Effect of Sodium Chloride and Sodium Bicarbonate on the

Jun 7, 2018 - Soft wheat flour doughs were prepared with different levels of salt (NaCl) or baking soda (NaHCO3). Oscillation rheology, elongational v...
0 downloads 0 Views 3MB Size
Subscriber access provided by Kaohsiung Medical University

Food and Beverage Chemistry/Biochemistry

Effect of sodium chloride and sodium bicarbonate on the physicochemical properties of soft wheat flour doughs and gluten polymerization Gengjun Chen, Laura Ehmke, Rebecca Miller, Pierre Faa, Gordon Smith, and Yonghui Li J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b01197 • Publication Date (Web): 07 Jun 2018 Downloaded from http://pubs.acs.org on June 8, 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 33

Journal of Agricultural and Food Chemistry

1

Effect of sodium chloride and sodium bicarbonate on the

2

physicochemical properties of soft wheat flour doughs and gluten

3

polymerization

4 5 6

Gengjun Chen1, Laura Ehmke1, Rebecca Miller1, Pierre Faa2, Gordon Smith1,

7

Yonghui Li1*

8 9 10

1

11

66506

12

2

13 14 15 16 17 18 19 20 21 22

Department of Grain Science and Industry, Kansas State University, Manhattan, KS

Frito-Lay North America, Plano, TX 75024

*Correspondence to: Yonghui Li, E-mail: [email protected], Ph: 785-532-4061, fax: 785-532-7010

Submit to Journal of Agricultural and Food Chemistry

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

23

Abstract: Soft wheat flour doughs were prepared with different levels of salt (NaCl)

24

and/or baking soda (NaHCO3). Oscillation rheology, elongational viscosity, and

25

extensibility of doughs were tested to evaluate the effect of salt and/or baking soda

26

on the physical properties of doughs. Furthermore, a series of physical-biochemical

27

analytical techniques were used to investigate gluten polymerization in doughs,

28

including Zeta potential analyzer, Fourier transform infrared spectroscopy (FTIR),

29

spectrophotometer, and reversed phase high performance liquid chromatography

30

(RP-HPLC). The addition of high levels of NaHCO3 (1.0 % fwb), either by itself or in

31

combination with NaCl, increased dough strength, elongational viscosity, and

32

viscoelasticity. RP-HPLC results demonstrated macromolecular aggregation of gluten

33

proteins in the presence of NaCl and/or NaHCO3. The addition of NaHCO3 or NaCl

34

also decreased both free sulfhydryl content and random coil structure of gluten

35

isolated from the doughs. Overall, NaCl and/or NaHCO3 induced the changes of

36

molecular conformation of gluten, which impacted the physicochemical qualities of

37

soft wheat flour dough. This study provides a better understanding of salt and baking

38

soda functionality in the formation of soft flour dough, which will support the

39

searching of feasible sodium reduction strategies in soft flour bakery products.

40 41 42

Keywords: sodium chloride, sodium bicarbonate, soft wheat flour doughs, gluten,

43

gliadin, glutenin, secondary structure, hydrophobicity, molecular interaction

44 45 46

2

ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33

Journal of Agricultural and Food Chemistry

47

Introduction

48

Although worldwide dietary patterns are different, wheat-based products are still the

49

major food source for most of the population. Soft wheat, accounting for about 15%

50

of total US wheat production, is widely used in various chemically leavened products

51

such as cookies, pretzels, cakes, crackers, etc. Baking soda (or sodium bicarbonate,

52

NaHCO3) is by far the most popular chemical leavening agent and offers many

53

advantages, such as versatility, easy of handling, nontoxicity, tastelessness and low

54

cost. The addition of NaHCO3 in bakery goods is typically around 0.5-2% (flour basis),

55

depending on the specific type of food products. Salt (or sodium chloride, NaCl) is

56

another indispensable ingredient for nearly all bakery products and cereal snacks. It

57

provides not only crucial sensory characteristics to food but also technological

58

functions that influence the processing and quality characteristics of dough and

59

cereal goods. Previous studies showed that salt increases dough resistance, elasticity,

60

extensibility, and mixing stability.1,2 Salt also contributes to the products coloring,

61

improves food texture, and inhibits microbial growth. The typical addition of salt in

62

bakery products is about 1-2% total flour weight basis (fwb). Despite the fact that

63

sodium is an essential nutrient for humans, it is only needed by the body in relatively

64

small quantities. Excessive sodium intake significantly increases blood pressure,3,4

65

which has been the major cause of cardiovascular diseases (CVDs) such as strokes,

66

heart failure and coronary heart disease (CHD). CVDs alone account for one third of

67

global deaths.5 Currently, Americans consume an average of 3,440 mg sodium per

68

day, which is 50% more than the upper limit of 2,300 mg per day recommended by

69

the 2015 Dietary Guidelines for Americans,6 thus sodium intake should be

70

significantly reduced. However, as discussed above, salt and baking soda are still the

71

two largest dietary sodium sources in baked products. For example, these products

72

contribute to 30% of our daily salt intake.7,8 Therefore, reducing salt and baking soda

73

in bakery products is an essential step for the overall sodium reduction goal.

74

Wheat doughs possess unique viscoelastic properties attributed to the gluten

75

proteins. Network formation of wheat gluten is essential for many cereal and bakery

76

products. When the flour is mixed with water and other ingredients, wheat protein is 3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

77

hydrated, interacts with other additives such as salt and/or baking soda, and

78

develops a unique polymeric network, which is associated with dough properties.

79

According to different solubility, wheat flour proteins are traditionally classified as

80

albumins (soluble in water), globulins (soluble in NaCl solution), prolamins (soluble in

81

ethanol solution), and glutelins (soluble in acid or alkali solution), which are also

82

defined as Osborne fractions.9,10 More than 80% of gluten consist of gliadins

83

(prolamins fraction) and glutenins (glutelins fraction).11 The gliadins act as a

84

plasticizer to influence the viscous properties; while the glutenins determine the

85

elasticity and cohesiveness of the dough.12,13 Goesaert et al.14 reported the gliadins

86

are noncross-linked proteins with molecular mass (Mr) of 30,000-75,000 (ω-, α- and

87

γ-type). The glutenins are divided into the predominant low molecular weight

88

glutenin subunits (LMW-GS, Mr of 40,000-70,000), and high molecular weight

89

glutenin subunits (HMW-GS, Mr of 60,000-90,000).15

90

Salt and baking soda provide many technological functions in bakery food

91

systems in addition to the saltiness and leavening action. It is still a major challenge

92

to remove them or reduce their amount without any negative effect on dough

93

rheological properties or product quality. One major reason is that there is no clear

94

understanding of how salt and/or baking soda contribute to dough properties,

95

especially the interactions with gluten. Although some research studies have been

96

carried out during the past,15-17 the fundamental basis of salt and baking soda in

97

dough systems have yet to be well understood. The lack of foundational knowledge

98

in the field has been a limiting factor in developing effective strategies to reduce

99

sodium, and an in-depth knowledge of the function of salt and/or baking soda in

100

doughs at a molecular level is crucial. This study was conducted to investigate the

101

influences of salt and/or baking soda on the rheological properties and physical

102

processibility of soft wheat flour dough and gluten macromolecular characteristics

103

and interaction during dough development.

104 105

Materials and Methods

106

Materials 4

ACS Paragon Plus Environment

Page 4 of 33

Page 5 of 33

Journal of Agricultural and Food Chemistry

107

All chemicals, solvents, and reagents used in the experiments were of analytical

108

grade, and were purchased from Fisher Scientific (Fairlawn, NJ, USA) or Sigma-Aldrich

109

(St. Louis, MO, USA), unless specified otherwise. Soft wheat flour (pretzel flour,

110

protein content 8.0%, moisture content 12.7%, fwb) was provide by Frito-Lay North

111

America.

112

Mixing properties

113

Dough samples were prepared with different levels of NaCl and/or NaHCO3 (%

114

fwb) are shown in Table 1. The treatment includes four levels of NaCl (0, 0.5, 0.7, and

115

1.0%, fwb) alone, three levels of NaHCO3 (0, 0.3, and 1.0%) alone, and four

116

combinations of NaCl and NaHCO3. Mixing properties were measured using

117

Mixograph following AACCI Approved Method 54-40.02.18 The optimum water

118

absorption of the control flour was determined and that absorption was used for all

119

treatments. All analyses were carried out in triplicate. Farinograph was tested and

120

developed for each treatment according to AACCI Approved Method 54-21.01,18 and

121

mixing parameters including water absorption, mix time, and dough stability were

122

obtained.

123

Dough extensibility

124

Dough extensibility and resistance to extension were determined via Kieffer test

125

using a TA-XTPlus texture analyzer (Stable Micro Systems, Godalming, UK) with

126

SMS/Kieffer rig (Stable Micro Systems). Doughs containing 100 g flour, 59 g water,

127

and predetermined amount of NaCl and/or NaHCO3 was mixed in a pin mixer with

128

the mixing time determined from mixograph as shown in Table 1. The prepared

129

dough was equilibrated for 30 min in a closed container before the measurement,

130

and then it was gently molded into a rectangle with minimal handling and set on the

131

grooved section of the Teflon former lubricated with the mineral oil. A lametta strip

132

was placed in each groove to aid the removal of dough strip. The cover block was

133

placed on top of the former and a clamp compressed the blocks together. Excess

134

dough forced out the sides of the former was trimmed off. The full-length dough

135

strips were selected and tested after resting for 30 mins. The following settings were

136

used: pre-test speed of 2 mm/s, test speed of 3.3 mm/s, post-test speed of 10 mm/s, 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

137

a distance of 75 mm and trigger force of 5 g. The dough was molded and rested in a

138

clamping tool for 30 min, and full-length dough strips were tested. The output

139

quantities were the resistance to extension (g) and the extensibility (mm). The

140

analysis was performed in triplicate.

141

Dynamic oscillatory rheology

142

A Bohlin CVOR 150 rheometer (Malvern Instruments, Southborough, MA)

143

equipped with a PP 20 parallel plate (gap size of 1200 μm) was utilized to evaluate

144

the oscillatory rheological properties of dough. Doughs were freshly prepared in the

145

pin mixer as described previously, and the dough was rested for 30 min in a closed

146

container before loading on the dynamic oscillatory rheometer. Excessive dough

147

sample outside the plate edge was trimmed and the paraffin oil was used on the

148

lateral surface to prevent moisture evaporating or drying. Amplitude scan was first

149

performed at strain values of 0.01-10% at a constant frequency of 1Hz at 25 °C to

150

establish the linear viscoelastic range. Frequency sweep was then performed from

151

0.1 to 100 Hz within the strain amplitude set at 0.1% (within linear viscoelasticity) at

152

25 °C. Storage modulus (G’), loss modulus (G”) and tan δ (G”/G’) as a function of

153

amplitude or frequency were recorded. Each treatment was analyzed in duplicate.

154

Elongational viscosity property

155

The TA.XTPlus (Texture Technologies, Scarsdale, NY, USA; and Stable Micro

156

Systems, Godalming, UK) was used to measure the elongational viscosity of the

157

doughs. Doughs were prepared similarly as previously described. The doughs were

158

sheeted, covered and rested for 10 min. Three 1 inch diameter disks were cut from

159

the dough and covered with mineral oil for lubricating and to prevent adhesion to

160

the probe. Lubricated uniaxial compression was performed with a 2 inch diameter

161

probe at 50% strain deformation at a speed of 0.4 mm/s. Three separate doughs with

162

3 disks per dough were tested for each treatment.

163

Extraction of gluten

164

Gluten fractions were extracted from soft wheat flour dough systems prepared

165

with the different levels of NaCl and NaHCO3 following AACCI Approved Method

166

38-10.01.18 Gluten samples were lyophilized, ground to powder, and stored in -20 °C 6

ACS Paragon Plus Environment

Page 6 of 33

Page 7 of 33

Journal of Agricultural and Food Chemistry

167

freezer until further analysis. Protein content was determined using a LECO TruMac

168

nitrogen analyzer (LECO Corp., St Joseph, MI, USA) according to AACCI Approved

169

Method 46-19.01,18 and a conversion factor of 5.7 was utilized for calculation from

170

nitrogen content.

171

RP-HPLC analysis

172

RP-HPLC measurement of gluten extracts was conducted according to a previous

173

method with some modification.19 Gluten (100 mg) was extracted stepwise twice

174

with 1.0 mL of 60% (v/v) aqueous ethanol for 12 min at room temperature to obtain

175

the gliadin fraction; and twice with 1.0 mL of a solution containing 50% (v/v)

176

propan-1-ol, 2 M urea, 1% (w/v) dithioerythritol and 0.05 M Tris-HCl for 45 min at

177

60 °C to obtain the glutenin fraction. Each extraction step started with vortex mixing

178

for 1 min at room temperature. All extracts were centrifuged (15,000 xg, 10 min,

179

20 °C), and corresponding extracts were combined and filtered with a 0.45 μm

180

membrane (Phenex™ filter membranes, Phenomenex, Torrance, CA, USA). A HP1050

181

Series HPLC (Agilent Technologies, Santa Clara, CA, USA) coupled with a diode array

182

detector (DAD) was used to analyze the extracts. Two fractions of extracts were

183

loaded (10 μL) onto a reversed-phase Aeris WIDEPORE XB-C18 column (3.6 μm, 150 x

184

4.6 mm, Phenomenex, Torrance, CA, USA). The elution solvents were water

185

containing 0.1% trifluoroacetic acid (A), and acetonitrile containing 0.1%

186

trifluoroacetic acid (B) with a linear gradient from 0 min 24 % B to 30 min 56 % B,

187

with flow rate of 0.6 mL/min, and temperature of 60 °C with detection at 210 nm.

188

The gliadins and glutenins were identified by comparison of their retention times and

189

spectra with protein standards while quantitative analysis was based on their peak

190

areas from the chromatograms. Extraction and measurements were performed in

191

duplicate.

192

Zeta potential

193

Zeta potential of samples was tested on a ZetaPALS Zeta potential analyzer

194

(Brookhaven Instruments Co., Holtsville, NY, USA) with hydrodynamic light scattering

195

and laser doppler electrophoresis. Zeta potential value was determined by the

196

method described by Chen et al.20 Gluten suspensions were prepared with a solid 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 33

197

content of 0.1%. Five runs comprising of 2 cycles were carried out for each

198

measurement. Each sample was tested in duplicate.

199

Sulfhydryl (SH) and disulfide (SS) group determination

200

The concentration of free SH was determined by Ellman’s reagent

201

(5’5-dithiobis(2-nirtobenzoic acid), DTNB) according to the method reported by

202

Rombouts et al.21 with some modifications. Briefly, gluten (30 mg) was dispersed in

203

3.0 mL of reaction buffer A (8 M urea, 3 mM EDTA, 1% SDS, 0.2 M Tris-HCl, and pH

204

8.0), and sample was vortexed for 30 s and shaken at 25 °C for 60 min. Afterwards,

205

0.3 mL of buffer B (10 mM DTNB in 0.2 M Tris-HCl, pH 8.0) was added to the sample

206

and mixed for another 60 min. The solution was then centrifuged at 13,600 xg for 15

207

min at room temperature, and the absorbance of the supernatant was detected at

208

412 nm using a double beam spectrophotometer (VWR UV-6300PC, Radnor, PA,

209

USA).

210

The concentration of total SH group was determined according to a previous

211

methods with some modification.22,23 Briefly, 1 mL of reaction buffer (3mM EDTA, 1%

212

SDS,

213

2-nitro-5-thiosulfobenzoate (NTSB)) was mixed with 10 mg of gluten . The solution

214

was vortexed for 30 s and mixed in a shaker for 60 min in the dark, and then

215

centrifuged at 13,600 xg for 15 min followed by diluting the supernatant (0.3 mL)

216

with 2.7 mL reaction buffer without NTSB. Absorbance was then measured at 412 nm

217

using a double beam spectrophotometer (VWR UV-6300PC, Radnor, PA, USA). The SH

218

content (CSH) was calculated according to the equation: CSH =

219

absorbance, ε is the extinction coefficient of 13,600 M-1cm-1, and b is the cell path

220

length. The SS content (Css) was calculated according to the equation: Css =

221 222

0.2M

  

Tris-HCl,

0.1M

sodium

sulfite,

pH

9.5,

 

and

0.5mM

, where A is the

, where CTSH is the total SH content, and CFSH is the free SH content.

Secondary structure of gluten

223

Secondary structures of gluten were analyzed using a PerkinElmer Spectrum 400

224

FT-IR/FT-NIR Spectrometer (PerkinElmer, Inc., Waltham, MA, USA) equipped with an

225

ATR accessory. A total of 64 scans were taken at a resolution of 4 cm-1 in the range of 8

ACS Paragon Plus Environment

Page 9 of 33

Journal of Agricultural and Food Chemistry

226

400-4000 cm-1. Gluten secondary structures were deconvoluted from the amide I

227

region (1600-1700 cm-1) via a secondary derivation. The spectra were analyzed using

228

OriginPro 2016 software to obtain the relative areas of the selected amide I region.

229

Surface hydrophobicity

230

The sodium dodecyl sulfate (SDS) binding capacity was used to measure protein

231

total surface hydrophobicity according to a previous method with some

232

modification.24 The SDS solution (0.1 mmol/L, 40 mL) was used to dissolve 10 mg

233

gluten , and was shaken for 60 min. After being dialyzed in distilled water for 48 h,

234

one milligram of inner dialyzate was mixed with 20 mL CHCl3, and then 5 mL of

235

methylene blue (0.024 g/L) was added into the CHCl3 layer. The mixture was

236

centrifuged at 2500 xg for 15 min, and then the absorbance of the mixture in the

237

lower layer was detected at 655 nm (VWR UV-6300PC, Radnor, PA, USA). A linearity

238

(R2 > 0.995) for the SDS standard solution within the range of 0.01–0.1 mmol/L was

239

obtained in the calibration data. The hydrophobicity value (H) = × 20 ×

240

where c is the concentration of SDS calculated by absorbance and standard curve.

241

Statistical analysis

. 

,

242

Data were analyzed using the Statistical Analysis System software v9.4 (SAS

243

Institute, Cary, NC, USA). Significant differences were found at the statistical presence

244

of a P value of 0.05.

245 246

Results and Discussion

247

Mixing properties of doughs

248

In the Mixograph study, water absorption was held constant at 59% (optimal for

249

the control) to investigate the effect of the NaCl and NaHCO3 on dough mixing

250

properties (Table 1, Figure 1). Mixing time of the control doughs was 3.50 min, and it

251

increased to different extents for other treatment groups (Table 1). Addition of

252

NaHCO3 and/or NaCl increased the mixing time compared to the control; while the

253

combination of 1.0% NaCl plus 1.0% NaHCO3, or 1.0% NaHCO3 alone contributed the

254

most to the increased mixing time (6.2 or 6.0 min). Moreover, NaCl had relatively less

9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

255

effect on mixing time than NaHCO3 when each ingredient was added individually.

256

Farinograph absorption slightly decreased with the addition of NaHCO3 or in the

257

combination groups (Table 1). Salt or baking soda influenced the mixing time and

258

stability of all treatment groups. The addition of 1% NaHCO3 exhibited the largest

259

strengthening effect, increasing the mixing time by 4 min and the stability by 9.2 min

260

compared with the control. Dough stability was affected more by NaHCO3 than NaCl;

261

for instance, the Farinogram curves of doughs containing 1.0% NaHCO3 or 0.5%

262

NaCl/0.3% NaHCO3 became much stronger; whereas 0.7% NaCl did not have a large

263

impact on the shape of the Farinogram curve (Figure 1). The results were consistent

264

with our finding from Mixograph experiments.

265

Extensibility and viscosity of doughs

266

A previous study suggested that dough strain hardening behavior was correlated

267

with the gluten network.25 Although 0.5% NaCl/0.3% NaHCO3, or 0.7% NaCl had

268

relatively minor effects on the dough rupture forces, dough strength was significantly

269

increased (P < 0.05) in the other treatment groups compared with the control (7.4 g)

270

(Table 2). The rupture force of dough with 0.5% NaCl/1.0% NaHCO3 was the largest

271

(20.3 g). Increased dough extensibility was achieved with salt alone at all the studied

272

levels, as well as with lower amount of NaHCO3 (0.3%), or combination of salt with

273

0.3% NaHCO3. As shown in Table 2, the addition of 1.0% NaHCO3 produced doughs

274

with the highest elongational viscosity (P > 0.05); while those groups with the

275

combination of NaCl/NaHCO3 also showed increased values compared to control

276

group. Furthermore, NaCl alone did not change dough viscosity even when the level

277

was 1.0% fwb. Overall, the presence of NaHCO3 in soft wheat flour doughs greatly

278

enhanced the strength and viscosity of dough, and it may be due to the stronger

279

polymeric interaction of gluten in these systems, which will be further discussed in

280

the next sections.

281

Dynamic rheometry of doughs

282

Oscillational rheometry measurement was carried out to investigate the

283

rheological properties of the dough. Strain and frequency sweeps curves are shown

284

in Figure 2. Loss modulus (G”), storage modulus (G’), and tan δ (G”/G’) of the 10

ACS Paragon Plus Environment

Page 10 of 33

Page 11 of 33

Journal of Agricultural and Food Chemistry

285

samples are the basic parameters used to describe the viscoelastic behavior of the

286

dough systems. Although the doughs treated with NaCl at high levels (1.0% of fwb) or

287

NaHCO3 all exhibited higher G´ and G” compared with the control, the difference of

288

the modulus between NaHCO3 containing doughs and control was much more

289

significant. It is noticeable that the combined addition of salt and baking soda

290

markedly changed the G” and G’ of doughs. These results were in accordance with

291

the previous reports.1,25,26 For instance, it was reported that adding more salt

292

enhanced the strength and elasticity of wheat flour dough.25 The tan δ (G’’/G’) is

293

referred to as a proper measure of viscoelasticity, and Yiannopoulos et al.27

294

suggested that wheat dough can be classified as a viscoelastic solid if the tan δ was

295

less than one. In our study, the tan δ values were decreased with the addition of

296

NaHCO3 or combinations among the groups at 0.1% strain (Figure 2E), which

297

indicated that the dough structure became relatively more elastic in the presence of

298

NaHCO3. A previous study discovered alkali enhanced dough strength;28 in the

299

meantime, the elasticity of buckwheat dough could be increased with the addition of

300

alkali.29 The reason salt or baking soda impacted the rheological properties of doughs

301

may be attributed to the structural changes of macromolecules (i.e. cross-links of

302

gluten polymers). However, other authors did not find any change in tan δ with NaCl

303

addition.30,31 The lack of consistency on the rheology tests might be due to the

304

different wheat gluten protein quality and quantity or dough systems.

305

Gluten fractions analysis

306

After extracting gluten from soft wheat flour dough systems with different levels

307

of NaCl and NaHCO3, the fractions of gliadin and glutenin were separated and

308

identified by RP-HPLC (Figure 3), and the results are presented in Table 3. The

309

linearity of the standard calibration curve ranged from 0.065 to 1.05 mg, in which

310

the correlation coefficient indicated a good linearity (R2 > 0.98). The limit of

311

detection (LOD) was 0.016 mg/mL, and the limit of quantification (LOQ) was 0.053

312

mg/mL. According to results of recovery experiments, the average recovery rate was

313

88%, which was calculated by spiking the control samples with 1% gliadins standard

314

(n=3); while the coefficient of variation for repeatability of determinations of 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

315

samples was ±5.3% (average of five determinations within two weeks), which

316

compared favorably with the previous study.19 The gluten samples were extracted

317

step wise and further analyzed through RP-HPLC into several elution ranges including

318

gliadin fraction (ω-, α-, and γ-gliadin), and glutenin fraction of bound gliadin

319

(ωb-gliadin), HMW-GS, and LMW-GS, respectively.13,32

320

Although ω-gliadin amount remained almost constant in all the groups, the

321

extractable α- and γ-gliadin decreased gradually with the addition of NaCl and/or

322

NaHCO3; therefore, the overall extractable gliadins also decreased. In the meantime,

323

the apparent HMW-GS and LMW-GS amounts both increased when adding NaHCO3

324

into the dough system. For instance, the HMW-GS extractability concentration was

325

increased significantly from 30 to 36 μg/mg in the group within NaCl 1.0%/NaHCO3

326

1.0% (P < 0.05). The total glutenin content was increased by up to 19% with the

327

addition of NaCl 1.0%/NaHCO3 1.0%. The present results suggested that

328

gliadin-glutenin polymeric interaction and cross-linking were enhanced via the

329

addition of salt and/or baking soda, which led to a reduced amount of extractable

330

gliadins and increased amount of extractable glutenins. NaHCO3 had a stronger

331

effect than NaCl, which was consistent with Wu et al.33 who also reported the

332

alkaline reagent impact association of gluten. Lagrain et al.34 pointed out that the

333

aggregation of gluten might be through oxidation of SH groups or SH-SS exchange

334

reactions in bread dough systems. The amount of extractable ω-gliadin was not

335

affected by the addition of salt and NaHCO3, because the ω-gliadin lack cysteine

336

groups involved in the interaction with other proteins.19 SS bonds, hydrophobic, or

337

electrostatic interactions might change the microstructure of gluten protein as

338

well.33,35,36 The investigation of the possible interaction of gluten will be discussed in

339

the next sections.

340

The ratio between gliadins to glutenins (gli/glu) influences the formation of

341

dough since the gliadins are responsible for the viscosity and extensibility of dough,

342

whereas glutenins play a key role in the elasticity of dough through polymeric

343

network formation.2,35 An appropriate ratio of gli/glu is required for the viscoelastic

344

properties of doughs and the final bakery goods quality.13,37 It is noteworthy that the 12

ACS Paragon Plus Environment

Page 12 of 33

Page 13 of 33

Journal of Agricultural and Food Chemistry

345

ratio significantly decreased with high levels of NaHCO3 or combined addition (P
0.05). A

394

significant change of β-turn structure was not observed (P > 0.05). However, it was

395

found that the formation of β-sheet was at the expense of β-turn during gluten

396

deformation in another study.42,43 An increase of β-sheet structure in the previous

397

studies was attributed to the effect of hydrogen bonding between the glutenin

398

molecules.16,42 Based on a proposed mechanism,9 the hydrogen bonds contribute to

399

the formation of more loop structure of the gluten when the molecules are hydrated,

400

and consequently the gluten aggregation are more pronounced. Therefore, NaCl

401

and/or NaHCO3 may influence the secondary conformation of gluten via inducing

402

hydrogen bonding formation and contribute to the development of gluten network

403

appearing as protein aggregation in the study.

404

Surface hydrophobicity 14

ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33

Journal of Agricultural and Food Chemistry

405

Gluten is a biomacromolecule with a 3-dimensional structure, in which

406

hydrophobic amino acid residues contribute to the tertiary conformation through

407

hydrophobic interaction. In the study, SDS binding was used to explain the degree of

408

hydrophobicity of gluten samples from soft wheat flour dough systems. This method

409

has been applied to determine the hydrophobicity of the insoluble proteins such as

410

ovalbumins.24 No significant differences of the degree of hydrophobicity of glutens as

411

affected by NaCl and/or NaHCO3 were found (P > 0.05); thus, the addition of salt or

412

baking soda at present levels does not influence the total surface hydrophobicity of

413

gluten in the current dough systems.

414

In the present study, the ratio of extractable gliadins to glutenins significantly

415

decreased within the high level of baking soda, implying that more aggregated and

416

less soluble protein network was formed. On other hand, the absolute values of zeta

417

potential were changed, which, might confirm the molecular interactions and

418

aggregations. In terms of the FTIR results, the secondary structure distributions were

419

altered, where both salt and baking soda could contribute to the reduced amount of

420

random coil structure. To further elucidate the mechanism of polymeric network

421

formation in dough systems, it was reported as a result of SS bonds formation by

422

oxidation of SH or interchange of SH-SS,29,44 which was also partially demonstrated

423

by our HPLC results of the aggregation of gliadins. The free SH content of gluten

424

reduced in the doughs treated with 1.0% NaHCO3 or combination, which suggested

425

that NaHCO3 may lead to more intense SH reactions during dough production.

426

However, there are other non-redox cross-linking worth studying, which may cause

427

the polymeric interaction of dough.

428

The inclusion of NaCl and NaHCO3 in soft flour doughs improved dough mixing

429

stability, strength, extensibility, and elasticity (to some extent), which was a result of

430

synergistic inter- and intra-molecular interactions. There are several conceptual

431

models proposed to explain how the interaction of ingredients affected dough

432

rheological behavior. It was believed that the hydration of gluten results in the

433

cohesion and adhesion properties of dough, as the mobility of hydrated gluten

434

increases in the network.16 Thus, adding NaCl or/and NaHCO3 could mitigate the 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

435

stickiness and enhance the strength of dough via the interaction of ions and water to

436

delay the hydration, which was confirmed by the longer mixing time during the

437

development of dough. Meanwhile, as a result of the interchange of SH-SS or

438

non-covalent interaction (hydrogen bonds or hydrophobic) on the molecules level,

439

the gluten polymerization could contribute to the change of dough rheological

440

behavior, which was partially consistent with our observation. However, the change

441

of the surface hydrophobic interactions was not significantly in our study. In summary,

442

the presence of NaHCO3 and/or NaCl induced the macromolecular aggregation via SH

443

related cross-linking or other non-redox reaction and altered secondary structure of

444

gluten, which resulted in the changes of physical and rheological properties of dough.

445

In order to achieve a significant amount of sodium reduction in wheat products and

446

develop high quality bakery foods, more in-depth studies are needed to address

447

protein polymerization and aggregation during the dough development.

448 449

Acknowledgement

450

This is contribution no. 18-316-J from the Kansas Agricultural Experimental Station.

451

Financial support was provided by Frito-Lay North America and the Kansas State

452

University, Department of Grain Science and Industry.

453 454

16

ACS Paragon Plus Environment

Page 16 of 33

Page 17 of 33

Journal of Agricultural and Food Chemistry

455

References

456

1. Beck, M.; Jekle, M.; Becker, T. Impact of sodium chloride on wheat flour dough for

457

yeast-leavened products. I. Rheological attributes. J. Sci. Food Agric. 2012, 92

458

(3), 585-592.

459 460

2. He, H.; Roach, R. R.; Hoseney, R. C. Effect of Nonchaotropic Salts on Flour Bread-Making Properties. Cereal Chem. 1992, 69 (4), 366-371.

461

3. Luft, F. C.; Zemel, M. B.; Sowers, J. A.; Fineberg, N. S.; Weinberger, M. H. Sodium

462

bicarbonate and sodium chloride: effects on blood pressure and electrolyte

463

homeostasis in normal and hypertensive man. J. Hypertens. 1990, 8 (7),

464

663-670.

465

4. Sacks, F. M.; Svetkey, L. P.; Vollmer, W. M.; Appel, L. J.; Bray, G. A.; Harsha, D.;

466

Obarzanek, E.; Conlin, P. R.; Miller, E. R.; Simons-Morton, D. G.; Karanja, N.;

467

Lin, P. H.; Grp, D. A.-S. C. R., Effects on blood pressure of reduced dietary

468

sodium and the dietary approaches to stop hypertension (DASH) diet. N. Engl.

469

J. Med. 2001, 344 (1), 3-10.

470

5. Roth, G. A.; Johnson, C.; Abajobir, A.; Abd-Allah, F.; Abera, S. F.; Abyu, G.; Ahmed,

471

M.; Aksut, B.; etc. Global, Regional, and National Burden of Cardiovascular

472

Diseases for 10 Causes, 1990 to 2015. J. Am. Coll. Cardiol. 2017, 70 (1), 1-25.

473

6. United States Department of Agriculture (USDA), & United Stated Department of

474

Health and Human Services (HHS). 2015-2020 Dietary Guidelines for

475

Americans. URL

476

(https://health.gov/dietaryguidelines/2015/resources/2015-2020_Dietary_G

477

uidelines.pdf) (2015).

478 479

7. Farahnaky, A.; Hill, S. E. The effect of salt, water and temperature on wheat dough rheology. J. Texture Stud. 2007, 38 (4), 499-510.

480

8. Angus, F. Dietary salt intake: sources and targets for reduction. Reducing salt in

481

foods: practical strategies. Woodhead Publishing: Witney, UK, 2007; 3-17.

482

9. Shewry, P. R. Wheat. J. Exp. Bot. 2009, 60 (6), 1537-1553.

483

10. Song, Y.; Zheng, Q. Dynamic rheological properties of wheat flour dough and

484

proteins. Trends Food Sci Technol. 2007, 18(3), 132-138. 17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

485

11. Domenek, S.; Brendel, L.; Morel, M. H.; Guilbert, S. Swelling Behavior and

486

Structural Characteristics of Wheat Gluten Polypeptide Films.

487

Biomacromolecules. 2004, 5, 1002-1008.

488

12. Joye, I. J.; Lagrain, B.; Delcour, J. A. Use of chemical redox agents and exogenous

489

enzymes to modify the protein network during breadmaking - A review. J

490

Cereal Sci. 2009, 50 (1), 11-21.

491

13. Wieser, H. Chemistry of gluten proteins. Food Microbiol, 2007, 24(2), 115-119.

492

14. Goesaert, H.; Brijs, K.; Veraverbeke, W. S.; Courtin, C. M.; Gebruers, K.; Delcour, J.

493

A. Wheat flour constituents: how they impact bread quality, and how to

494

impact their functionality. Trends Food Sci Technol, 2005, 16(1-3), 12-30.

495

15. Lindsay, M. P.; Skerritt, J. H. The glutenin macropolymer of wheat flour doughs:

496

structure-function perspectives. Trends Food Sci Technol. 1999, 10, 247-253.

497

16. Tuhumury, H. C. D.; Small, D. M.; Day, L. The effect of sodium chloride on gluten

498

network formation and rheology. J Cereal Sci. 2014, 60 (1), 229-237.

499 500 501 502 503

17. Ukai, T.; Matsumura, Y.; Urade, R. Disaggregation and reaggregation of gluten proteins by sodium chloride. J. Agric. Food. Chem. 2008, 56 (3), 1122-1130. 18. AACC. Approved methods of the American Association of Cereal Chemists; The Association: St. Paul, MN, 2000. 19. Wieser, H.; Antes, S.; Seilmeier, W. Quantitative determination of gluten protein

504

types in wheat flour by reversed-phase high-performance liquid

505

chromatography. Cereal Chem. 1998, 75(5), 644-650.

506

20. Chen, P.; Zhang, L. Interaction and properties of highly exfoliated soy

507

protein/montmorillonite nanocomposites. Biomacromolecules. 2006, 7(6),

508

1700-1706.

509

21. Rombouts, I.; Jansens, K. J. A.; Lagrain, B.; Delcour, J. A.; Zhu, K. X., The impact of

510

salt and alkali on gluten polymerization and quality of fresh wheat noodles. J

511

Cereal Sci. 2014, 60 (3), 507-513.

512

22. Chan, K. Y.; Wasserman, B. P. Direct Colorimetric assay of free thiol-groups and

513

disulfide bonds in suspensions of solubilized and particulate cereal proteins.

514

Cereal Chem. 1993, 70(2), 240-240. 18

ACS Paragon Plus Environment

Page 18 of 33

Page 19 of 33

Journal of Agricultural and Food Chemistry

515 516 517

23. Thannhauser, T. W.; Konishi, Y.; Scheraga, H. A. Analysis for disulfide bonds in peptides and proteins. Methods Enzymol. 1987, 143, 115-119. 24. Kato, A., Matsuda, T.; Matsudomi, N.; Kobayashi, K. Determination of protein

518

hydrophobicity using sodium dodecyl sulfate binding method. J. Agric. Food.

519

Chem. 1984, 32(2), 284-288.

520

25. McCann, T. H.; Day, L. Effect of sodium chloride on gluten network formation,

521

dough microstructure and rheology in relation to breadmaking. J Cereal Sci.

522

2013, 57(3), 444-452

523

26. Larsson, H. Effect of pH and sodium chloride on wheat flour dough properties:

524

Ultracentrifugation and rheological measurements. Cereal Chem, 2002, 79(4),

525

544-545.

526

27. Yiannopoulos, S.; Kontogiorgi, A.; Poulli, E.; Krokida, M. Effect of Carob Flour

527

Addition on the Rheological Properties of Gluten-Free Breads. Food Bioproc

528

Tech. 2014, 7(3), 868-876.

529 530 531 532

28. Ong, Y. L.; Ross, A. S.; Engle, D. A. Glutenin Macropolymer in Salted and Alkaline Noodle Doughs. Cereal Chem. 2010, 87(1), 79-85. 29. Shiau, S. Y.; Yeh, A. I. Effects of alkali and acid on dough rheological properties and characteristics of extruded noodles. J Cereal Sci. 2001, 33(1), 27-37.

533

30. Lynch, E. J.; Dal Bello, F.; Sheehan, E. M.; Cashman, K. D.; Arendt, E. K.

534

Fundamental studies on the reduction of salt on dough and bread

535

characteristics. Food Res Int. 2009, 42(7), 885-891.

536

31. Salvador, A.; Sanz, T.; Fiszman, S. M. Dynamic rheological characteristics of wheat

537

flour-water doughs. Effect of adding NaCl, sucrose and yeast. Food

538

Hydrocolloid. 2006, 20(6), 780-786.

539

32. Shewry, P. R.; Halford, N. G.; Belton, P. S.; Tatham, A. S. The structure and

540

properties of gluten: an elastic protein from wheat grain. Philos Trans R Soc

541

Lond B Biol Sci. 2002, 357(1418), 133-142.

542

33. Wu, J. P.; Beta, T.; Corke, H., Effects of salt and alkaline reagents on dynamic

543

rheological properties of raw oriental wheat noodles. Cereal Chem. 2006, 83

544

(2), 211-217. 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

545

34. Lagrain, B.; Thewissen, B. G.; Brijs, K.; Delcour, J. A. Impact of redox agents on the

546

extractability of gluten proteins during bread making. J. Agric. Food. Chem.

547

2007, 55(13), 5320-5325.

548

35. Butow, B. J.; Gras, P. W.; Haraszi, R.; Bekes, F., Effects of different salts on mixing

549

and extension parameters on a diverse group of wheat cultivars using 2-g

550

mixograph and extensigraph methods. Cereal Chem. 2002, 79 (6), 826-833.

551 552 553

36. Letang, C.; Piau, M.; Verdier, C. Characterization of wheat flour-water doughs. Part I: Rheometry and microstructure. J. Food Eng. 1999, 41(2), 121-132. 37. Barak, S.; Mudgil, D.; Khatkar, B. S. Relationship of gliadin and glutenin proteins

554

with dough rheology, flour pasting and bread making performance of wheat

555

varieties. LWT-Food Sci Technol. 2013, 51(1), 211-217.

556 557 558

38. Doane, T. L.; Chuang, C. H.; Hill, R. J.; Burda, C., Nanoparticle zeta-Potentials. Acc. Chem. Res. 2012, 45 (3), 317-326. 39. Bruneel, C.; Lagrain, B.; Brijs, K.; Delcour, J. A. Redox agents and

559

N-ethylmaleimide affect the extractability of gluten proteins during fresh

560

pasta processing. Food Chem. 2011, 127(3), 905-911.

561

40. Wagner, M.; Morel, M. H.; Bonicel, J.; Cuq, B. Mechanisms of Heat-Mediated

562

Aggregation of Wheat Gluten Protein upon Pasta Processing. J. Agric. Food.

563

Chem. 2011, 59(7), 3146-3154.

564

41. Netto, L. E. S.; de Oliveira, M. A.; Monteiro, G.; Demasi, A. P. D.; Cussiol, J. R. R.;

565

Discola, K. F.; Demasi, M.; Silva, G. M.; Alves, S. V.; Faria, V. G.; Horta, B. B.,

566

Reactive cysteine in proteins: Protein folding, antioxidant. defense, redox

567

signaling and more. Comp Biochem Physiol C Toxicol Pharmacol. 2007, 146

568

(1-2), 180-193.

569

42. Wellner, N.; Bianchini, D.; Mills, E. N. C.; Belton, P. S., Effect of selected

570

Hofmeister anions on the secondary structure and dynamics of wheat

571

prolamins in gluten. Cereal Chem. 2003, 80 (5), 596-600.

572

43. Wellner, N.; Mills, E. N. C.; Brownsey, G.; Wilson, R. H.; Brown, N.; Freeman, J.;

573

Halford, N. G.; Shewry, P. R.; Belton, P. S., Changes in protein secondary

574

structure during gluten deformation studied by dynamic Fourier transform 20

ACS Paragon Plus Environment

Page 20 of 33

Page 21 of 33

Journal of Agricultural and Food Chemistry

575 576

infrared spectroscopy. Biomacromolecules. 2005, 6 (1), 255-261. 44. Lagrain, B.; Brijs, K.; Delcour, J. A. Reaction Kinetics of Gliadin-Glutenin

577

Cross-Linking in Model Systems and in Bread Making. J. Agric. Food. Chem.

578

2008, 56(22), 10660-10666.

579

45. Georget, D. M. R.; Belton, P. S. Effects of temperature and water content on the

580

secondary structure of wheat gluten studied by FTIR spectroscopy.

581

Biomacromolecules. 2006, 7(2), 469-475.

582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

605 606

Figure Captions

607

Figure 1. Typical Mixograms and Farinograms of soft wheat flour doughs with (P0) 0%

608

NaCl, (P1) 1.0% NaHCO3, (P2) 0.5% NaCl/0.3% NaHCO3, and (P3) 0.7% NaCl.

609 610

Figure 2. Dynamic rheological properties of the soft wheat flour doughs with

611

addition of salt/baking soda, (A) loss modulus G” as a function of frequency at a

612

strain of 0.1%, (B) storage modulus G’ as a function of frequency at a strain of 0.1%,

613

(C) loss modulus G” as a function of strain at a frequency of 1Hz, (D) storage modulus

614

G’ as a function of strain at a frequency of 1Hz, and (E) tan δ as a function of

615

frequency at a strain of 0.1%.

616 617

Figure 3. Typical RP-HPLC chromatographs of gluten extract separated from soft

618

wheat flour dough samples of (A) gliadins in control; and (B) glutenins in control,

619

using a Aeris WIDEPORE XB-C18 column (3.6 μm, 150 x 4.6 mm) with UV detection at

620

a wavelength of 210 nm. The flow rate was 0.6 mL/min at 60 °C and injection volume

621

was 10 µL. The mobile phases were water containing 0.1% trifluoroacetic acid

622

(solvent A), and acetonitrile containing 0.1% trifluoroacetic acid (solvent B).

623 624

Figure 4. Zeta potential of glutens isolated from soft wheat flour dough added by

625

salt/baking soda in deionized water with solid content of 0.1%.

626 627

Figure 5. Typical FTIR scan of gluten extracted from soft wheat flour dough showing

628

the different bands: (A) (C=O lipid vibration at around 1750 cm-1), (B) (amide I

629

1600-1700 cm-1), and (C) (amide II 1500-1580 cm-1).

630 631

Figure 6. FTIR deconvolution and second-derivative spectra of amide I region for

632

gluten extracted from soft wheat flour dough. A full discussion of the assignment is

633

given as extended chains (1600-1615 cm-1), β-Sheet (1624-1640, 1681 cm-1), random

634

coil (1640-1650 cm-1), α-helix (1650-1660 cm-1), β-turn (1660-1670, 1694 cm-1).45 22

ACS Paragon Plus Environment

Page 22 of 33

Page 23 of 33

Journal of Agricultural and Food Chemistry

635 636

Table 1. Mixing properties of soft wheat flour dough samples.

NaCl (% fwb)

NaHCO3 (% fwb)

0 0.5 0.7 1.0 0 0 0.5 0.5 1.0 1.0

0 0 0 0 0.3 1.0 0.3 1.0 0.3 1.0

Mixograph Water Abs (% fwb) 59 59 59 59 59 59 59 59 59 59

Mix Time (min) 3.50a 5.00c 4.25b 4.25b 6.25d 6.00d 5.00c 5.75d 5.25c 6.25d

Farinograph Water Abs (% fwb) 55.0 54.0 53.4 52.8 53.8 54.7 54.0 54.8 54.3 54.8

Mix Time (min) 3.0 3.3 3.0 3.6 3.2 7.0 4.3 4.3 4.2 4.1

Stability (min) 6.3 7.0 10.5 8.5 14.2 15.5 13.0 11.4 9.3 11.6

637

abcd

638

different at P < 0.05. Statistics were not run on farinograph data as there was only

639

one replicate.

Means with different superscripts within the same column are significantly

640

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

641

642 643

Page 24 of 33

Table 2. Dough properties with different NaCl and/or NaHCO3 addition. NaCl (% fwb)

NaHCO3 (% fwb)

0 0.5 0.7 1.0 0 0 0.5 0.5 1.0 1.0

0 0 0 0 0.3 1.0 0.3 1.0 0.3 1.0

Keiffer Test Force (g) 7.4f 12.4cde 8.5ef 12.7bcde 13.7bcd 15.0bc 9.9def 20.3a 14.3bcd 17.3ab

Distance (mm) 37.2bc 47.9ab 54.1a 49.9abc 57.0a 29.6c 55.7a 30.8c 51.7ab 32.7c

abcdef

Elongational Viscosity (Pa.s) 6477cd 4433e 6307cde 5046de 6403cd 12110a 8943b 9821b 6858c 9798b

Means with different superscripts within the same column are significantly different at P < 0.05.

24

ACS Paragon Plus Environment

Page 25 of 33

Journal of Agricultural and Food Chemistry

644 645

Table 3. Gluten contents of RP-HPLC fractions isolated from soft wheat flour doughs at the different levels of salt and baking soda addition.

Sample

HPLC elution range concentration (μg/mg) ωgliadin

αgliadin

γgliadin

Total gliadins

ωbgliadin

HMW- LMWGS GS

Ratio Total glutenins (gli:glu)

Control

15.0a

107.1a

91.5ab

213.7a

3.1a

30.7b

73.1ab

106.9c

2.0a

NaCl 0.5%

14.5a

106.4ab

90.3abc

211.3ab

3.2a

32.5ab

73.7ab

112.6bc

1.9a

NaCl 0.7%

15.1a

106.5a

88.0abc

209.7abc

3.5a

34.3ab

74.2ab

118.5abc

1.8abc

NaCl 1.0%

14.4a

103.2abc

83.0bc

200.6bcde

4.1a

34.4ab

80.6ab

125.7ab

1.6bc

NaHCO3 0.3%

14.3a

102.5abc

94.7a

211.4ab

3.2a

31.1ab

72.8ab

113.7abc

1.9ab

NaHCO3 1.0%

13.7a

99.5abcd

85.0bc

198.1cde

3.9a

33.0ab

81.3a

124.8ab

1.6bc

NaCl 0.5%/NaHCO3 0.3%

14.7a

100.2abcd

90.1a

205.0abcd

3.8a

30.8ab

71.9b

113.1bc

1.8ab

NaCl 0.5%/NaHCO3 1.0%

14.2a

97.3cd

85.1bc

196.6de

4.3a

33.8ab

79.5ab

124.2ab

1.6c

NaCl 1.0%/NaHCO3 0.3%

13.7a

98.3bcd

84.6bc

196.7de

4.3a

36.7a

79.3ab

126.9a

1.6c

NaCl 1.0%/NaHCO3 1.0%

14.3a

93.9d

82.2c

190.4e

4.3a

36.2ab

80.2ab

127.2a

1.5c

646 647 648 649 650 651 652 653 654 655

abcde

Means with different superscripts within the same column are significantly different at P < 0.05.

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 33

Table 4. The effect of baking soda or salt on the SH and SS content, secondary structure and surface hydrophobicity of glutens separated from soft wheat flour doughs.

Total SH contentA Free SH contentA SS contentA Extended Chains (%)B β-Sheet (%)B Random coil (%)B α-helix (%)B β-turn (%)B Hydrophobicity

control

salt 0.5%

salt 0.7%

salt 1.0%

soda 0.3%

soda 1.0%

56.8±1.5a 6.1±0.1a 25.4±0.5a 5.2±0.6a 19.7±1.4a 35.2±0.1a 31.4±0.8a 7.6±0.7a 13.4±1.5a

55.6±2.0a 5.2±0.1abc 25.2±0.7a 7.9±0.4a 22.0±1.8a 22.1±1.4b 36.6±0.2a 7.6±0.4a 13.1±1.1a

57.2±2.9a 5.3±0.1abc 26.2±0.9a 6.5±1.2a 23.2±1.0a 24.5±1.9b 32.9±1.6a 8.8±0.5a 13.4±1.0a

58.5±1.5a 5.0±0.2bc 26.8±0.5a 8.5±0.4a 24.3±2.1a 25.2±1.6b 32.9±5.1a 8.9±1.0a 13.3±0.7a

56.5±1.3a 5.5±0.1ab 25.5±0.9a 5.6±1.2a 22.6±0.4a 24.9±3.0b 28.8±0.9a 8.0±0.1a 13.8±0.9a

59.9±1.8a 3.9±0.2d 27.9±0.8a 5.8±0.5a 25.2±0.1a 23.3±1.0b 33.8±1.8a 7.6±1.0a 13.7±0.7a

abcd

salt 0.5% /soda 0.3% 56.3±3.3a 5.4±0.1ab 25.4±1.1a 8.4±0.4a 22.6±0.3a 23.2±0.4b 37.6±0.4a 8.1±0.1a 12.9±0.8a

salt 1.0% /soda 0.3% 58.2±3.1a 5.3±0.3abc 26.4±1.3a 7.9±0.1a 23.1±0.4a 28.9±0.4ab 39.4±1.0a 5.9±0.6a 13.4±1.1a

salt 0.5% /soda 1.0% 56.3±4.8a 4.4±0.1cd 25.9±1.7a 6.2±0.5a 21.9±2.4a 28.0±2.1ab 30.4±1.4a 6.7±1.0a 13.1±1.4a

salt 1.0% /soda 1.0% 57.1±4.6a 4.4±0.3cd 26.3±1.7a 6.8±1.4a 24.2±3.0a 28.1±0.2ab 31.8±1.5a 5.7±1.4a 13.5±1.1a

Means with different superscripts within the same row are significantly different at P < 0.05. Value is represented as the mean ± standard deviation (n=2). A SH and SS contents were expressed by nmoles/mg sample. B Gluten secondary structures were assignment as extended chains (1600-1615 cm-1), β-Sheet (1624-1640, 1681 cm-1), random coil (1640-1650 cm-1), α-helix (1650-1660 cm-1), β-turn (1660-1670, 1694 cm-1).45

26

ACS Paragon Plus Environment

Page 27 of 33

Journal of Agricultural and Food Chemistry

Figure 1.

27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

control salt 0.7% soda 0.3% salt 0.5%/soda 0.3% salt 1.0%/soda 0.3%

A 3.2E+04

salt 0.5% salt 1.0% soda 1.0% salt 0.5%/soda 1.0% salt 1.0%/soda 1.0%

Page 28 of 33

control salt 0.7% soda 0.3% salt 0.5%/soda 0.3% salt 1.0%/soda 0.3%

B 6.5E+04

salt 0.5% salt 1.0% soda 1.0% salt 0.5%/soda 1.0% salt 1.0%/soda 1.0%

2.2E+04 G'(Pa)

G'' (Pa)

4.5E+04

1.2E+04

2.5E+04

5.0E+03

2.0E+03 0.1

1 Frequency (Hz) soda 1.0% salt 0.5% salt 1.0% salt 0.5%/soda 0.3% salt 1.0%/soda 0.3%

C

0.1

10 control salt 0.7% soda 0.3% salt 0.5%/soda 1.0% salt 1.0%/soda 1.0%

1

Frequency (Hz)

control salt 0.7% soda 0.3% salt 0.5%/soda 0.3% salt 1.0%/soda 0.3%

D 4.2E+04

10 salt 0.5% salt 1.0% soda 1.0% salt 0.5%/soda 1.0% salt 1.0%/soda 1.0%

1.3E+04

G' (Pa)

9.0E+03

2.2E+04

5.0E+03 1.2E+04

1.0E+03 0.10%

1.00% Strain

10.00%

2.0E+03 0.10%

1.00%

Strain

E

0.4 tan δ

G'' (Pa)

3.2E+04

control

salt 0.5%

salt 0.7%

salt 1.0%

soda 0.3%

soda 1.0%

salt 0.5%/soda 0.3%

salt 0.5%/soda 1.0%

salt 1.0%/soda 0.3%

salt 1.0%/soda 1.0%

0.2 1

Frequency (Hz)

10

Figure 2. 28

ACS Paragon Plus Environment

10.00%

Page 29 of 33

Journal of Agricultural and Food Chemistry

Figure 3.

29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 30 of 33

8 5.57 4

3.92

Zeta potential (mV)

1.13

1.93

0

-3.69

-4 -5.73

-5.86 -8 -9.53 -12

-12.88 -16

Figure 4.

30

ACS Paragon Plus Environment

-11.46

Page 31 of 33

Journal of Agricultural and Food Chemistry

Figure 5.

31

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 6.

32

ACS Paragon Plus Environment

Page 32 of 33

Page 33 of 33

Journal of Agricultural and Food Chemistry

Enhanced dough physical and rheological properties with NaCl and/or NaHCO3 were a result of synergistic inter- and intra-molecular interactions 279x157mm (300 x 300 DPI)

ACS Paragon Plus Environment