus Environment

Page 2 of 35

Page 3 of 35

54

Journal of Agricultural and Food Chemistry



INTRODUCTION

55

Myofibrillar proteins (MPs) are the main constituent of total muscle protein (55% to 60%).1

56

The quality of MP is mainly responsible for the properties of processed meat products.2 It is

57

necessary to start with high-quality MP to produce a well-formed, uniformly textured,

58

physicochemically stable meat product, as the MP serves to stabilize fat globules and entrap

59

water.1, 3 MP is prone to attack by the reactive oxygen species that are generated during food

60

processing (i.e., cooking, drying and storing), which results in direct conversion of amino acid

61

residues to carbonyls and cross-links.4 These oxidative modifications can improve the

62

polymerization and aggregation of proteins, but can also alter the secondary and tertiary structure

63

of proteins, leading to changes in the physical properties of proteins, such as hydrophobicity.5

64

Consequently, protein oxidation has been linked to deteriorated quality of meat products, such as

65

loss in juiciness, increase in cooking loss, and toughness of meat.6 There is an increasing volume

66

of research on using antioxidants to prevent protein oxidation in the meat product industry.7

67

There is a growing focus on using natural preservatives due to their safety and their perceived

68

non-toxicity compared to synthetic additives in the meat industry.7 Extracts from herbs and spices

69

are widely used as natural alternatives. These extracts provide promising antioxidant effects that

70

are attributed to phenolic compounds, due to their hydrogen-donating capacity and

71

metal-chelating potential. Polyphenols are of upmost interest due to their health-beneficial

72

effects.8 Phenolic derivatives have been widely and commonly used to inhibit oxidative

73

modification in meat products.9

74

However, aside from the positive effect on scavenging free radicals, some researchers found

75

that polyphenols had negative effects on the quality of processed meat products. The addition of a 3 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

76

phenolic-rich extract from dog-rose to beef patties had an antioxidant effect on lipids and proteins

77

but increased cooking loss and storage loss.10 The addition of 1000 ppm green tea extract

78

decreased the content of thiobarbituric acid reactive substances (TBARS) and carbonyls but

79

deteriorated the emulsion properties of meat products, resulting in increased cooking loss and

80

texture instability.11 The addition of 500 ppm white grape extract reduced the TBARS and protein

81

carbonyls but also reduced the desired thiols in beef patties.12

82

In recent years, it has been found that polyphenols can be oxidized and converted to quinines,

83

which can irreversibly react with the thiol and amino moieties of proteins.8 Polyphenols react with

84

a cysteinyl thiol derivative (N-benzoylcysteine methyl ester) though formation of thiol-quinone

85

adducts under radical oxidation conditions.13 When 4-Methylcatechol was used to inhibit protein

86

oxidation, the formation of thiol-quinone adducts led to an additional loss of thiol groups.14

87

Reactive quinones, derived from rosmarinic acid, will form adducts with thiol compounds, such

88

as cysteine, glutathione, and peptides, derived from myosin.15 Moreover, quinones can also

89

irreversibly react with the anime groups of proteins.8, 16 Intermolecular disulphide (S-S) linkages

90

due to oxidation of thiol groups are one of the principal forces for gelation of muscle protein

91

during heating. A significant decrease in the storage modulus (G'), gel strength, and cook yield of

92

MP emulsion gels occurs when thiols of MP were modified by N-ethylmaleimide before

93

heating.17 Despite all these recent reports, the effects of thiol-quinone and amino-quinone adducts

94

between polyphenol and meat proteins on the quality of food products in processing are relatively

95

unreported.10-12 Furthermore, it is well known that high NaCl concentrations result in more

96

soluble MP with a swollen appearance, which enhances protein-protein interactions. However,

97

effects of NaCl concentrations on the adducts of thiol-quinone and amino-quinone have yet to be 4 / 35

ACS Paragon Plus Environment

Page 4 of 35

Page 5 of 35

98

Journal of Agricultural and Food Chemistry

thoroughly investigated.

99

Tea is the second-most widely consumed beverage worldwide, second only to water.18 Green

100

tea extract contains polyphenolic compounds known as catechins, which may constitute up to 30%

101

of the dry weight of tea leaves.18-19 The most abundant catechin is (-)-epigallocatechin-3-gallate

102

(EGCG), which has been reported to be responsible for many of the effects of green tea.20 Green

103

tea extracts are broadly used to inhibit lipid oxidation in emulsion-type meat products, e.g.

104

various bologna, wieners, and frankfurters, which generally contain at least 30% fat.21 However,

105

the influence of thiol-quinone and amino-quinone adducts between catechins and protein in these

106

meat products on the emulsifying properties has not been investigated.

107

In this work, the effects of EGCG addition on the emulsifying properties of pork MP emulsion

108

gel under controlled oxidizing conditions (H2O2) were investigated. Cooking loss and textural

109

profiles of MP emulsion gel were observed. Chemical, structural, thermal and rheological

110

properties of EGCG-modified MP were analyzed to discern the internal mechanism. The effects

111

of salt concentration on adducts between EGCG and MP were also studied.

112

 MATERIALS AND METHODS

113

Materials. Fresh pork from the Longissimus muscle of pig carcasses (large white crossbred,

114

within 2 days post mortem) was bought from a local supermarket (Haoyouduo, Yangling, Shaanxi,

115

China). Chemicals used in present study were of reagent grade and purchased from Aladdin

116

Industrial Corporation (Fengxian, Shanghai, China) or Sigma-Aldrich Co. (St. Louis, MO, USA).

117

Myofibrillar Protein (MP) Preparation. MP was prepared according to published methods,

118

with some modifications.5, 22 Briefly, 200 g of fresh pork was firstly homogenized with a blender

119

(Joyung Co., Ltd. Jinan, China) in 1 L isolation buffer [20 mM PBS, 150 mM NaCl, 25 mM KCl, 5 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

120

3 mM MgCl2, and 4 mM EDTA at pH 7.0]. The suspension was further homogenized with an

121

Ultra Turrax dispersing instrument (IKA T18-Digital, Staufen, Germany) in an ice bath. After

122

filtratration through gauze, the homogenate was spun twice at 2000 × g for 15 min at 4 °C. Then,

123

the pellets were washed twice with 4 L of 0.1 M NaCl solution. The collected pellets were washed

124

again with 4 L of 20 mM PBS pH 7.0. The protein concentration was measured using the Biuret

125

method.23

126

Oxidative Treatments with EGCG. MP suspensions were prepared with 20 mM PBS buffer

127

(pH 7.0). Nine different reaction mixtures (final protein concentration, 40 mg/mL) were made

128

with various EGCG (0, 100 and 1000 mg/kg protein) and NaCl (0, 0.2, and 0.6 M) concentrations

129

in five replicates according to a total factorial design. Mixtures were oxidized at 4 °C for 12 h

130

using 5 mM H2O2, 10 mM FeCl3, and 100 mM ascorbic acid. Sodium azide (final concentration,

131

0.02% by wt) was used to prevent microbial growth in the mixtures. Addition of Trolox

132

C/butylated hydroxytoluene (BHT)/ EDTA (each at 1 mM final concentration) was applied to stop

133

the oxidation reaction. A non-oxidized, EGCG-free MP suspension was used as the control.

134

Carbonyl Analysis. The carbonyl levels in the treated MP samples were analyzed according to

135

the method of Oliver et al.24 Briefly, carbonyl levels, in the form of hydrazones, were detected by

136

2,4-dinitrophenylhydrazine (DNPH). Absorbance was measured at 370 nm with an absorption

137

coefficient of 22000 M−1 cm−1 to calculate the carbonyl content.

138

Free Amines. Free anime levels in the treated MP samples were analyzed as described by Liu

139

et al.25 Briefly, 2,4,6-trinitrobenzenesulfonic acid (TNBS) was used to measure free anime levels

140

with absorbance at 420 nm.

141

Total Sulfhydryl Content. Total sulfhydryl content in the treated MP was measured using 6 / 35

ACS Paragon Plus Environment

Page 6 of 35

Page 7 of 35

Journal of Agricultural and Food Chemistry

142

absorbance at 412 nm with the indicator 5,5′-Dithiobis (2-nitrobenzoic acid) (DTNB).26-27

143

Absorption coefficients of 13600 M-1 cm-1 were used for calculating the total sulfhydryl content.

144

Surface

Hydrophobicity.

Surface

hydrophobicity

was

determined

using

145

1-anilino-8-naphthalenesulfonate (ANS) as hydrophobic fluorescent probe. A series of MP

146

solutions with a protein concentration ranging from 0.05 to 2.0 mg/mL were prepared. An aliquot

147

of 4 mL of each MP solution was mixed by vortex with 20 µL of 8 mM ANS solution. The

148

fluorescence intensity of all samples was measured after static reaction for 2 min using a

149

PerkinElmer LS-55 spectrofluorometer (Waltham, MA, USA). The excitation and emission

150

wavelengths were 390 and 470 nm, respectively. Both excitation slit width and emission slit width

151

were set at 10 nm. Protein surface hydrophobicity was expressed by linear regression slope on a

152

curve plotting fluorescence intensity against protein concentration.

153

Electrophoresis. Polymerization of protein was investigated using sodium dodecyl sulfate

154

polyacrylamide gel electrophoresis (SDS−PAGE) with a 4% polyacrylamide stacking gel and 12.5%

155

separating gel.5, 28 The MP solution (2 mg/mL) was mixed with 4-fold sample buffer containing

156

10% β-mercaptoethanol (βME) or 0% βME, and then boiled for 5 min. After electrophoresis using

157

10 µL of each sample, 0.1% Coomassie Brilliant Blue was used to show protein bands on the gel.

158

After eliminating the free dye, the band profile on each gel was recorded with a Gel Doc XRTM

159

System (Bio-Rad Laboratories, Hercules, CA).

160

Thermal Analysis by Differential Scanning Calorimetry (DSC). DSC was performed on a

161

TA Instruments Q2000 DSC (TA Instruments, New Castle, DE, USA). Raw samples (40 mg/mL)

162

were accurately weighed (approximately 15 mg) and sealed in hermetic aluminum pans. An

163

empty pan was used as control. Thermal scan was performed from 30 to 80 °C at a 5 °C /min rate. 7 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

164

The temperature maximum (Tm) for protein transition was estimated using Universal Analysis

165

software (TA Instruments, Ver. 4.3A).

166

Emulsion Gel Preparation. Oil-in-water (O/W) emulsions were prepared by dispersing

167

soybean oil in MP suspension (v:v = 1:4) using an Ultra Turrax dispersing instrument (IKA

168

T18-Digital, Staufen, Germany) operating at 11000 rpm for 1 min. Five grams of each emulsion

169

were added into glass vials (2.5 cm i.d., 5 cm height) and incubated from room temperature to

170

74°C, and stayed at 74°C for 10 min.

171

Cooking Loss. Raw emulsions were weighed before cooked. The emulsion gels were chilled to

172

room temperature after cooking to a core temperature of 74 °C. The weight of the samples was

173

recorded again after removing the juice from the surface of the emulsion gels. Cooking loss was

174

calculated as: ୑బ ି୑భ

175

Cooking loss (%) =

176

where M0 is the weight of the sample prior to cooking and M1 is the weight of the sample after

177

୑భ

× 100%

cooking and chilling.

178

Gel Strength. The gel strength of the MP emulsion gels was measured using a TA-XT plus

179

texture analyzer (Stable Micro Systems Co. Ltd., Surrey, UK) fitted with a cylindrical probe

180

(P/0.5). Analysis was performed using the following conditions: pre-speed, 1.00 mm/s; trigger

181

force, 5 g; test speed, 1.00 mm/s; and post-speed, 1.00 mm/s. The data acquisition rate was 200

182

pps.

183

Dynamic Rheological Testing During Gelation. Viscoelastic characteristics of emulsion were

184

measured during the heat-induced gelation with a Model AR1000 rheometer (TA Instruments,

185

West Sussex, UK) in an oscillatory mode equipped with parallel plates geometry (40 mm 8 / 35

ACS Paragon Plus Environment

Page 8 of 35

Page 9 of 35

Journal of Agricultural and Food Chemistry

186

diameter). Five gram of each sample was added between parallel plates (1.0 mm gap) and silicon

187

oil was used to cover the exposed rim to prevent dehydration. Gelation was induced by heating

188

the protein emulsion from 30 to 80 °C at a 2 °C/min rate. Shear force of samples was measured

189

with a sinusoidal strain at a 2% amplitude and an oscillating frequency of 0.1 Hz during heating.

190

Changes in the storage modulus (G′) were monitored continuously.29

191

Confocal Laser Scanning Microscopy (CLSM). The microstructure of the MP emulsion gels

192

was imaged with a confocal laser scanning microscope (A1R, Nikon Inc., Tokyo, Japan). A Fast

193

Green and Nile Red mixture (0.038%, w/v) was used to stain the protein and oil droplets,

194

respectively. The images were taken with a 10 x magnification lens by CLSM.

195

Scanning Electron Microscopy (SEM). The morphology of the gels was observed using a

196

Hitachi-S-4800 field emission scanning electron microscope (Hitachi High Technologies Corp.,

197

Tokyo, Japan). Cubic samples (1 × 0.8 × 0.5 cm3) obtained from gels were fixed using 2.5%

198

glutaraldehyde in 0.1 M PBS (pH 7.4) for 24 h. Samples were then washed using 0.1 M PBS (pH

199

7.4) for 20 min, and then fixed in 1% osmium tetraoxid prepared using PBS buffer (pH 7.4) for 5

200

h. After washing three times with 0.1 M PBS (pH 7.4), fixed samples were then dehydrated in

201

gradient ethanol solutions (50, 60, 70, 80, 90, and 95%, and three times with 100%).

202

Statistical Analysis. All measurements were repeated five times. Data from the analysis were

203

collected and subjected to statistical analyses. Two-way analysis of variance (ANOVA) was

204

applied to assess the effect of the different concentrations of EGCG and NaCl (SPSS 20.0,

205

Chicago, IL, USA). A LSD was applied when ANOVA found significant differences between

206

different treatments. The statistical significance was set at P < 0.05.

207

 RESULTS AND DISCUSSION 9 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

208

Results of this work showed that the addition of EGCG and NaCl significantly affected all

209

measured parameters, such as the chemical and emulsifying properties of MP under oxidative

210

stress. Cooking loss, gel strength and microstructure of the modified MP emulsion gels were also

211

influenced by addition of EGCG and NaCl.

212

Effect of EGCG and NaCl on the Chemical and Structural Properties of MP. Reactive

213

oxygen species (ROS) readily oxidize the lysine, arginine, proline and threonine residues within

214

proteins into carbonyls (aldehydes and ketones). Carbonyl levels are a well-known index of the

215

level of protein oxidation in meat products during processing.6 Application of oxidation stress to

216

pork MP significantly increased carbonyl content compared to the non-oxidized MP (Figure 1 A).

217

The addition of EGCG reduced the carbonyl content, especially for MP treated with a high dose

218

of EGCG (Figure 1 A). This result was in accordance with previous studies showing that

219

polyphenols prevented the formation of carbonyls in protein.15, 30

220

Free amine content is another indicator of protein oxidation,5 since the ε-NH2 groups of some

221

amino residues can be readily converted into carbonyls through a deamination process under

222

oxidation stress.22 Under oxidative stress in the present study, the ε-NH2 group content was higher

223

in the absence of EGCG (Figure 1 B). In the presence of EGCG, the ε-NH2 groups showed

224

additional loss that was EGCG-dose dependent (Figure 1 B). This could be due to the formation

225

of amine–quinone adducts between the quinone of EGCG and the free amines of protein under

226

oxidative stress, since quinones can irreversibly react with the anime groups of proteins.8, 16

227

The free amine and carbonyl contents of the oxidatively stressed MP were also significantly

228

affected by NaCl addition (Table 1, Figure 1A & B). More soluble MP was available at higher

229

NaCl concentration.31 Consequently, the ε-NH2 groups of lysine residues on the protein surface 10 / 35

ACS Paragon Plus Environment

Page 10 of 35

Page 11 of 35

Journal of Agricultural and Food Chemistry

230

were more prone to be attacked by the quinone groups of EGCG at high NaCl concentration (0.6

231

M NaCl).5

232

MP is rich in thiol groups, which are susceptible to attack by ROS and subsequent conversion

233

to intra- and intermolecular disulfide bond linkages.5 Compared to the non-oxidized MP (control),

234

the thiol group content of oxidized MP was significantly lower in the absence of EGCG (Figure 1

235

C). Furthermore, the thiol group levels showed an EGCG-dose-dependent decrease. This could be

236

caused by the formation of thiol–quinone adducts.14-15 These results were in accordance with

237

previous studies.11, 30 The thiol group levels in oxidized MP were also significantly decreased with

238

the increases in NaCl concentration (Table 1). This could also be explained by the same changes

239

in protein status that affected the ε-NH2 groups.31

240

Protein surface hydrophobicity is a marker of protein unfolding. The surface hydrophobicity of

241

the oxidized MP, in the absence of EGCG, was significantly higher (P < 0.05) than the

242

non-oxidized MP (Control, Figure 1 D). The surface hydrophobicity of oxidized MP significantly

243

decreased in the presence of EGCG, especially at the high dose of EGCG (1000 ppm) (Figure 1

244

D). This decrease in surface hydrophobicity with increased EGCG may be due to protein

245

aggregation caused by EGCG addition, which would partially shield the effect of unfolding in

246

oxidized MP. Protein aggregation has been proposed to be the reason for loss of surface

247

hydrophobicity in oxidized MP samples.3 A combination of high NaCl and high EGCG

248

concentrations seems to result in greater aggregation of oxidized MP (Table 1 & Figure 1 D).

249

Consequently, the oxidized MP treated with 1000 ppm EGCG had the lowest surface

250

hydrophobicity with 0.6 M NaCl (Figure 1 D).

251

The interaction between EGCG and NaCl had significant effects on these chemical and 11 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

252

structural markers of the protein (carbonyl, free amine, thiol group and surface hydrophobicity;

253

Table 1). The MP gained a swollen appearance at high NaCl concentration, which would promote

254

modification by quinone of EGCG and the ·OH generated by the hydroxyl-radical-generating

255

system.31

256

Effect of EGCG and NaCl on the DSC Thermal Properties of MP. Non-oxidized MP

257

(control) showed two clear endothermic peaks with Tmax values at 60.7 and 69.6 °C (Figure 2); the

258

first was attributed to myosin, and the second to actin.32 Oxidative stress at 5 mM H2O2 tended to

259

diminish the second peak, which was consistent with previous studies.3 The addition of EGCG

260

remarkably shifted all the peaks to lower temperatures, especially for the MP treated with the

261

highest dose of EGCG (Figure 2). The total heat of denaturation (∆H) was remarkably reduced

262

upon oxidation, and EGCG addition further exacerbated the decrease in ∆H, especially for the MP

263

treated with the high dose of EGCG (Figure 2). The high concentration of NaCl (0.6 M) lowered

264

the Tmax and ∆H of MP compared to MP with low concentration of NaCl (0 and 0.2 M) (Figure 2).

265

These results indicated that the higher EGCG dose led to significant MP denaturation, which

266

could be accelerated by high NaCl concentration.

267

Effect of EGCG and NaCl on the SDS−PAGE Patterns of MP. Oxidative stress can result in

268

intra- and inter-molecular cross-links between proteins.4-5 Compared to the non-oxidized MP

269

(control), the band intensities in the oxidized MP for both myosin heavy chain (MHC) and actin

270

decreased. In oxidized MP, the intensities of both bands decreased with increasing EGCG,

271

resulting in formation of polymers appearing at the top of the stacking gel (Figure 3 A). This

272

indicated that addition of EGCG improved the polymerization of MP, resulting in the reduction of

273

MHC and actin bands (Figure 3 A). However, most MHC and actin bands were recovered when 12 / 35

ACS Paragon Plus Environment

Page 12 of 35

Page 13 of 35

Journal of Agricultural and Food Chemistry

274

treated with βME except for the MP with high dose of EGCG (Figure 3 B). This indicated that

275

these polymers were largely caused by disulfide bonds.5, 33 In the reduced samples, oxidized MP

276

treated with EGCG clearly had more polymer bands on the top of both the stacking and separating

277

gels compared to the oxidized MP without EGCG. This indicated that more covalent bonds that

278

were not disulfides might be formed due to the addition of EGCG. Combined with the data on the

279

EGCG-dose dependent decrease of thiol groups and free amines, this indicates that quinone-thiol

280

adducts and amine-quinone adducts likely formed. Jongberg et al. found that thiol-quinone

281

adducts were formed between quinone of 4-methylcatechol and protein thiol groups.14 Tang et al.

282

also found that quinone of rosmarinic acid can react with compounds containing thiol groups

283

(R−SH), forming thiol−quinone adducts.15 Moreover, quinone of polyphenol can react with amino

284

side chains of polypeptides.8,34 Hence, amine-quinone adducts can contribute to the formation of

285

non-disulfide polymerizations. Cysteine and lysine residues of protein have nucleophilic side

286

chains that may form covalent bonds with the quinone formed by oxidation of the phenol by

287

nucleophilic addition. The proposed reaction mechanism in Figure 4 explains how the

288

thiol−quinone adducts and the amine-quinone adducts could form.

289

Effect of EGCG and NaCl on the Cooking Loss and Gel Strength of MP Emulsion Gel.

290

Cooking loss was analyzed to access the emulsifying stability of different MP emulsion gels

291

(Figure 5). EGCG addition had significant effects on cooking loss of the MP emulsion gel (Table

292

1). At each NaCl concentration, the MP emulsion gel containing EGCG had significantly higher

293

cooking loss than the MP emulsion gel without EGCG, especially for 1000 ppm EGCG (0 M

294

NaCl, 36.96%; 0.2 M NaCl, 41.88%; 0.6 M NaCl, 39.75%) (Figure 5). At lower NaCl

295

concentrations, the MP emulsion gel treated with 1000 ppm EGCG had a 1.88- to 2.16-fold higher 13 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

296

cooking loss than MP emulsion gel without EGCG (Figure 5). NaCl addition significantly

297

increased cooking loss of the MP emulsion gel (Table 1). With 0.6 M NaCl, the cooking loss of

298

the MP emulsion gel treated with 1000 ppm EGCG was 9.46-fold higher compared to the MP

299

emulsion gel without EGCG (Figure 5). Covalent bonds induced by heating have important

300

contributions to the homogeneous, three-dimensional network of the gel, especially disulfide

301

bonds. The formation of thiol−quinone and amine-quinone adducts blocked the formation of

302

covalent bonds, such as disulfide bonds and active carbonyl−NH2 interactions.5 This resulted in a

303

poor protein network, and consequently, higher cooking loss of MP emulsion gel treated with

304

1000 ppm EGCG, especially with 0.6 M NaCl (Figure 5).

305

Gel strength was also measured to evaluate the properties of the MP emulsion gels. Both

306

EGCG and NaCl had significant effects on gel strength of the MP emulsion gel (Table 1). At 0 M

307

NaCl and 0.2 M NaCl, the high dose of EGCG significantly decreased the gel strength (Figure 5).

308

The high dose of EGCG modified the thiol groups and free amines by additive reaction, hence

309

blocking the cross-linking during heat-induced emulsion gel processing. Consequently, the high

310

dose of EGCG decreased the force needed to compress the sample to reach a given deformation

311

(gel strength). Jongberg et al. found that a high dose of green tea extract increased the cooking

312

loss and decreased the firmness/hardness of a meat emulsion gel, indicating that a high dose of

313

EGCG damaged the emulsifying stability of the emulsion gels.11 However, the combination of

314

high EGCG (1000 ppm) and high NaCl (0.6 M) significantly increased the gel strength of the MP

315

emulsion (Figure 5). At high ionic strength (0.6 M NaCl), it is possible that the MP could have a

316

higher degree of modification, resulting in highly denatured MP and high level of aggregation.

317

Consequently, the MP emulsion gel treated with high EGCG and high NaCl was more compact 14 / 35

ACS Paragon Plus Environment

Page 14 of 35

Page 15 of 35

Journal of Agricultural and Food Chemistry

318

due to protein aggregation caused by the high hydrophobic force. Moreover, the molecular size of

319

the MP could become larger in the presence of 1000 ppm EGCG at 0.6 M NaCl, which might

320

hinder relative motion between protein molecules. As a result, at 0.6 M NaCl, the MP emulsion

321

gel treated with 1000 ppm EGCG had higher gel strength compared to MP emulsion gel treated

322

with 0 or 100 ppm EGCG (Figure 5). The negative charges on the surface of MP could be

323

neutralized by the positive sodium ions at higher NaCl concentration, which may also cause

324

aggregation of proteins.35 This could also account for the higher gel strength of the MP emulsion

325

gel treated with 1000 ppm EGCG at 0.6 M NaCl compared with that at 0 and 0.2 M NaCl.

326

Dynamic Rheological Properties of MP Emulsion Gel. Dynamic oscillatory analyses are

327

commonly used to assess the viscoelastic characteristics of emulsion gels.36 The storage modulus

328

(G′), which represents the amount of recoverable energy stored in the elastic gel, of the MP

329

emulsion gel was analyzed and plotted as G′ versus temperature (Figure 6). The oxidized sample

330

had significantly higher final G′ compared to the control (Figure 6), suggesting that interactions

331

between MP were promoted by oxidation. 16 The G′ of the MP emulsion gel showed an EGCG

332

dose-dependent decrease at 0 and 0.2 M NaCl, especially in the sample prepared with 1000 ppm

333

EGCG (Figure 6 A & B). It is possible that the high dose of EGCG promoted the formation of

334

thiol-quinone and amino-quinone adducts, which blocked the formation of intermolecular

335

covalent bonds between proteins and prevented formation of ordered, heat-induced,

336

three-dimensional protein network. As a result, G′ was reduced in the MP emulsion gel treated

337

with a high dose of EGCG at a low NaCl concentration. The G′ of the MP emulsion gel treated

338

with a low dose of EGCG (100 ppm) was higher than the control (Figure 6), suggesting that

339

addition of 100 ppm EGCG did not significantly prevent the cross-linking induced by heating 15 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

340

under oxidative stress nor the formation of other chemical interactions/bonds between MP and did

341

not inhibit the gelation of MP compared to the control.37 Previous studies have shown that the

342

structure and function of MP can be affected by ionic strength,31 and that high ionic strength (>

343

0.4 M NaCl) can improve the transverse expansion of MP.38 The swollen MP enhanced the

344

protein−protein interactions and the formation of a thicker layer surrounding the lipid droplet to

345

improve the stability of the MP emulsion gel treated with high NaCl concentration.31, 39 It is

346

surprising that the MP treated with a high dose of EGCG (1000 ppm) had significantly higher G′

347

at 0.6 M NaCl (Figure 6 C). At high NaCl concentration (0.6 M NaCl), the MP was in an evenly

348

swollen state, which would aggravate unfolding of protein due to interaction with EGCG. The

349

significantly higher hydrophobic force initiated by the high surface hydrophobicity might cause

350

the shrinking of the MP gel with a compact structure during heating, resulting in higher G′ (Figure

351

6 C).31, 40 Moreover, the molecular size of the swollen MP would become larger due to covalent

352

and non-covalent interactions with high EGCG (1000 ppm), which might increase entanglements

353

between protein chains. This could further improve the G′ of MP emulsion gel. Therefore, at low

354

NaCl concentrations, the EGCG-mediated prevention of covalent interactions between proteins

355

caused a decrease of G' (Figure 6 A & B). On the other hand, at high NaCl concentration,

356

hydrophobic forces induced by unfolding of protein caused by high level of EGCG contributed to

357

the increase of G' (Figure 6 C).

358

Effect of EGCG and NaCl on the Microstructure of MP Emulsion Gel. CLSM imaging was

359

used to reveal distribution of lipid droplets in the MP emulsion gels (Figure 7), especially since

360

cooking loss of emulsion gels is largely dependent upon gel microstructure.31 The protein network

361

is formed by cross-links upon heating, which can bind and embed water and oil droplets.41 Oil 16 / 35

ACS Paragon Plus Environment

Page 16 of 35

Page 17 of 35

Journal of Agricultural and Food Chemistry

362

droplets had uniform size and even distribution in the control MP emulsion gel (Figure 7 A). The

363

oxidized MP emulsion gel treated with 0 or 100 ppm EGCG showed distributions of oil droplets

364

similar to control (Figure 7). These images indicated that oxidation (5 mM H2O2) and low level of

365

EGCG did not significantly jeopardize the internal structure of the MP emulsion gels. Stability of

366

MP emulsion gels was evenly matched compared to the control, as evidenced by the similar

367

cooking loss and gel strength values (Figure 5). In the presence of NaCl, the sizes of the oil

368

droplets coated with protein become smaller and more evenly dispersed in the oxidized MP

369

emulsion gel treated with 0 or 100 ppm EGCG, especially at high NaCl concentration (0.6 M)

370

(Figure 7). When the concentration of EGCG was at 1000 ppm and NaCl levels increased, oil was

371

coalesced into larger droplets (Figure 7). The oil droplets in the MP emulsion gels were likely

372

stabilized by an interfacial protein film and the gel matrix that restricted their movement.31 At

373

1000 ppm EGCG, adducts between quinone of EGCG and –SH and –NH2 groups of the amino

374

acid residues might prevent interaction covalent bonds, such as S–S and carbonyl–NH2 covalent

375

bonds, between interfacial proteins and proteins in the gel matrix, resulting in coalescence of and

376

larger emulsion droplets (Figure 4 & 7).

377

The microstructure of the protein matrix was investigated using SEM images as well (Figure 8).

378

The unoxidized MP emulsion gel (control) exhibited a continuous protein network structure

379

(Figure 8). With 0 or 0.2 M NaCl, oxidation contributed to light protein aggregation compared to

380

control (Figure 8). Moreover, the higher dose of EGCG (1000 ppm) resulted in further MP

381

aggregation and the more compact gel structure, which might be caused by the higher

382

hydrophobic forces that come with higher levels of modification. The MP emulsion gel had a

383

more porous three-dimensional network at 0.6 M NaCl concentration compared to MP emulsion 17 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

384

gels at low NaCl concentrations in absence of EGCG (Figure 8). However, addition of 1000 ppm

385

EGCG and 0.6 M NaCl led to severe protein aggregation, shrinking the gel matrix (Figure 8). The

386

shrinking of the matrix could lead to the higher cooking loss and gel strength seen in the MP

387

emulsion gel treated with 1000 ppm EGCG and 0.6 M NaCl (Figure 5).

388

Interfacial protein interactions and protein in continuous phase can prevent the coalescence of

389

lipid droplets, by forming a three-dimensional network in meat products.17 Addition of 1000 ppm

390

EGCG significantly increased cooking loss but deteriorated the gel strength of the MP emulsion

391

gel. At low NaCl concentration (≤0.2 M NaCl), high dose of EGCG decreased the gel strength

392

and increased cooking loss (Figure 5). However, at high NaCl concentration (0.6 M NaCl), a high

393

dose of EGCG increased both the gel strength and the cooking loss (Figure 5). Together, these

394

results indicated that addition of high dose of EGCG jeopardized the stability of the MP emulsion

395

gel. A schematic representation of the formation of MP emulsion gels with or without EGCG is

396

shown in Figure 9. The loss of thiol groups and free amines and the appearance of polymers in the

397

reducing SDS-PAGE images indicated that thiol-quinone and amino-quinone adducts were

398

formed. These adducts blocked the formation of covalent bonds between the proteins absorbed on

399

the surface of lipid droplets and the proteins in the continuous phase (non-absorbed proteins),

400

resulting in a poor three-dimensional gel network. Consequently, the cooking loss of the

401

EGCG-treated emulsion gel significantly increased and the texture became softer with

402

significantly lower gel strength (Figure 5). Moreover, a high dose of EGCG promoted further

403

unfolding of the protein compared to the oxidized MP.30 Hydrophobic forces caused protein

404

aggregation and shrunk the emulsion gel, which further increased the cooking loss. Shrinking and

405

aggregation of the emulsion gel was shown by the compact microstructure of the emulsion gel 18 / 35

ACS Paragon Plus Environment

Page 18 of 35

Page 19 of 35

Journal of Agricultural and Food Chemistry

406

treated with the high dose of EGCG (Figure 8). However, high NaCl concentration (0.6 M NaCl)

407

resulted in evenly swollen MP, which aggravated unfolding of protein due to interaction with

408

EGCG. The hydrophobic forces might be too high to mask the effect of covalent adducts on the

409

texture of the emulsion gel. Consequently, the emulsion gel treated with the high dose of EGCG

410

had much more compact microstructure (Figure 8) and higher gel strength compared to other

411

samples (Figure 5) at high NaCl concentration (0.6 M NaCl). Other interactions, such as hydrogen

412

bonds or ionic bonds, between protein molecules were not depicted in Figure 9.

413

In conclusion, EGCG can be oxidized and converted to quinine, which can irreversibly react

414

with the thiol and amino moieties in protein under oxidative stress. These irreversible adducts are

415

more prevalent at higher EGCG doses and prevent interfacial proteins and gel matrix proteins

416

from interacting through covalent bonds, such as S–S and carbonyl–NH2 covalent bonds. In

417

addition, the high dose of EGCG induced protein aggregation. Consequently, the MP emulsion gel

418

structure was jeopardized, and its emulsifying properties became unstable, resulting in

419

significantly higher cooking loss and texture deterioration. Notably, high NaCl concentration

420

magnified the modification of MP by the high dose of EGCG, contributing to the poorer quality

421

of the MP emulsion gel. The results of present paper may help the meat processing industry to

422

fully understand how antioxidants such as EGCG interact with NaCl and alter emulsion-type meat

423

products. This understanding can lead to optimized formulations designed to produce high quality

424

muscle-based foods.

425

ACKNOWLEDGMENTS

426

We would like to thank Dr Anita K. Snyder, Donald Danforth Plant Science Center, America

427

for her helpful advices and assistance with the English language. This work was supported by the 19 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

428

National Natural Science Fund for Young Scholars (Grant No.: 31601497; 31401515), the China

429

Postdoctoral Science Foundation Project (Grant No.: 2016M591857). All authors read,

430

commented on, and approved the final manuscript.

431

REFERENCES

432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467

1.

Sun, X. D.; Holley, R. A., Factors Influencing Gel Formation by Myofibrillar Proteins in Muscle Foods.

Compr. Rev. Food Sci. F. 2011, 10 (1), 33–51. 2.

Lund, M. N.; Heinonen, M.; Baron, C. P.; Estevez, M., Protein oxidation in muscle foods: A review. Mol. Nutr.

Food Res. 2011, 55 (1), 83-95. 3.

Li, C.; Xiong, Y. L.; Chen, J., Oxidation-induced unfolding facilitates myosin cross-linking in myofibrillar protein by

microbial transglutaminase. J. Agr. Food Chem. 2012, 60 (32), 8020-8027. 4.

Chen, L.; Hackman, R. M.; Li, C.; Xu, X.; Zhou, G.; Feng, X., Different Physicochemical, Structural and Digestibility

Characteristics of Myofibrillar Protein from PSE and Normal Pork before and after Oxidation. Meat Sci. 2016, 121, 228-237. 5.

Feng, X.; Li, C.; Ullah, N.; Hackman, R. M.; Chen, L.; Zhou, G., Potential Biomarker of Myofibrillar Protein

Oxidation in Raw and Cooked Ham: 3-Nitrotyrosine Formed by Nitrosation. J. Agr. Food Chem. 2015, 63 (51), 10957-10964. 6.

Estévez, M., Protein carbonyls in meat systems: A review. Meat Sci. 2011, 89 (3), 259-279.

7.

Vossen, E.; Doolaege, E. H. A.; Moges, H. D.; Meulenaer, B. D.; Szczepaniak, S.; Raes, K.; Smet, S. D., Effect of

sodium ascorbate dose on the shelf life stability of reduced nitrite liver pates. Meat Sci. 2012, 91 (1), 29-35. 8.

Ozdal, T.; Capanoglu, E.; Altay, F., A review on protein–phenolic interactions and associated changes. Food Res.

Int. 2013, 51 (2), 954–970. 9.

Karre, L.; Lopez, K.; Getty, K. J. K., Natural antioxidants in meat and poultry products. Meat Sci. 2013, 94 (2),

220-7. 10. Utrera, M.; Morcuende, D.; Rui, G.; Estévez, M., Role of Phenolics Extracting from Rosa canina L. on Meat Protein Oxidation During Frozen Storage and Beef Patties Processing. Food Bioprocess Technol. 2015, 8 (4), 854-864. 11. Jongberg, S.; Terkelsen, L. d. S.; Miklos, R.; Lund, M. N., Green tea extract impairs meat emulsion properties by disturbing protein disulfide cross-linking. Meat Sci. 2015, 100, 2-9. 12. Jongberg, S.; Skov, S. H.; Tørngren, M. A.; Skibsted, L. H.; Lund, M. N., Effect of white grape extract and modified atmosphere packaging on lipid and protein oxidation in chill stored beef patties. Food Chem. 2011, 128 (2), 276-283. 13. Fujimoto, A.; Masuda, T., Chemical Interaction between Polyphenols and a Cysteinyl Thiol under Radical Oxidation Conditions. J. Agr. Food Chem. 2012, 60 (20), 5142-5151. 14. Jongberg, S.; Lund, M. N.; Waterhouse, A. L.; Skibsted, L. H., 4-methylcatechol inhibits protein oxidation in meat but not disulfide formation. J. Agr. Food Chem. 2011, 59 (18), 10329-35. 15. Tang, C. B.; Zhang, W. G.; Dai, C.; Li, H. X.; Xu, X. L.; Zhou, G. H., Identification and Quantification of Adducts between Oxidized Rosmarinic Acid and Thiol Compounds by UHPLC-LTQ-Orbitrap and MALDI-TOF/TOF Tandem Mass Spectrometry. J. Agr. Food Chem. 2014, 63 (3), 902-911. 16. Cao, Y.; Xiong, Y. L., Chlorogenic acid-mediated gel formation of oxidatively stressed myofibrillar protein. Food Chem. 2015, 180, 235-243. 17. Wu, M.; Xiong, Y. L.; Chen, J., Role Of Disulphide Linkages Between Protein-Coated Lipid Droplets And The Protein Matrix In The Rheological Properties Of Porcine Myofibrillar Protein–Peanut Oil Emulsion Composite Gels. 20 / 35

ACS Paragon Plus Environment

Page 20 of 35

Page 21 of 35

468 469 470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511

Journal of Agricultural and Food Chemistry

Meat Sci. 2011, 88 (3), 384–390. 18. Graham, H. N., Green tea composition, consumption, and polyphenol chemistry. Prev. Med. 1992, 21 (3), 334-50. 19. Chan, E. W. C.; Lim, Y. Y.; Chew, Y. L., Antioxidant activity of Camellia sinensis leaves and tea from a lowland plantation in Malaysia. Food Chem. 2007, 102 (4), 1214-1222. 20. Bose, M.; Lambert, J. D.; Ju, J.; Reuhl, K. R.; Shapses, S. A.; Yang, C. S., The major green tea polyphenol,(-)-epigallocatechin-3-gallate, inhibits obesity, metabolic syndrome, and fatty liver disease in high-fat–fed mice. J. Nutr. 2008, 138 (9), 1677-1683. 21. Jiang, J.; Xiong, Y. L., Role of interfacial protein membrane in oxidative stability of vegetable oil substitution emulsions applicable to nutritionally modified sausage. Meat Sci. 2015, 109, 56-65. 22. Feng, X.; Li, C.; Jia, X.; Guo, Y.; Lei, N.; Hackman, R. M.; Chen, L.; Zhou, G., Influence of sodium nitrite on protein oxidation and nitrosation of sausages subjected to processing and storage. Meat Sci. 2016, (116), 260-267. 23. Gornall, A. G.; Bardawill, C. J.; David, M. M., Determination of serum proteins by means of the biuret reaction. J. Biol. Chem. 1949, 177 (2), 751-766. 24. Oliver, C. N.; Ahn, B.-W.; Moerman, E. J.; Goldstein, S.; Stadtman, E. R., Age-related changes in oxidized proteins. J. Biol. Chem. 1987, 262 (12), 5488-5491. 25. Liu, G.; Xiong, Y.; Butterfield, D., Chemical, Physical, and Gel‐forming Properties of Oxidized Myofibrils and Whey- and Soy-protein Isolates. J. Food Sci. 2000, 65 (5), 811-818. 26. Ellman, G. L., Tissue sulfhydryl groups. Arch. Biochem. Biophys. 1959, 82 (1), 70-77. 27. Martinaud, A.; Mercier, Y.; Marinova, P.; Tassy, C.; Gatellier, P.; Renerre, M., Comparison of oxidative processes on myofibrillar proteins from beef during maturation and by different model oxidation systems. J. Agr. Food Chem. 1997, 45 (7), 2481-2487. 28. Chen, L.; Feng, X. C.; Zhang, Y. Y.; Liu, X. B.; Zhang, W. G.; Li, C. B.; Ullah, N.; Xu, X. L.; Zhou, G. H., Effects of ultrasonic processing on caspase-3, calpain expression and myofibrillar structure of chicken during post-mortem ageing. Food Chem. 2015, 17 (7), 280-287. 29. Martinez-Alvarenga, M.; Martinez-Rodriguez, E.; Garcia-Amezquita, L.; Olivas, G.; Zamudio-Flores, P.; Acosta-Muniz, C.; Sepulveda, D., Effect of Maillard reaction conditions on the degree of glycation and functional properties of whey protein isolate–Maltodextrin conjugates. Food Hydrocolloid. 2014, 38, 110-118. 30. Cao, Y.; Xiong, Y. L., Chlorogenic acid-mediated gel formation of oxidatively stressed myofibrillar protein. Food Chem. 2015, 180, 235-243. 31. Zhou, F.; Sun, W.; Zhao, M., Controlled Formation of Emulsion Gels Stabilized by Salted Myofibrillar Protein under Malondialdehyde (MDA)-Induced Oxidative Stress. J. Agr. Food Chem.2015, 63 (14), 3766–3777. 32. Sun, J.; Wu, Z.; Xu, X.; Li, P., Effect of peanut protein isolate on functional properties of chicken salt-soluble proteins from breast and thigh muscles during heat-induced gelation. Meat Sci. 2012, 91 (1), 88-92. 33. Li, C.; Xiong, Y. L.; Chen, J., Protein Oxidation at Different Salt Concentrations Affects the Cross‐Linking and Gelation of Pork Myofibrillar Protein Catalyzed by Microbial Transglutaminase. J. Food Sci. 2013, 78 (6), C823-C831. 34. Strauss, G.; Gibson, S. M., Plant phenolics as cross-linkers of gelatin gels and gelatin-based coacervates for use as food ingredients. Food Hydrocolloid. 2004, 18 (1), 81-89. 35. Wu, J.; Shi, M.; Li, W.; Zhao, L.; Wang, Z.; Yan, X.; Norde, W.; Li, Y., Pickering emulsions stabilized by whey protein nanoparticles prepared by thermal cross-linking. Colloid. Surface. B 2015, 127, 96-104. 36. Jimenez-Colmenero, F.; Cofrades, S.; Herrero, A. M.; Solas, M. T.; Ruiz-Capillas, C., Konjac gel for use as potential fat analogue for healthier meat product development: Effect of chilled and frozen storage. Food Hydrocolloid. 2013, 30 (1), 351-357. 37. Wang, H.; Pato, M.; Pietrasik, Z.; Shand, P., Biochemical and physicochemical properties of thermally treated 21 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

512 513 514 515 516 517 518 519 520 521 522

natural actomyosin extracted from normal and PSE pork Longissimus muscle. Food Chem. 2009, 113 (1), 21-27. 38. Offer, G.; Trinick, J., On the mechanism of water holding in meat: The swelling and shrinking of myofibrils. Meat Sci. 1983, 8 (8), 245-81. 39. Li, C.; Xiong, Y. L.; Chen, J., Protein Oxidation at Different Salt Concentrations Affects the Cross-Linking and Gelation of Pork Myofibrillar Protein Catalyzed by Microbial Transglutaminase. J. Food Sci. 2013, 78 (6), C823–C831. 40. Moreno, H. M.; Bargiela, V.; Tovar, C. A.; Cando, D.; Borderias, A. J.; Herranz, B., High pressure applied to frozen flying fish ( Parexocoetus brachyterus ) surimi: Effect on physicochemical and rheological properties of gels. Food Hydrocolloid. 2015, 48 (4), 127-134. 41. Gibis, M.; Schuh, V.; Weiss, J., Effects of carboxymethyl cellulose (CMC) and microcrystalline cellulose (MCC) as fat replacers on the microstructure and sensory characteristics of fried beef patties. Food Hydrocolloid. 2015, 45, 236-246.

22 / 35

ACS Paragon Plus Environment

Page 22 of 35

Page 23 of 35

Journal of Agricultural and Food Chemistry

523

Figure and Table Captions:

524

Figure 1. Effect of EGCG on the physicochemical and structural characteristics of MP at

525

different NaCl concentrations. EGCG concentrations were 0, 100 and 1000 ppm. Salt

526

concentrations were 0, 0.2 and 0.6 M. Control represents unoxidized MP. Values are mean ± SD.

527

Different lowercase letters indicate significant differences (P < 0.05).

528

Figure 2. Representative DSC thermal curves of MP treated with EGCG at different NaCl

529

concentrations. EGCG concentrations were 0, 100 and 1000 ppm. Salt concentrations were 0, 0.2

530

and 0.6 M. Control represents unoxidized MP.

531

Figure 3. Images after SDS–PAGE of pork Myofibrillar Protein (MP) treated with EGCG and/or

532

NaCl. Native (A, - βME) and reducing (B, + βME) SDS-PAGE conditions. EGCG concentrations

533

were 0, 100 and 1000 ppm. Salt concentrations were 0, 0.2 and 0.6 M. Control represents

534

unoxidized MP.

535

Figure 4. Proposed reactions of EGCG quinone derivatives with MP 8, 15.

536

Figure 5. Effect of EGCG addition on cooking loss (A) and gel strength (B) of MP emulsion gels

537

at different NaCl concentrations. EGCG concentrations were 0, 100 and 1000 ppm. Salt

538

concentrations were 0, 0.2 and 0.6 M. Control represents unoxidized MP. Values are mean ± SD.

539

Different lowercase letters indicate significant differences (P < 0.05).

540

Figure 6. Storage modulus (G′) of MP emulsion gels prepared with A, 0 M NaCl; B, 0.2 M NaCl;

541

C, 0.6 M NaCl and different dosages of EGCG (0, 100 or 1000 ppm). Control represents

542

unoxidized MP.

543

Figure 7. Representative CLSM images of MP emulsion gels prepared with EGCG (0, 100 and

544

1000 ppm) and NaCl (0, 0.2, and 0.6 M). Control represents unoxidized MP. 23 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

545

Figure 8. SEM images of the microstructure of MP emulsion gels prepared with EGCG (0, 100

546

and 1000 ppm) and NaCl (0, 0.2, and 0.6 M). Control represents unoxidized MP.

547

Figure 9. Schematic representation of the proposed interactions of oil droplets and MP in

548

emulsion gels prepared with EGCG and different NaCl concentrations, emphasizing the

549

importance of protein disulfides and hydrophobic forces for texture.

550

Table 1. Significance values for the main effects of EGCG, NaCl, and their interaction on protein

551

chemical and structural markers, cooking loss and gel strength in MP Emulsion Gels.

552

24 / 35

ACS Paragon Plus Environment

Page 24 of 35

Page 25 of 35

Journal of Agricultural and Food Chemistry

90

f

A

Free amine (nM/mg protein)

Carbonyl (nM/mg protein)

2.0

e 1.5

b cg

1.0

bc

g d

d

h a

0.5

0.0

B

a

80 b

b

70

d

60

c

c 50

c

c

e

40 30

f

20 10 0

0MNaCl

0.2MNaCl

Control

0ppm

0.6MNaCl

100ppm

0MNaCl

1000ppm

0.2MNaCl

Control

0ppm

0.6MNaCl

100ppm

1000ppm

553

50

1500

C

a

d b

b

D b

Surface hydrophobicity

Thiol groups (nM/mg protein)

60

b

b

f

40 30 c e

20

g

1000

ad

d ae

e

e

500

c

c

g

10 0

0 0MNaCl Control

0.2MNaCl 0ppm

100ppm

0.6MNaCl

0MNaCl

1000ppm

Control

554 555

b

Figure 1 25 / 35

ACS Paragon Plus Environment

0.2MNaCl 0ppm

100ppm

0.6MNaCl 1000ppm

Journal of Agricultural and Food Chemistry

556 557

Figure 2 26 / 35

ACS Paragon Plus Environment

Page 26 of 35

Page 27 of 35

Journal of Agricultural and Food Chemistry

558

559 560

Figure 3

27 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

561 562

Figure 4

563 564 565 566 567 568 569 570 571 572 573 574

28 / 35

ACS Paragon Plus Environment

Page 28 of 35

Page 29 of 35

Journal of Agricultural and Food Chemistry

A

2.16 folds

1.88 folds 50%

9.46 folds

d

Cooking loss (%)

cd

c

40%

30%

b a

b

a

a

20%

10% e

f

0% 0M NaCl

0.2M NaCl

Control

575

0ppm

100ppm

0.6M NaCl 1000ppm

0.7

B

c

0.6

Gel strength (N)

a 0.5

a

a

a

a

a

a

0.4 b

b

0.3 0.2 0.1 0.0 0M NaCl Control

0.2M NaCl 0ppm

100ppm

0.6M NaCl 1000ppm

576 577

Figure 5

29 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

578

579 580

Figure 6 30 / 35

ACS Paragon Plus Environment

Page 30 of 35

Page 31 of 35

Journal of Agricultural and Food Chemistry

581 582

Figure 7 31 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

583 584

Figure 8

32 / 35

ACS Paragon Plus Environment

Page 32 of 35

Page 33 of 35

Journal of Agricultural and Food Chemistry

585 586

Figure 9

33 / 35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

587

Table 1 carbonyls thiol groups free amines Surface Hydrophobicity Cooking loss Gel strength EGCG