Molecular Assembly of Wheat Gliadins into ... - ACS Publications

Sep 14, 2015 - and Reiko Urade*,‡. †. Research Reactor Institute, Kyoto University, 2-1010 Asashiro-nishi, Kumatori-cho, Sennan-gun, Osaka 590-049...
0 downloads 0 Views 955KB Size
Subscriber access provided by CMU Libraries - http://library.cmich.edu

Article

Molecular Assembly of Wheat Gliadins into Nanostructures: A Small-Angle X-Ray Scattering Study of Gliadins in Distilled Water over a Wide Concentration Range Nobuhiro Sato, Aoi Matsumiya, Yuki Higashino, Satoshi Funaki, Yuki Kitao, Yojiro Oba, Rintaro Inoue, Fumio Arisaka, Masaaki Sugiyama, and Reiko Urade J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.5b02902 • Publication Date (Web): 14 Sep 2015 Downloaded from http://pubs.acs.org on September 17, 2015

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 30

Journal of Agricultural and Food Chemistry

Molecular Assembly of Wheat Gliadins into Nanostructures: A Small-Angle X-Ray Scattering Study of Gliadins in Distilled Water over a Wide Concentration Range Nobuhiro Sato*,†, Aoi Matsumiya‡, Yuki Higashino‡, Satoshi Funaki‡, Yuki Kitao‡, Yojiro Oba†, Rintaro Inoue†, Fumio Arisaka§, Masaaki Sugiyama**†, and Reiko Urade***,‡

Corresponding authors *

**

***



Phone: +81 72 451 2499. Fax: +81 72 451 2633. E-mail: [email protected] Phone: +81 72 451 2670. Fax: +81 72 451 2670. E-mail: [email protected] Phone: +81 774 38 3760. Fax: +81 774 38 3765. E-mail: [email protected]

Research Reactor Institute, Kyoto University, 2-1010 Asashiro-nishi, Kumatori-cho,

Sennan-gun, Osaka 590-0494 JAPAN ‡

Division of Agronomy and Horticultural Science, Graduate School of Agriculture, Kyoto

University, Gokasho, Uji, Kyoto 611-0011 JAPAN §

Life Science Center, College of Bioresource Science, Nihon University, 1866 Kameino,

Fujisawa 252-0880 JAPAN

E-mail addresses of coauthors A. Matsumiya , Y. Higashino , S. Funaki , Y. Kitao , Y. Oba , R. Inoue , F. Arisaka

- 1Plus - Environment ACS Paragon

Journal of Agricultural and Food Chemistry

Page 2 of 30

1

ABSTRACT: Gliadin, one of the major proteins together with glutenin composing

2

gluten, affects the physical properties of wheat flour dough. In this study, nanoscale

3

structures of hydrated gliadins extracted into distilled water were investigated

4

primarily by small-angle X-ray scattering (SAXS) over a wide range of

5

concentrations.

6

analyses of SAXS profiles indicate that gliadins are present as monomers together

7

with small amounts of dimers and oligomers in a very dilute solution. The SAXS

8

profiles also indicate that interparticle interference appears above 0.5 wt% because of

9

electrostatic repulsion among gliadin assemblies. Above 15 wt%, gliadins form gel-

10

like hydrated solids. At greater concentrations, a steep upturn appears in the low-q

11

region owing to the formation of large aggregates, and a broad shoulder appears in the

12

middle-q region showing density fluctuation inside. This study demonstrates that

13

SAXS can effectively disclose nanostructure of hydrated gliadin assemblies.

Gliadins are soluble in distilled water below 10 wt%.

Guinier

14 15

KEYWORDS: gliadin; wheat protein; molecular nanostructures; viscoelastic

16

properties; SAXS; protein crosslinking

17

2 - Environment ACS Paragon -Plus

Page 3 of 30

Journal of Agricultural and Food Chemistry

18

INTRODUCTION

19

The rheological properties of dough are strongly dependent on the physical

20

properties of gluten,1 which is formed by mixing wheat flour with water. Gluten

21

consists of two component proteins: glutenin and gliadin. The strength, elasticity, and

22

extensibility of dough owe much to the viscoelastic properties of each component

23

protein. Glutenin is a polymer linked by intermolecular disulfide bonds, subunits of

24

which are classified into high- and low-molecular-weight ones.2,3 High-molecular-

25

weight glutenin subunits contribute to the elasticity of dough through a network

26

structure connected by intermolecular disulfide bonds.2,4 In contrast, gliadin is a

27

monomeric protein that is responsible for the viscosity of dough.5 They are classified

28

into α-, γ-, and ω-gliadins according to their primary structure,2,6 and usually are

29

soluble in dilute acids or 60–70% ethanol-water solutions. It has been shown that

30

noncovalent intermolecular interactions exert a dominant effect on the viscoelastic

31

properties of hydrated gliadins.7 It has also been demonstrated that gliadins weaken

32

the strength of dough and soften gluten.8-11 Previous studies on the sulfur-rich α- and

33

γ-gliadins in dilute acids or ethanol-water solutions have indicated that the C-terminal

34

domains of those gliadins are rich in α-helical content and adopt a compact globular

35

structure stabilized by disulfide bonds.2,12-17 The N-terminal repetitive domains are

36

nonglobular and contain elements of a poly-L-proline II structure and β-reverse

37

turns.18-20 Because gliadins have known to have poor water solubility,21 analyses such

38

as electrophoresis and viscoelastic measurements have been performed on gliadins

39

that were extracted with an ethanol-water solution22 or dilute acetic acid.23,24

40

However, it is not known if gliadins in alcohols or acids show the same behavior as

41

those in water. In addition, gliadins exist as a hydrated solid in food. It is therefore

42

debatable whether data obtained for gliadins in solution are valid for hydrated solids.

3 - Environment ACS Paragon -Plus

Journal of Agricultural and Food Chemistry

Page 4 of 30

43

Detailed investigation of the structure of gliadins in relation to water content is

44

required to clarify the relationship between superstructures and rheological properties

45

of hydrated solid gliadins.

46

Classical methods utilizing a 60–70% alcohol or acid solution were primarily

47

used to extract gliadins from wheat flour. In 2008, we found that most gliadins could

48

be extracted from dough made with 0.51 M NaCl by washing with distilled water.25

49

The extracted proteins were primarily found to be gliadin monomers by N-terminal

50

amino acid sequencing and analytical ultracentrifugation. These gliadins are free

51

from the drawback of higher-order structural changes that may occur through

52

treatment with alcohols or acids and, accordingly, are expected to exhibit the same

53

structure and properties as gliadins in dough. It was also found that the gliadin

54

preparation was soluble in distilled water up to 10 wt% and readily aggregated when

55

NaCl or other salts were added.25 Thus, the solubility behavior of extracted gliadins

56

needs to be clarified in relation to the structure of the proteins. Small-angle X-ray

57

scattering (SAXS) is known to be a powerful method for analyzing the nanoscale

58

structure of various materials. It provides ensemble-averaged structural information

59

on the mass distribution and shape of non-crystalline molecules in a non-destructive

60

manner. Hence, it has been utilized widely for the nanoscale structural analysis of

61

soft matters such as polymer gels, emulsions, and liquid crystals.26,

62

characteristic merits of SAXS also can be used effectively for the structural analysis

63

of proteins such as gliadins, not only of the gliadin molecule itself, but also of

64

aggregated forms at high concentrations. Some results have been reported to date

65

concerning the structural analysis on gluten proteins by SAXS. Matsushita et al.

66

investigated the structure of high-molecular-weight glutenin subunits by SAXS in 1-

67

propanol / acetic acid / water mixed solutions in the concentration range of 1.0–10.0

4 - Environment ACS Paragon -Plus

27

The

Page 5 of 30

Journal of Agricultural and Food Chemistry

68

mg/mL.28 From Guinier and cross-section Guinier plots, they determined glutenins to

69

have a rod-like shape. In another SAXS study, Thomson et al.20, 29 examined α-, γ-,

70

and ω-gliadins and high-molecular-weight glutenin subunits in the alcohol/water

71

mixed solutions in the concentration range about 1–10 mg/mL and analyzed their

72

SAXS profiles by means of Guinier and cross-section Guinier plots. They found that

73

gliadins in this concentration range have a prolate ellipsoid shape with a length of 15–

74

20 nm and a diameter of about 3.5 nm. These results shed light on the nanostructure

75

of gluten protein molecules in dilute solution, but their structure in the alcohol-

76

containing water is not necessarily the same as that in alcohol-free water. In addition,

77

the structure of gluten proteins at high concentrations is still unclear. In this study, we

78

focus on gliadins extracted with distilled water and report on the SAXS analysis of

79

their structures over a wide range of concentrations in aqueous solutions. We also

80

correlate the results of an electrophoresis measurement and a sedimentation velocity

81

experiment with the nanostructure of gliadins established by SAXS.

82 83 84

MATERIALS AND METHODS Materials.

Gliadins used in this study were prepared by the following

85

procedure, the details of which are described in our previous paper.25 Dough was

86

prepared by mixing and kneading 100 g of wheat flour (Super KingTM, Nisshin Flour

87

Milling, Inc.; 13.8% protein, 0.42% ash, 14% water) with 67 mL of 0.5 M NaCl in a

88

mixer (KN-200, Taisho Denki, Tokyo, Japan) for 20 min, and then repeatedly

89

kneaded with one hand in 500 mL of distilled water for 15 min at 15 °C. The first and

90

second wash solutions were discarded. The third to sixth washing solutions were

91

collected and centrifuged at 18,000 × g for 10 min at 20 °C. After NaCl was added to

92

the supernatant to adjust its concentration to 0.5 M, the precipitated gliadins were

5 - Environment ACS Paragon -Plus

Journal of Agricultural and Food Chemistry

93

collected by centrifugation at 18,000 × g for 20 min at 20 °C, suspended in 300 mL of

94

distilled water, dialyzed five times against 3 L of distilled water for 8 h at 4 °C, and

95

finally freeze-dried. The gliadins were also extracted from wheat flour with 70%

96

ethanol according to the method described by Tatham and Shewry.13 Protein was

97

assayed with an RC.DC protein assay kit from Bio-Rad Laboratories (Hercules, CA).

98

Sedimentation Velocity Experiments. Sedimentation velocity experiments

99

on 0.02 wt% gliadins dissolved in distilled water were performed with an Optima XL-

100

I unit (Beckman Coulter, Inc. Brea, CA, USA) using an An50Ti rotor at 20°C.

101

Charcoal-filled Epon double sector cells were used with the reference sectors filled

102

with distilled water.

103

successive scans.

104

obtained by use of the SEDFIT program.30, 31 The partial specific volume of gliadin

105

was set to 0.73, which is the default value.32 A protocol for calculating the molecular

106

weight corresponding to each peak using the Svedberg equation was implemented in

107

SEDFIT, where a common frictional ratio, f/f0, was assumed for all molecular species.

108

Electrophoretic Analysis of Gliadin Preparations. Gliadins extracted into

109

distilled water or 70% ethanol were subjected to 2D-PAGE isoelectric focusing and

110

SDS-PAGE.33 ,34 Gliadins were treated with a 2D clean-up kit (GE Healthcare UK

111

Ltd.) and applied to 7-cm Ready-Strip IPG Strips (Bio-Rad Laboratories). Isoelectric

112

focusing was performed using a Protean IEF Cell (Bio-Rad Laboratories). The strips

113

were then subjected to SDS-PAGE and stained with Coomassie Brilliant Blue R-250.

All data were acquired without time intervals between

The sedimentation coefficient distribution function, c(s), was

114

SAXS Experiments. The SAXS experiments were carried out at the beam

115

line BL-10C of Photon Factory, a synchrotron radiation facility of Institute of

116

Materials Structure Science, High Energy Accelerator Research Organization (KEK).

117

Dilute solution samples were measured in a 1-mm thick aluminum cell with 20-µm

6 - Environment ACS Paragon -Plus

Page 6 of 30

Page 7 of 30

Journal of Agricultural and Food Chemistry

118

thick quartz windows. Viscous hydrated solids were measured in a 1-mm thick PTFE

119

sandwich cell with windows of 7.5-µm thick Kapton® films (TORAY-DuPont). The

120

X-ray wavelength was 0.1488 nm, and the observed range of scattering vector

121

(denoted by q) was 0.06–2.5 nm−1 as calibrated with silver behenate. The scattered

122

beam was detected with two-dimensional area detectors. The typical X-ray exposure

123

time was 300 s. Samples were used for SAXS measurements at least four days after

124

preparation. Measured two-dimensional data were converted into one-dimensional

125

scattering profiles followed by the standard procedure of data correction for cell and

126

background scatterings, transmission, and beam intensity.

127

Crosslinking Experiments.

Gliadin solutions were incubated with and

128

without 5 mM 3,3'-dithiobis(sulfosuccinimidyl propionate) (DTSSP) from Dojindo

129

(Tokyo Japan) at 25 °C for 60 min. After addition of N-ethylmaleimide (20 mM), the

130

solutions were incubated for 10 min at room temperature and then freeze-dried.

131

Laemmli's SDS sample buffer31 without reducing reagent was added to the gliadin

132

solutions, which were then boiled for 5 min. The suspension was centrifuged at

133

10,000 × g for 40 min at room temperature. The supernatant (30 µg protein) was

134

subjected to SDS-PAGE under non-reducing conditions.

135

supernatant (100 µg protein) was subjected to SDS-PAGE with a NativePAGETM 3–

136

12 % gel (Invitrogen, Carlsbad, CA) and electrophoresed with 50 mM bis(2-

137

hydroxyethyl)amino-tris(hydroxymethyl)methane-HCl (Bis-Tris HCl) buffer having a

138

pH of 6.8, including 50 mM tricine and 0.1 % SDS as running buffer, under

139

nonreducing conditions. After electrophoresis, the sample lane was cut from the gel

140

and immersed in Laemmli's SDS sample buffer with 5% 2-mercaptoethanol for 30

141

min and subjected to SDS-PAGE under reducing conditions. The proteins were

142

stained with Coomassie Brilliant Blue R-250.

7 - Environment ACS Paragon -Plus

Another part of the

Journal of Agricultural and Food Chemistry

143 144

RESULTS AND DISCUSSION

145

Gliadin-rich proteins extracted from NaCl-containing dough with

146

distilled water. Previously we found that gliadins can be extracted with distilled

147

water from dough prepared with wheat flour and a 0.51 M NaCl aqueous solution.25

148

When the dough made with 100 g wheat flour was washed with 500 mL distilled

149

water, starch, albumins, and globulins were eluted from the dough into the first and

150

second wash solutions. However, gliadins were eluted into the third to sixth wash

151

solutions when the NaCl content in the dough becomes less than 5 mg/100 g dough.25

152

In the present study, we also observed abundant gliadins in the third and subsequent

153

extracts by repeating the extraction procedure. The amount of extracted proteins

154

(2.63 g from 100 g flour) was almost same as that of gliadins extracted with 70%

155

ethanol (2.53 g from 100 g flour). 2D-PAGE shows that the composition of proteins

156

extracted by our procedure was almost same as that of gliadins extracted with 70%

157

ethanol except for a minor spot at 75 kDa in distilled water, which is probably due to

158

contamination (Figure 1).

159



160

The gliadins obtained by our procedure can be dissolved in distilled water to

161

yield a transparent solution at concentrations of 10 wt% or less, while they form gel-

162

like hydrated solids at concentrations of 15 wt% or more. Gliadins resting in a plastic

163

tube at 40 wt% or more do not flow even upon turning the tube upside down, which

164

suggests the presence of a network structure.

165

Molecular weight distribution. A sedimentation velocity analysis revealed

166

the molecular weight distribution of the gliadin in 0.02% distilled water solutions as

167

shown in Figure 2. The distribution curve shows the largest peak at a sedimentation

8 - Environment ACS Paragon -Plus

Page 8 of 30

Page 9 of 30

Journal of Agricultural and Food Chemistry

168

coefficient corresponding to a molecular weight of 25.7 kDa indicating that most of

169

the extracted gliadins are in the monomeric form. A small, broad peak found at a

170

molecular weight of ca. 103 kDa suggests that associated gliadins exist in addition to

171

monomeric gliadins.

172

Nanostructure of gliadins in distilled water as revealed by SAXS. Figure 3

173

shows SAXS profiles of gliadins in distilled water at various concentrations: (a)

174

solution components of low concentration samples (≤ 10 wt%) and (b) high

175

concentration gel-like solid samples (≥ 15 wt%). Because of a large difference in the

176

physical state between the two regions, we discuss each concentration region

177

separately.

178



179

In the low concentration region, the scattering intensity increases gradually

180

with increasing concentration. A small shoulder appears near q = 0.3 nm−1 at 0.5 wt%

181

and gradually shifts to 0.13 nm−1 at 10 wt% with increasing intensity. The emergence

182

of this shoulder indicates the presence of considerable interparticle interference.

183

Therefore, it is required that the behavior of isolated gliadin molecules be observed at

184

concentrations below 0.5 wt%. We first attempted to evaluate the size of gliadin

185

molecules at concentrations below 0.5 wt% by means of a Guinier analysis of the

186

scattering curves. However, the results could not be well fitted with a simple Guinier

187

relationship, I = A exp(−Rg2 q2 /3), indicating a polydispersity of scatterers. Thus, we

188

employed the three-component Guinier equation expressed as I(q) = A1 exp(−Rg12 q2

189

/3) + A2 exp(−Rg22 q2 /3) + A3 exp(−Rg32 q2 /3). As shown in Figure 4a, the scattering

190

profiles are well fitted by this equation; the fitting parameters are summarized in

191

Table 1. The value of the smallest radius of gyration ranges from 4.10 to 6.46 nm.

192

Thomson et al.20 previously reported Rg values of 3.55, 3.80, and 4.60 nm for α-, γ-,

9 - Environment ACS Paragon -Plus

Journal of Agricultural and Food Chemistry

193

and ω-gliadins in 70% ethanol, respectively. By comparing their data with the Rg1

194

values of our data, we can say that gliadin molecules are present as monomers in the

195

0.025 wt% solution, although some molecules with larger Rg values such as gliadin

196

dimers and oligomers are also present. The presence of these larger molecules is

197

consistent with the results of the sedimentation velocity experiment described above.

198

The values of Rg1 increase slightly with increasing concentration indicating that some

199

gliadin monomers associate to dimers at higher concentrations. The parameters Rg2

200

and Rg3 in Table 1 are much larger than that of gliadin monomers and likely

201

correspond to larger aggregates or a high-molecular-weight contaminant (e.g.

202

glutenin).

203



204



205

When anisotropic particles, such as those possessing a rod-like shape, are

206

examined by SAXS, a cross-section Guinier plot may be used to obtain information

207

about the cross-section size. Figure 4b shows the cross-section Guinier plots for the

208

three low concentration samples. The scattering curves in the low-q region were well

209

fitted with straight lines obeying the relationship qI(q) = A exp(−Rc2 q2 /2). The

210

resulting values for the cross-sectional radius of gyration, Rc, are listed in Table 2.

211

These values are more than twice as large as those reported by Thomson et al.20 The

212

reason for this is due in part to the aggregation of gliadin molecules, but also may be

213

ascribable to structural differences between gliadins in ethanol and in water. The rod

214

length, Λ, obtained from the equation, Rg2 − Rc2 = Λ 2 / 12, is also presented in Table 2.

215

The radii and length become larger at 0.5 wt%. However, we think that this change is

216

due to an increasing number of aggregated gliadins rather than to a change of

10 - Environment ACS Paragon-Plus

Page 10 of 30

Page 11 of 30

Journal of Agricultural and Food Chemistry

217

molecular size itself; accordingly, the average value of each dimension becomes

218

larger.

219



220

In the 1–10 wt % concentration range, Guinier plots do not yield reasonable

221

physical values due to interparticle interference. Instead, the correlation length, L,

222

was obtained from the Bragg equation, L = 2π /qs, where qs is the magnitude of the

223

scattering vector of the shoulder appearing above 1 wt%. As shown in Table 3, L

224

increases with increasing concentration, which suggests that the size of aggregated

225

gliadins becomes larger and that the distance between aggregates increases under

226

these conditions.

227



228

At concentrations above 15 wt%, gliadins are insoluble in water and form gel-

229

like hydrated solids. The SAXS profiles in this region are displayed in Figure 3b.

230

The scattering curve at 15 wt% is similar to that of the 10 wt% solution. This fact

231

indicates that the nanostructure of these two samples are similar despite the difference

232

in water solubility. It is suggested that only intermolecular association occurs in this

233

concentration range without a significant change in the size and distribution of

234

aggregated gliadins. As concentration increases further, the shoulder near 0.13 nm-1

235

diminishes while the upturn in the low-q region grows and a broad shoulder appears

236

in the higher-q region. When the low-q changes are examined closely, the 0.13 nm-l

237

shoulder shifts to the even lower q and finally protrudes out of the observable q-range.

238

This transition suggests the evolving formation of gliadin aggregates.

239

concentration increases, isolated gliadin aggregates coalesce to form larger aggregates

240

resulting in the growth of the low-q upturn. Concurrently, the density fluctuation

241

inside the aggregates also increases, which is manifested as a broad peak in the high-q

11 - Environment ACS Paragon-Plus

As

Journal of Agricultural and Food Chemistry

242

region. The characteristic peaks observed in this region are summarized in Table 3.

243

When the concentration reaches 70 wt%, the broad shoulder is almost unrecognizable,

244

and only the upturn in the low-q region remains.

245

Association of gliadins examined by crosslinking experiments. The SAXS

246

analysis described above indicates noncovalent association of gliadins in solution.

247

Thus, crosslinking experiments were performed with DTSSP, a crosslinking molecule

248

that is cleavable with reducing agents after crosslinking, to examine the specificity of

249

association between gliadin molecules. By applying nonreducing SDS-PAGE to the

250

soluble fraction after centrifugation, it is found that the gliadin bands disappear almost

251

entirely from the DTSSP-treated samples at concentrations of 8 wt% and above

252

(Figure 5a). This suggests that gliadins crosslink to form large polymers at 8 wt% and

253

above and become insoluble in an SDS solution. Monomeric and crosslinked gliadins

254

are detected in the supernatant for 0.5–7 wt% samples treated with DTSSP. The most

255

prominent bands of crosslinked gliadins correspond to 50–75 kDa in size. Gliadins at

256

7 wt% were analyzed by 2D-PAGE (Figure. 5b). By the first nonreducing SDS-

257

PAGE with bis-Tris-HCl buffer containing tricine, crosslinked gliadins were detected

258

as ~55–65 kDa (i), ~65–75 kDa (ii), ~90–120 kD (iii), ~140–180 kDa (iv) in addition

259

to the large smear bands (Figure 5b, panel 1). The crosslinked gliadins were cleaved

260

to monomers in the gel after the first nonreducing SDS-PAGE by reduction with 2-

261

mercaptoethanol and separated by the second reducing SDS-PAGE. The crosslinked

262

~55–65 kDa bands on the first nonreducing SDS-PAGE are composed of 30–35 kDa

263

gliadins, which have been shown to be a mixture of α- and γ-gliadins by N-terminal

264

amino acid sequence analysis.25 The 65–75 kDa bands on the first nonreducing SDS-

265

PAGE are composed of ~37 and ~40 kDa gliadins, which have been identified as α-

266

gliadins.25 These results suggest that some gliadins associate specifically in distilled

12 - Environment ACS Paragon-Plus

Page 12 of 30

Page 13 of 30

Journal of Agricultural and Food Chemistry

267

water. The crosslinked ~55–65 kDa and ~65–75 kDa bands are predicted to be

268

dimers based upon their sizes. In addition, the 90–120 and140–180 kDa smear bands

269

on the first nonreducing SDS-PAGE were shown to be composed of 30–35, ~37, and

270

~40 kDa gliadins, suggesting that these bands originate from the formation of

271

oligomers.

272



273

Plausible model for gliadin association in distilled water. From above

274

results, we present the model shown in Figure 6 for the nanoscale structures of

275

gliadins in distilled water. In the 0.025–0.5 wt% concentration range, gliadins are

276

dispersed in water as isolated particles (Figure 6a), because no shoulder peak is found

277

due to interparticle interference.

278

populations of dimers and oligomers are also present based on the results of the

279

sedimentation velocity measurement and the SAXS analyses.

280

reported that α-, γ-, and ω-gliadins in 70% ethanol show Rg values of 3.55, 3.80, and

281

4.60 nm, respectively, at concentrations below 1 wt%.

282

obtained in this study are slightly larger. This suggests that the structures of gliadin

283

monomers in distilled water differ somewhat from those in ethanol. The cross-

284

sectional radius of gyration and the length of gliadins at the lowest concentration are

285

2.50 nm and 11.3 nm, respectively, which is consistent with a rod-like shape.

286

Although some gliadins form dimers and oligomers at the lowest concentration, the

287

shape of the associated particles is rod-like. The frictional ratio of 1.52 obtained from

288

the sedimentation velocity analysis supports the rod-like model.

Most of the particles are monomers, but small

Thomson et al.

However, the Rg values

289

In the 0.5–10 wt% concentration range, gliadin molecules self-assemble and

290

form small soluble aggregates (Figure 6b) by noncovalent interactions. Interparticle

291

interference is apparent, but gliadins remain soluble in distilled water up to at least 10

13 - Environment ACS Paragon-Plus

Journal of Agricultural and Food Chemistry

292

wt%. This interparticle interference arises from the electrostatic repulsion between

293

small aggregates, which is confirmed by the fact that the SAXS shoulder peak

294

vanishes when NaCl is added to the solution at a concentration of 10 mM (data not

295

shown). The correlation length among small aggregates gradually increases with

296

increasing concentration indicating that the interparticle distance is becoming greater.

297

At 15–20 wt%, gliadins become insoluble in distilled water but the SAXS

298

profiles are similar to those at 10 wt% concentration (Figure 6c). We suggest that the

299

small aggregates become more closely spaced and come in contact with one another

300

to form a continuous network without significantly changing aggregation structure. In

301

the SAXS profiles at concentrations above 30 wt%, the shoulder near 0.13 nm-1

302

disappears and broad peaks at 0.3–1.0 nm-1 become noticeable. In this region, the

303

space is almost entirely filled with gliadin molecules and a definite correlation length

304

disappears; however, density fluctuation is present inside the aggregated gliadins

305

(Figure 6d). Above 50 wt% (Figure 6e), the broad peaks diminish and only the upturn

306

in the low-q region remains, suggesting that gliadins form large aggregates of more

307

than 100 nm in size.

308



309

Finally the influence of NaCl on the structure of associated gliadins should be

310

noted. In this study, NaCl acts as a water-solubilizer of gliadins in dough, but it is

311

well known that NaCl also causes the precipitation of gliadin in aqueous solutions.25

312

Because these two observations appear to be contradictory, it is of interest to clarify

313

the effect of NaCl addition on the state of association of gluten proteins in relation to

314

the intermolecular interactions among them. This issue will be addressed in our next

315

paper, which will focus on the difference between the association states of highly

14 - Environment ACS Paragon-Plus

Page 14 of 30

Page 15 of 30

Journal of Agricultural and Food Chemistry

316

concentrated gliadins hydrated with distilled water and the precipitates produced by

317

NaCl addition.

318

In conclusion, SAXS analyses have clarified the state of association of

319

gliadins in distilled water over a wide concentration range. At low concentrations (≤

320

0.5 wt%), gliadin molecules are present as monomers plus a minor amount of dimers

321

and oligomers. At greater concentrations, gliadin molecules begin to coalesce (0.5–1

322

wt%) and then show interparticle interference among aggregated molecules (1–10

323

wt%). In these regions, the correlation length increases as concentration increases,

324

because the aggregates grow in size and their mutual separation distance increases.

325

Crosslinking experiments indicate that association of gliadin molecules at 0.5–10 wt%

326

occurs in a specific manner. Both dimers and oligomers are formed, but some are

327

preferentially composed only of α-gliadins instead of a mixture of α- and γ-gliadins.

328

Gliadin becomes insoluble at 15 wt%, although the SAXS profile does not show a

329

significant difference from that at 10 wt%.

330

continuous network of aggregated gliadins. As concentration increases further, the

331

space is filled with gliadin molecules while a density fluctuation remains at ~40 wt%,

332

and finally the large aggregates are formed.

This is due to the formation of a

333 334

Abbreviations Used

335

SDS, sodium dodecyl sulfate; PAGE, polyacrylamide gel electrophoresis; Bis-Tris,

336

bis(2-hydroxyethyl)amino-tris(hydroxymethyl)methane;

337

dithiobis[sulfosuccinimidylpropionate]; SAXS, small-angle X-ray scattering.

338 339

Acknowledgment

15 - Environment ACS Paragon-Plus

DTSSP,

3,3'-

Journal of Agricultural and Food Chemistry

340

This work was supported in part by JSPS KAKENHI Grant Number 26660111 (R.U.),

341

15H02042 (M.S.), and 24310068 (M.S.), MEXT KAKENHI Grant Number 26102524

342

(M.S.), and a grant from the Tojuro Iijima Foundation for Food Science and

343

Technology (R.U.). This work has been performed under the approval of the Photon

344

Factory Program Advisory Committee (Proposal No. 2014G127 and 2014G162).

345 346

References

347

(1) Hamer, R. J.; MacRitchie, F.; Weegels, P. L. Structure and functional properties of

348

gluten. In Wheat: Chemistry and Technology, 4th ed.; Khan, K., Schewry, P. R., Eds.;

349

AACC International Press: St. Paul, Minnesota, 2009; pp. 153-178.

350

(2) Shewry, P. R.; D'Ovidio, R.; Lafiandra, D.; Jenkins, J. A.; Mills, E. N. C.; Békés,

351

F. Wheat Grain Proteins. In Wheat: Chemistry and Technology, 4th ed.; Khan, K.,

352

Schewry, P. R., Eds.; AACC International Press: St. Paul, Minnesota, 2009; pp. 223-

353

298

354

(3) Cornec, M.; Popineau, Y.; Lefebvre, J. Characterisation of gluten subfractions by

355

SE-HPLC and dynamic rheological analysis in shear. J. Cereal Sci. 1994, 19, 131-139.

356

(4) Anjum, F. M.; Khan, M. R.; Din, A.; Saeed, M.; Pasha, I.; Arshad, M. U. Wheat

357

gluten: High molecular weight glutenin subunits—structure, genetics, and relation to

358

dough elasticity. J. Food Sci. 2007, 72, R56–R63.

359

(5) Uthayakumaran, S.; Gras, P. W.; Stoddard, F. L.; Békés, F. Effect of varying

360

protein content and glutenin-to-gliadin ratio on the functional properties of wheat

361

dough. Cereal Chem. 1999, 76, 389-394.

362

(6) Kasarda, D. D.; Adalstein, A. E.; Laird, N. F. γ-Gliadins with α-type structure

363

coded on chromosome 6B of the wheat (Triticum aestivum L.) cultivar ”Chinese

16 - Environment ACS Paragon-Plus

Page 16 of 30

Page 17 of 30

Journal of Agricultural and Food Chemistry

364

Spring.” In Proc. 3rd Int. Workshop on Gluten Proteins; Laztity, A., Békés, F., Eds.;

365

World Scientific Publishing: Singapore, 1988; pp. 20-29.

366

(7) Marting, S. E.; Mulvaney, S. J.; Cohen, C. Effect of moisture content on

367

viscoelastic properties of hydrated gliadin. Cereal Chem. 2004, 81, 207-219.

368

(8) Fido, R. J.; Békés, F.; Gras, P. W.; Tatham, A. S. Effects of α-, β-, γ- and ω-

369

gliadins on the dough mixing properties of wheat flour. J. Cereal Sci. 1997, 26, 271-

370

277.

371

(9) Khatkar, B. S.; Fido, R. J.; Tatham, A. S.; Schofield, J. D. Functional properties of

372

wheat gliadins. II. Effects on dynamic rheological properties of wheat gluten. J.

373

Cereal Sci. 2002, 35, 307-313.

374

(10) Khatkar, B. S.; Barak, S.; Mudgil, D. Effects of gliadin addition on the

375

rheological, microscopic and thermal characteristics of wheat gluten. Int. J. Biol.

376

Macromol. 2013, 53, 38-41.

377

(11) Lee, C. C.; Mulvaney, S. J. Dynamic viscoelastic and tensile properties of gluten

378

and glutenin gels of common wheats of different strength. J. Agric. Food Chem. 2003,

379

51, 2317-2327.

380

(12) Kasarda D. D.; Bernardin, J. E.; Gaffield, W. Circular dichroism and optical

381

rotatory dispersion of α-gliadin. Biochemistry 1968, 7, 3950-3957.

382

(13) Tatham A. S.; Shewry P. R. The conformation of wheat gluten proteins. The

383

secondary structures and thermal stabilities of α-, β-, γ- and ω-gliadins. J. Cereal Sci.

384

1985, 3, 103-113.

385

(14) Tatham, A. S.; Field, J. M.; Smith, S. J.; Shewry, P. R. The conformations of

386

wheat gluten proteins, II*, aggregated gliadins and low molecular weight subunits of

387

glutenin. J. Cereal Sci. 1987, 5, 203-214.

17 - Environment ACS Paragon-Plus

Journal of Agricultural and Food Chemistry

388

(15) Tatham, A. S.; Massont, P.; Popineaut, Y. Conformational studies of peptides

389

derived by the enzymic hydrolysis of a gamma-type gliadin. J. Cereal Sci. 1990, 11,

390

1-13.

391

(16) Purcell J. M.; Kasarda, D. D.; Wu C-S. C. Secondary structures of wheat α- and

392

ω-gliadin proteins: Fourier transform infrared spectroscopy. J. Cereal Sci. 1988, 7,

393

21-32.

394

(17) Blanch, E. W.; Kasarda, D. D.; Hecht, L.; Nielsen, K.; Barron, L. D. New insight

395

into the solution structures of wheat gluten proteins from Raman optical activity.

396

Biochemistry 2003, 42, 5665-5673.

397

(18) Cole, E. W.; Kasarda, D.; Lafiandra, D. The conformational structure of Α-

398

gliadin: Intrinsic viscosities under conditions approaching the native state and under

399

denaturing conditions. Biochim. Biophs. Acta, 1984, 787, 244-251.

400

(19) Shewry, P. R.; Miles, M. J.; Thomson, N. H.; Tatham, A. S. Scanning probe

401

microscopes—Applications in cereal science. Cereal Chem. 1997, 74, 193-199.

402

(20) Thomson, N. H.; Miles, M. J.; Popineau, Y.; Harries, J.; Shewry, P.; Tatham, A.

403

S. Small angle X-ray scattering of wheat seed-storage proteins: α-, γ- and ω-gliadins

404

and the high molecular weight (HMW) subunits of glutenin. Biochim. Biophys. Acta

405

1999, 1430, 359-366.

406

(21) Shewry, P. R.; Tatham, A. S. The characteristics, structures and evolutionary

407

relationship of prolamins. In Seed Proteins; Shewry P. R., Eds.; Kluwer Academic

408

Publications: Dordrecht, The Netherlands, 1999; pp. 11-33.

409

(22) Osborne, T. B. The proteins of wheat-kernel; Carnegie Institution: Washington

410

DC, 1907.

411

(23) Blish, M. J.; Sandstedt, R. M. An improved method of the preparation of wheat

412

gliadin. Cereal Chem. 1926, 3, 144-149.

18 - Environment ACS Paragon-Plus

Page 18 of 30

Page 19 of 30

Journal of Agricultural and Food Chemistry

413

(24) Bernerdin, J. E.; Kasarda, D. D.; Mecham, D. K. Preparation and characterization

414

of α-gliadin. J. Biol. Chem. 1967, 242, 445-450.

415

(25) Ukai, T.; Matsumura, Y.; Urade, R. Disaggregation and reaggregation of gluten

416

proteins by sodium chloride. J. Agric. Food Chem. 2008, 56, 1122-1130.

417

(26) Guinier, A.; Fournet, G. Small-Angle Scattering of X-rays; Wiley-Interscience: 

418

New York, 1955.

419

(27) Glatter, O.; Kratky, O. Small angle X-ray scattering; Academic Press: London,

420

1982.

421

(28) Matsushima, N; Danno, G.; Sasaki, N.; Izumi, Y. Small-angle X-ray scattering

422

study by synchrotron orbital radiation reveals that high molecular weight subunit of

423

glutenin is a very anisotropic molecule. Biochim. Biophys. Res. Commun. 1992, 186,

424

1057-1064.

425

(29) Thomson, N. H.; Miles, M. J.; Tatham, A. S.; Shewry, P. R. Molecular images of

426

cereal proteins by STM. Ultramicroscopy 1992, 42, 1204-1213.

427

(30) Schuck, P. Size distribution analysis of macromolecules by sedimentation

428

velocity ultracentrifugation and Lamm equation modeling. Biophys. J. 2000, 78,

429

1606-1619.

430

(31) Schuck, P.; Perugini, M. S.; Gonzales, N. R.; Howlett, G. J.; Schubert, D. Size-

431

distribution analysis of proteins by analytical ultracentrifugation: strategies and

432

application to model systems. Biophys. J. 2002, 82, 1096-1111.

433

(32) Laue, T. M.; Shah, B. D.; Ridgeway, T. M.; Pelletier, S. L. Computer-aided

434

interpretation of analytical sedimentation data for proteins. In Analytical

435

ultracentrifugation in biochemistry and polymer science. Harding, S. E., Rowe, A. J.,

436

Horton, J. C., Eds.; Royal Society of Chemistry: Cambridge, United Kingdom, 1992;

437

pp. 90-125.

19 - Environment ACS Paragon-Plus

Journal of Agricultural and Food Chemistry

438

(33) O'Farrell, P. H. High resolution two-dimensional electrophoresis of proteins. J.

439

Biol. Chem. 1975, 250, 4007-4021.

440

(34) Laemmli, U. K. Cleavage of structural proteins during the assembly of the head

441

of bacteriophage T4. Nature 1970, 227, 680-685.

442

20 - Environment ACS Paragon-Plus

Page 20 of 30

Page 21 of 30

Journal of Agricultural and Food Chemistry

Figure Captions Figure 1. Two-dimensional PAGE of gliadin fractions. (a) Gliadin extracted from dough made with wheat flour and a 0.5 M sodium chloride solution by washing with distilled water. (b) Gliadin extracted from wheat flour with 70% ethanol.

Figure 2. Sedimentation velocity analysis of 0.02 wt% gliadins in distilled water. (a) Raw data of sedimentation velocity scans with best-fit curves based on the distribution shown in b.

(b) Concentration distribution versus the sedimentation

coefficient, s, obtained from the fitting result of a. The peaks appearing at 2.06 and 4.76 S correspond to molecular weights of 25.7 and 103 kDa, respectively.

Figure 3. SAXS profiles of gliadins in distilled water. (a) Solution components at low concentrations, (b) paste-form samples at high concentrations. The curves are vertically shifted for clarity.

Figure 4. (a) Guinier plots for dilute aqueous solutions of gliadins. Scattering profiles are fitted by a three-component Guinier equation (thick lines). Small arrows indicate the points at which the relationship qRg1 = 1.3 holds (Rg1: smallest radius of gyration). (b) Cross-section Guinier plots for aqueous solutions of gliadins. Small arrows indicate the points at which the relationship qRc = 1 holds (Rc: cross-sectional radius of gyration).

Figure 5. Detection of the self-assembled gliadins in solution. Gliadin solutions (0.5, 1, 3, 5, 6, 7, 8, 9, 10 wt% in distilled H2O) were incubated at 25˚C for 60 min in the

21 - Environment ACS Paragon-Plus

Journal of Agricultural and Food Chemistry

absence (-) or presence of 5 mM DTSSP (+). After NEM-treatment for 10 min, the solutions were boiled in Laemmli’s SDS sample buffer without reducing agent. The soluble fraction was separated by centrifugation and then subjected to nonreducing SDS-PAGE (a).

Gliadins from the 7 wt% solution treated with DTSSP were

separated by nonreducing SDS-PAGE on 3-12% gel with Bis-Tris running buffer (b, panel 1), reducing SDS-PAGE (b, panel 2) and 2D-PAGE of nonreducing SDS-PAGE on 3-12% gel with Bis-Tris running buffer and reducing SDS-PAGE (b, panel 3). Proteins were stained with Coomassie Brilliant Blue.

Figure 6.

Schematic illustrations of the nanostructures of gliadin assemblies in

distilled water over a wide range of concentrations. (a) Gliadins are present as watersoluble isolated monomers and a few dimers at 0.025–0.5 wt%. (b) Gliadin molecules self-assemble to form small aggregates (dashed line circles) at 0.5 –15 wt%. The distance between the aggregates is estimated to be ~40 nm. (c) Gliadins begin to form continuous networks at 15–20 wt%.

(d) Gliadin molecules fill the space, but

condensed regions due to density fluctuation (dashed line circles) appear above 30 wt%. The distance between dense domains is estimated to be 14 nm at 40 wt%. (e) Above 50 wt% the density fluctuation almost vanishes, but large aggregates over 100nm in size are formed.

22 - Environment ACS Paragon-Plus

Page 22 of 30

Page 23 of 30

Journal of Agricultural and Food Chemistry

Tables

Table 1. Radii of gyration calculated by fitting with a three-component Guinier equation. conc / wt% Rg1 / nm Rg2 / nm Rg3 / nm 0.025 4.10±0.74 13.7±1.1 33.2±0.9 0.1 5.68±0.20 16.4±0.7 34.9±1.2 0.5 6.46±0.05 16.7±0.1 40.0±0.3

Table 2. Radius of gyration (Rg1), cross-sectional radius of gyration (Rc), and rod length (Λ) calculated from Guinier and cross-section Guinier plots. conc / wt% Rg1 / nm Rc / nm Λ / nm 0.025 4.10±0.74 2.65±0.05 10.8±3.4 0.1 5.68±0.20 2.71±0.01 17.3±0.8 0.5 6.46±0.05 3.59±0.003 18.6±0.2

Table 3. Correlation lengths (L) calculated from Bragg's law. conc / wt% qs / nm-1 1 0.210±0.003 3 0.169±0.002 5 0.165±0.001 10 0.139±0.003 15 0.150±0.004 20 0.144±0.002 25 0.145±0.001 30 0.147±0.002 35 0.304±0.005 40 0.434±0.001 50 0.665±0.002 60 0.741±0.008

23 - Environment ACS Paragon-Plus

L 29.9 ±0.4 37.2 ±0.4 38.0 ±0.2 45.3 ±0.8 41.9 ±1.0 43.4 ±0.5 42.7 ±0.4 40.7 ±0.4 20.6 ±0.3 14.5 ±0.05 9.45 ±0.04 8.48 ±0.09

Journal of Agricultural and Food Chemistry

Figures

Figure 1

24 - Environment ACS Paragon-Plus

Page 24 of 30

Page 25 of 30

Journal of Agricultural and Food Chemistry

Figure 2.

25 - Environment ACS Paragon-Plus

Journal of Agricultural and Food Chemistry

Page 26 of 30

Figure 3

b

2

10

10 wt% 5 3 1 0.5 0.1 0.025

1

10

0

10 Intensity

-1

10

-2

10

-3

15 20 35 25 40 30 50 60 70 wt%

Intensity

a

10

-4

10

-5

10

-6

10

0.1

q / nm

-1

1

0.1

26 - Environment ACS Paragon-Plus

q / nm

-1

1

Page 27 of 30

Journal of Agricultural and Food Chemistry

Figure 4

a

b 0.5 0.1 0.025 wt%

qRc = 1 qI

Intensity

0.5 0.1 0.025 wt%

qRg1 = 1.3 0.05

0.10 2

q / nm

0.15

0.05 0.10 0.15 0.20 0.25 0.30

-2

2

q / nm

27 - Environment ACS Paragon-Plus

-2

Journal of Agricultural and Food Chemistry

Figure 5

28 - Environment ACS Paragon-Plus

Page 28 of 30

Page 29 of 30

Journal of Agricultural and Food Chemistry

Figure 6

a

b

d

e

c

29 - Environment ACS Paragon-Plus

Journal of Agricultural and Food Chemistry

TOC Graphics

15 wt%

Gliadin 10 wt% 0.025

Intensity

70

aqueous solutions 0.1

q / nm

-1

hydrated solids 1

0.1

q / nm

-1

1

30 - Environment ACS Paragon-Plus

Page 30 of 30