Sodium alginate ( ) 0.1% wt – bovine serum albumin ( ) 0.2% wt

8. All tests were conducted in triplicates with freshly prepared samples. Analysis of variance. 148. (ANOVA) was used for data analysis (SPSS 21, Chic...
0 downloads 11 Views 7MB Size
Subscriber access provided by - Access paid by the | UCSB Libraries

Article

Complex Coacervation of Milk Proteins with Sodium Alginate Elham Ghorbani Gorji, Abdul Waheed, Roland Ludwig, Jose Luis Toca-Herrera, Gerhard Schleining, and Sara Ghorbani Gorji J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.7b03915 • Publication Date (Web): 28 Feb 2018 Downloaded from http://pubs.acs.org on March 1, 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.

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 34

Journal of Agricultural and Food Chemistry

ζ= 9.73

3.0 pHopt

ζ= 6.97

Absorbance at 400nm

2.5

ζ= -39.66

2.0 pHφ2 1.5

pHφ1

ζ= -50.76

1.0 0.5

pHC

0.0 0

1

2

3

4

5

6

7

pH

Sodium alginate ( ) 0.1% wt – bovine serum albumin ( ) 0.2% wt

ACS Paragon Plus Environment

8

Journal of Agricultural and Food Chemistry

Complex Coacervation of Milk Proteins with Sodium Alginate

1 2 3

Authors: Elham Ghorbani Gorjia, Abdul Waheedb, Roland Ludwig a, José Luis Toca-Herrerac,

4

Gerhard Schleininga and Sara Ghorbani Gorjid

5

Affiliations:

6

a

7

(BOKU), Vienna, Austria

8

b

9

c

Department of Food Science and Technology, University of Natural Resources and Life Sciences

Faculty of Agriculture, University of Hohenheim, Stuttgart, Germany

Institute for Biophysics, Department of Nanobiotechnology, University of Natural Resources and

10

Life Sciences Vienna (BOKU), Muthgasse 11, A-1190 Vienna, Austria

11

d

12

Australia

The University of Queensland, School of Agriculture and Food Science, Brisbane 4072, QLD,

1 ACS Paragon Plus Environment

Page 2 of 34

Page 3 of 34

Journal of Agricultural and Food Chemistry

13

Abstract

14

Beta-lactoglobulin (BLG) and bovine serum albumin (BSA) coacervate formation with sodium

15

alginate (ALG) was investigated by turbidimetric analysis, zeta potential, particle size, viscosity,

16

transmission electron

17

measurements as a function of pH (1.0-7.0) and, protein:alginate mixing ratio (1:1, 1.5:1 and 2:1,

18

1:0 and 0:1 w/w). Critical pH values of phase transitions for BSA-ALG complexes (pHC, pHɸ1,

19

and pHɸ2) representing the formation of soluble and insoluble complexes of a protein-ALG

20

mixture (2:1) at pH 4.8, pH 4.2 and pH 1.8, respectively. In case of BLG-ALG, Critical pH values

21

(pHC, pHɸ1, and pHɸ2) were found to be 4.8, 4.2 and 1.6, respectively. The pHopt values, expressed

22

by the highest optical density, were pH 2.8 for BSA-ALG and 2.4 for BLG-ALG. TEM and zeta

23

potential results showed that maximum coacervate formation occurred at pH 4.2 for both protein-

24

polysaccharide solutions. The interaction between BLG-ALG and BSA-ALG was spontaneous

25

exothermic at pH 4.2 according to ITC measurements. The findings of this study provide insights

26

to have a thorough understanding about the nature of interactions between milk proteins and ALG

27

and formulate the new applications for food, pharmaceutical, nutraceutical and cosmetics

28

applications.

microscopy (TEM) and isothermal titration calorimetric (ITC)

2 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 34

29

1. INTRODUCTION

30

Proteins and polysaccharides as natural functional ingredients are extensively used in the food,

31

biomedical industries and pharmaceutical 1. The manipulation of the macromolecular interactions

32

between these biopolymers is a crucial element in their application for different purposes

33

Protein-polysaccharide complexation occurs from an electrostatic interaction between the

34

oppositely charged biopolymers in an aqueous phase 5. This attractive interaction can take place in

35

either a single phase (soluble) or in two phase (insoluble) 6. Complex coacervation is the

36

formation of insoluble complexes which leads to a liquid-liquid macroscopic phase separation,

37

with solvent rich and complex rich phases

38

order to decrease the free energy of the system 8. The electrostatically driven formation of

39

complexes, can be influenced by several physicochemical parameters in the system i.e. type of

40

biopolymer, biopolymer concentration, mixing ratio, biopolymers reactive groups, size, pH,

41

charge density, temperature, and ionic buffer strength 9,10.

42

The interaction of proteins and polysaccharide gained considerable attention during the last decade

43

according to their usability in different areas such as food, cosmetic industries and pharmaceutical,

44

where they were used as bioactive delivery devices, fat replacers, gels, emulsion stabilizer, edible

45

films, coatings and for food texturisation 11,12. The effects of biopolymer ratio, total concentration

46

and pH changes were assessed for the β-lactoglobulin-acacia gum system. Being investigated by

47

Weinbreck et al., the effect of biopolymers charge density on their interaction on complex

48

formation between whey protein isolate and two other types of polysaccharides, λ- carrageenan

49

and a lactic bacteria exopolysaccharide lead to a precipitate in the complexation of whey protein

50

isolate- λ- carrageenan because of the high charge density of the polysaccharide 9,13.

51

Alginate is an anionic, linear and water-soluble polysaccharide extracted from three different

52

species of brown algae and is consist of varying amounts of 1–4 linked α-L-guluronic and β-D-

5,7

2–4

.

. The aggregation of soluble complexes occurs in

3 ACS Paragon Plus Environment

Page 5 of 34

Journal of Agricultural and Food Chemistry

53

mannuronic acid residues with pKa values of 3.4 – 4.4. The sequence of monomers and the

54

molecular weight of this biopolymer specifies its physical properties 14,15. Alginate is mainly used

55

in the food, pharmaceutical, and cosmetic industries due to its special characterisation such as

56

biocompatibility, biodegradability, gelling properties, immunogenicity, non-toxicity, and

57

relatively low cost 16. Studies showed that alginate is able to go through complex coacervation or

58

precipitation with several proteins like pea protein, albumin, whey protein, β-lactoglobulin, canola

59

protein, and gelatin A. The formed coacervates and precipitates depend on the type of protein but

60

also pH, total biopolymer concentration, mixing ratio, and ionic strength17,18. Bovine serum

61

albumin (BSA) is a highly soluble globular protein that is found in whey, milk, and plasma and

62

has 585 amino acid residues, and a molecular weight of 66.5 kDa19. It contains a high number of

63

amino acids with acidic side chains (203 aspartates and glutamates) and with basic side chains

64

(164 lysines, arginines and histidines), most of them surface exposed which results in an

65

isoelectric point of 5.0. The secondary structure of this protein is mainly composed of α-helices

66

(66%) with a remaining content of β-sheets turn and side chains (34%) 20. Convenient structural,

67

physicochemical, and biological properties makes BSA a proper model protein for pharmaceutical,

68

cosmetic and food systems

69

acid residues

70

aspartates and glutamates) and with basic side chains (18 lysines, arginines and histidines), most

71

of them surface exposed which results in an isoelectric point of 5.2. This protein is highly

72

functional with thermal aggregation, gelation and surface-active properties 23. These two proteins

73

are frequently used as model proteins in the development of pharmaceutical, nutraceutical,

74

cosmetic, and food formulations 24. They have electro negative hydrophilic domains at pH values

75

above their isoelectric points and an excess of positive charges below pH of isoelectric points and

76

this characteristic make them to be capable of forming complexes with various polyelectrolytes,

22

21

. β-lactoglobulin is a small, globular whey protein with 162 amino

. It contains also a high number of amino acids with acidic side chains (27

4 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

77

especially with potent charge polysaccharides 25. Electrostatic interactions lead the association and

78

the phase behaviour in the protein/polysaccharide complexes. The formation, the structure, and the

79

property of the complex are affected by physicochemical conditions like pH, ionic concentration,

80

intrinsic features, distribution of charges, size, stiffness of polysaccharide chains, etc.9,13.

81

The objectives of current study are to characterize the interactions occurring between the proteins

82

BSA and BLG and the anionic polysaccharide ALG in aqueous solutions. Moreover, the effect of

83

solvent pH and biopolymer mixing ratio on complex formation involving milk proteins with ALG

84

was determined. The interactions between these biopolymers in solution have been studied using

85

analytical techniques, including spectroscopy, size measurements, zeta potential, isothermal

86

titration calorimetry and different microscopy techniques. Each of these techniques will provide

87

some information on the nature of the interaction between these biopolymers in a broad range of

88

pH values.

89 90

2.

91

2.1 Sample preparation

92

β-Lactoglobulin and bovine serum albumin were obtained from Sigma (Germany). Guluronate-

93

rich sodium alginate (Protanal® LF 10/60 Sodium Alginate) was kindly provided by FMC

94

Biopolymers. 0.2% wt BSA and BLG solutions were made separately in Milli Q water (ultrapure

95

water). The solutions were stirred for 30 min and kept in 4°C for 12 hours. 0.1% wt ALG solution

96

was prepared by mixing sodium alginate in Milli Q water and stirring overnight (12h). The

97

mixtures of BSA-ALG and BLG-ALG were prepared by dissolving each protein solutions with

98

ALG solution.

99

measurements and 2:1 for other measurements were prepared.

100

2.2 Determination of critical pH values by spectrophotometry

MATERIAL AND METHODS

Mixing ratios of Protein:ALG 0:1, 1:1, 1.5:1, 2:1 for spectrophotometry

5 ACS Paragon Plus Environment

Page 6 of 34

Page 7 of 34

Journal of Agricultural and Food Chemistry

101

Protein and sodium alginate dispersions at different protein-ALG ratios of 0:1, 1:1, 1.5:1, 2:1 and

102

1:0 were prepared using the stock solutions. The prepared dispersions were acidified gradually

103

(from pH 7 to 6.80, 6.60, 6.40 to 1) by addition of 0.05 M, 0.1 M, 0.2 M and 2 M HCl solution

104

under magnetic stirring at each pH value. The turbidity (optical density) of proteins and sodium

105

alginate dispersions mixtures as a function of pH from ~7 to ~1 was measured using a UV/visible

106

light spectrophotometer (UV 1800 Shimadzu, Japan) at maximum absorbance wave length

107

obtained before for the solutions (~400 nm). Measurements were done in cuvettes (plastic, 1 cm

108

path length) and for the blank reference Milli-Q water was used. ALG, BSA and BLG solutions

109

were used as controls at 0.1% wt concentration. The critical pH values i.e. pHC: development of

110

soluble complexes, pHφ1: development of insoluble complexes, pHopt: maximum optical density,

111

pHφ2: dissolution of complexes (Table 1) were calculated based on the intersection point of two

112

curve tangents.

113

2.3 Zeta potential and size measurements

114

The electrophoretic mobility of samples was detected using a Malvern Zetasizer NS 200 apparatus

115

(Malvern Instruments, Germany) fitted with a DTS1070 sample cell. This instrument applies a

116

photon correlation spectroscopy (PCS) technique to measure particle size in constant random

117

thermal, or Brownian motion. Before the measurement, the sample solutions were diluted to 100

118

to 500 times with ultra-pure water to avoid the multiple scattering effects. All measurements were

119

done in triplicate at 25°C and the results were expressed as the mean of three measurements.

120

2.4 Transmission electron microscopy (TEM) and optical microscopic images

121

Images of proteins-polysaccharide interaction were captured with JEOL JEM-2100F (Model EM-

122

20014) transmission electron microscope at 7800X. The samples were absorbed on the surface of

123

a carbon coated copper grid and fixed with a negative staining method (with 1 % uranyl acetate in

124

milli Q water). Due to the high magnification of TEM, the solutions were diluted to 100 times 6 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

125

with milli Q water to observe the coacervates while in case of microscopic analysis, the solutions

126

were used without any dilution. Furthermore, ALG-proteins images were obtained by using an

127

optical microscope (Olympus BX51).

128

2.5 Isothermal titration calorimetry (ITC)

129

The enthalpic and entropic changes of BLG (0.2%)-ALG (0.1%) and BSA (0.1%)-ALG (0.1%)

130

interactions at 25°C was measured using conducted ITC 200 Micro calorimeter (Northampton,

131

MA, USA). Separate solutions of BLG, BSA and ALG were made by stirring gently for 8h in 5

132

mM sodium citrate buffer (pH 4.2). Sodium citrate buffer was used to eliminate

133

protonation/deprotonation effects during the measurement. Sodium citrate buffer was for control

134

experiments. Before the measurements, the solutions were degassed for 8 min under vacuum. This

135

was done by a device provided by the ITC apparatus. ALG solution was put into the injector

136

stirrer syringe (40 µL). Portions of 2 µL (except for the first injection that was 0.4 ml) of ALG

137

solution (0.1% (w/w)) were injected sequentially into the titration cell initially containing either

138

BLG or BSA solutions or the pure buffer. Each injection took 1sec with extra equilibration time

139

between consecutive injections of 150sec

140

blank titration of ALG solution into citrate buffer was done and the measured heat was subtracted

141

from experiments with proteins present to obtain correct enthalpy change. The change in enthalpy

142

per gram ALG (kcal/g), injected into the reaction cell was reported as results. Thermodynamic

143

parameters including binding stoichiometry (N), affinity constant (Ka), enthalpy (∆H) and entropy

144

(∆S) changes were calculated by iterative curve fitting of the binding isotherms using the one-

145

binding-site model (assuming one independent binding site exists for any positively charged

146

amino acids) and was plotted against the protein:ALG molar ratio.

147

2.6 Statistical analysis

18

. The speed of stirrer was set at 750 rpm. First the

7 ACS Paragon Plus Environment

Page 8 of 34

Page 9 of 34

Journal of Agricultural and Food Chemistry

148

All tests were conducted in triplicates with freshly prepared samples. Analysis of variance

149

(ANOVA) was used for data analysis (SPSS 21, Chicago, IL, USA). When F-values were

150

significant (P < 0.05) in ANOVA, Duncan’s multiple range test was used to compare treatment

151

means.

152

3. RESULTS AND DISCUSSIONS

153

3.1 pH-dependent complexation of sodium alginate/protein by spectrophotometry

154

Turbidimetric measurements have been broadly employed as a practical tool to monitor the

155

complex coacervation process in various protein polysaccharide systems

156

measurements as a function of pH was performed to evaluate the complexation between ALG

157

(0.1% wt) with proteins (BSA and BLG) at different protein concentrations (0.1, 0.15 and 0.2%

158

wt). Under the same conditions, homogenous biopolymer solutions of ALG, BSA and BLG

159

(0.1%wt) were investigated separately as control samples.

160

Turbidity of the solution can be changed by the change in size of aggregates and mass; thus any

161

change in turbidity could be caused by formation and dissociation of protein-ALG complexes.

162

Accordingly, turbidimetric measurements can demonstrate the complex coacervation process

163

between protein and polysaccharide 29.

164

ALG solution alone, stayed transparent from pH 7 to 1 Particles formed were not big enough to

165

scatter light strongly because of electrostatic repulsion between ALG solution (Figure 1)18.

166

Whereas, at lower pH values, there was a minor increase in turbidity, that could be due to

167

protonation of ALG carboxyl groups and consequently the degree of the electrical charge on the

168

ALG molecules reduced18. The BSA and BLG solutions (without ALG) had a smaller peak,

169

individually, at pH ranged about from 4 to 5.2 that was a result of aggregation and the following

170

precipitation of BLG and BSA (Figure 1). The electrostatic repulsion between proteins decreases

171

around their pI and initiates aggregates pI 30,31. 8 ACS Paragon Plus Environment

26–28

. Turbidimetric

Journal of Agricultural and Food Chemistry

172

At pH 7.00, both BLG-ALG and BSA-ALG dispersions were clear, indicating no absorbance at

173

400 nm (Figure 1&2). Due to acid titration of BSA-ALG and BLG-ALG dispersions, the four

174

phases of turbidity changes were indicated. The presence of ALG to each protein solution, at a 1:1

175

protein-ALG mixing ratio caused a significant shift in the initial rise in turbidity from comparing

176

to each protein pure solution. The absorbance values were low and constant until pHC. Later on

177

the absorbance increased, implying attractive interaction between ALG and proteins 32. At this pH

178

the formation of soluble and stable complexes is initiated

179

second boundary pH value (pHϕ1), the interaction between the two macromolecules became

180

stronger and phase separation occurred. The absorbance value continued to increase to maximum

181

value at pHopt. From this phase forward the absorbance values decrease abruptly with lowering pH

182

that is due to the sedimentation of particles in protein-ALG dispersions. This was followed by an

183

increase in optical density during pH titration and then the curve went down to pHɸ2. Ultimately,

184

in very low pH values because of the low charges of ALG chains, as well as the repulsion forces

185

between the positively charged proteins, the protein-ALG coacervates could redissolve into

186

soluble complexes, or even into non-interacting protein molecules and polysaccharide chains 4.

187

When the concentration of protein was increased from 0.1 to 0.15 and 0.2%wt in the solution, a

188

slightly higher turbidity curve was found comparing to the lower concentration of protein. It was

189

hypothesized that the presence of a highly charged polysaccharide at a concentration of 0.1% (wt)

190

caused adequate electrostatic repulsion to hinder protein aggregation at lower protein

191

concentration, 0.1% (wt), and small reticence of structure formation can contribute to the minor

192

reduction in maximum intensity. When the protein: ALG ratio increased (i.e. protein increased and

193

ALG levels decreased), delay of aggregate growth was less substantial where the shifting of pH-

194

dependent turbidity profiles to more acidic pH was reduced. With the addition of ALG, the shift to

195

lower pH values suggest structures with increased stability at low pH values relative to protein 9 ACS Paragon Plus Environment

9,32

. By decreasing pH passing the

Page 10 of 34

Page 11 of 34

Journal of Agricultural and Food Chemistry

196

aggregates in the absence of ALG. Klemmer et al. observed similar trends for mixtures of pea

197

protein isolate and alginate polysaccharides with biopolymer ratio of 1:1 to 20:133.

198

3.2 Size measurement of protein/sodium alginate

199

Particle size analysis is mostly used for monitoring formation and growth of electrostatic

200

complexes between protein and polysaccharide. Accordingly, during complex coacervation the

201

particle size of BSA-ALG and BLG-ALG was examined to provide further insight into their

202

interaction. Protein solutions at all pH values had a particle size ranging from 0.5 to 20 nm (Figure

203

2 (c) & (d)). The particle size measurements show that when the pH is reduced to 4.8 the proteins

204

tended to form large aggregates. It has been revealed that lowering pH to less than pI of the

205

proteins (4.9 for BSA and 5.2 for BLG) leads to aggregation of proteins 34. Hence the presence of

206

larger particles causing high peaks is negligible

207

positive charges on proteins so the proteins exist in the form of individual molecules again as a

208

result of repulsion forces. Hence, large aggregates are changed to smaller particles. Therefore, at

209

pH 1.8 the particles are smaller for both proteins.

210

Three separate peaks were detected in BSA-ALG solution, at pH 4.8 (Figure 2a). The maximum

211

particle diameters of 33 nm, 12 nm, and 4 nm, respectively. The 4 nm peak might be assigned to

212

BSA molecules, whereas according to ALG size reported by Motwani et al. bigger peak with

213

diameters of 33 nm and 12 nm might be from ALG suggesting no or small interactions between

214

ALG and BSA at pH 4.8 36. In case of BLG-ALG, at the same pH (4.8), two peaks were found at

215

47 nm and 17 nm. It could be due to polymer interaction between ALG and BLG or self-

216

association of biopolymers in solution changing the particle (size) distribution of both ALG and

217

BLG.37. As the pH decreases to 4.2 the size distribution develops into a more homogenous pattern

218

comparing to higher pH values, suggesting the development of complexes between proteins and

219

ALG.

8,35

. However, lowering pH develops more

10 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

220

With more decrease in pH up to 2.8 in BSA-ALG, four different peaks were observed. The

221

particle size ranged from o.6 nm to more than 1000 nm but the percentage of smaller particles was

222

touched to maximum number (about 8%) at pHopt while in case of BLG-ALG, smaller particles

223

reached a maximum number of 22% (at pHopt) as shown in Figure 2a and 2b. This result was in

224

accordance with the development of insoluble complexes. With the decrease of pH, the proteins

225

and ALG molecules exposed more binding sites for interaction, which generates the development

226

of insoluble complexes. At pHɸ2, the insoluble complexes initiated to dissolve and dissociate

227

because of advanced protonation of carboxyl groups on the ALG structure 8. In BSA-ALG

228

solution at pH 1.8, three peaks were seen. The largest one at around 100nm could be contributed

229

to ALG precipitates that occurs in pH values lower than pKa of ALG 38. The smaller peaks appear

230

to be the particles caused by degradation of protein in low pH value. It was suggested that there

231

were no interactions between BSA and ALG because of the protonation of anionic groups i.e.

232

carboxyl groups on polysaccharide structure thus, similar net charges were carried by both

233

biopolymers, at this pH value (Fig. 2a) 8. The rise in the particle size of ALG-proteins mixtures at

234

pHɸ2 could be due to the denaturation and unfolding of BSA and BLG at low pH values 39. Liu et

235

al. (2015) found the similar results in flaxseed gum and BSA solutions at different pH values (5.6

236

to 2.0) 8.

237

3.3 Zeta potential of protein-sodium alginate

238

The formation of insoluble and soluble protein-ALG complexes is generally initiated by by the

239

electrostatic attractions between the two biopolymers 6. For a better understanding of electrostatic

240

interactions between ALG and proteins in different pH values the electrokinetic potential at the

241

slip plane of moving particles, was measured that is zeta potential. The zeta potential of ALG

242

solution was negative in entire pH range investigated, which decreased from of -19.05 to -46.24

11 ACS Paragon Plus Environment

Page 12 of 34

Page 13 of 34

Journal of Agricultural and Food Chemistry

243

(pH from 4.80 to 2.80). The decrease in pH value caused the ionization of carboxylate residues on

244

the ALG molecular structure 40.

245

In cases of protein suspensions, as shown in Table 1, during the acid titration, the zeta potential

246

value was increased through protonation of carboxyl groups and protein amine 6. The pH, at which

247

zeta potential reaches zero, is often considered as isoelectric point (pI). However, the pI of both

248

proteins were slightly lower than what was suggested previously due to the use of minimum

249

concentration of these proteins in the solution and the charge contribution of some impurities in

250

proteins

251

the reason that proteins were below their pI and accordingly had a net positive charge. Hosseini et

252

al. (2013) found similar results by using BLG in higher concentration 35.

253

Factors such as polysaccharide type and concentration affect the absolute value of the zeta-. ALG

254

has a higher negative charge as compared to many other polysaccharides. It should be noted that

255

the zeta-potential of the mixtures represents the net zeta value of protein-ALG complexes and

256

non-interacting individual biopolymers

257

showed an increasing trend in general with decreasing pH. Zeta potentials at low pH values

258

indicate that the strongest protein-ALG interactions showed in spectrophotometry occurred when

259

the electrical charge of the mixtures was nearly neutralized. Phase separation can appear with a

260

negative total charge therefore, at phase separation neutrality is not a requirement and a generic

261

rule for protein-polyanionic polymers 43. Beyond these points, the zeta-potential values remained

262

rather constant at limiting values reflecting an excess of polysaccharide. As a result, negatively

263

charged polysaccharide molecules associated with positively charged protein and drove charge

264

reversal. The zeta-potential profile of the protein-ALG mixed systems exhibited a similar trend

265

with lower intensity.

41

. The zeta potential of the protein suspension was positive at pHɸ1, pHopt and pHɸ2, for

35,42

. The zeta-potential of BLG-ALG and BSA-ALG also

12 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

266

Complexation between BSA-ALG and BLG-ALG- at pH 4.8 and 4.2 caused low negative zeta

267

potential values between ~-40 mV - -50 mV (table 1) that makes it a stable system since colloidal

268

particles with zeta potential more than +30 mV and more negative than -30 mV are normally

269

considered stable 31.

270

3.4 Transmission electron microscopy and light microscopic images

271

Optical microscopy was used to study the effect of pH on BSA-ALG and BLG-ALG complexes at

272

protein-ALG ratio of 2:1 (w/w) with a total biopolymer concentration of 0.3% (w/w). The images

273

revealed that microstructures of dispersions were different at each critical pH values. At pH 4.8,

274

there were less number of coacervates as shown in Figure 3a, 4e, 5a and 6e while at pH 4.2

275

(Figures 4b, 4f, 5b and 6f), maximum coacervation with minimum particle aggregates was visible.

276

The particle size of coacervates ranged from 150 to 350 nm in both complexes (BSA-ALG and

277

BLG-ALG complexes). Gelatin and cashew gum coacervation with a ratio of 1:2.5 had the

278

coacervates about 20 µm at pH 4 to 4.2 44.

279

More aggregates were observed at pHopt due to higher turbidity while at pHɸ2, less number of

280

aggregates were visible due to the dissociation of the complexes (Figure 3-6 c, d, g and h). The

281

size of the aggregates in both protein complexes with ALG ranged from 0.20 to 200 µm at pHopt

282

and pHɸ2. These outcomes were also justified by the results of particle size and turbidity

283

measurement.

284

In case of ALG-BSA at pH 4.2, clear black coloured particles with a diameter around 0.4 to 1 µm

285

(Figure 3b) were seen. These particles are expected to be BSA aggregates, due to their irregular

286

shape. On the same image, the smaller white particles were visible with a diameter about 150 nm.

287

Due to the round clear structure of these particles, they could be coacervates 45. However, in case

288

of BLG-ALG, at the same pH value, the black aggregates are larger (0.4 to 3 µm) but the

289

coacervates are smaller than BSA-ALG (Figure 3 b) This could be because of the low molecular 13 ACS Paragon Plus Environment

Page 14 of 34

Page 15 of 34

Journal of Agricultural and Food Chemistry

46

290

weight of BLG than BSA

. Considering our results at pH 4.2 for BSA-ALG and BLG-ALG, it

291

could be used for further application such as encapsulation due to maximum coacervate

292

production.

293

Figures 4c, d and Figure 6c, d show an increase in particle polydispersity. A large number of

294

BSA-ALG and BLG-ALG particles at pHopt and pHɸ2 displayed larger aggregates than at pHC and

295

pHɸ1 indicating the maximum number of complexes at pHopt due to highly charged biopolymers

296

and precipitation of biopolymers at pHɸ2. These results are in accordance with our zeta potential

297

results showing a low positive charge for protein-ALG complexes at these two pH values

298

revealing a dispersion with unstable particles and zeta potential lower than +30 mV 31.

299

3.5 Isothermal titration calorimetry (ITC)

300

Isothermal titration calorimetry is a useful technique to assess energetic and binding parameters in

301

the complex coacervation process by titration of one biopolymer with another in a reaction

302

chamber and measuring the heat absorbed or released

303

resulting from the injections of ALG dispersions into protein (BSA and BLG) solutions at pH 4.2

304

and 25oC are shown in Figure 7 and. 8. In case of both BLG-ALG and BSA-ALG, the area under

305

each peak showed the heat exchange within the cell having BLG after each ALG injection.

306

Initially, a sequence of strong successive exothermic peaks of decreasing intensity was initially

307

observed. With more injections, the released binding energy continuously declined and reached to

308

a state of thermodynamic stability after the 32nd injection for BLG-ALG when a molar ratio of

309

0.75 was reached (Figure 7) and 36th injection for BSA-ALG when a molar ratio of 0.065 was

310

reached (Figure 8).

311

Exothermicity is imparted to nonspecific electrostatic neutralization of opposite charges carried by

312

two biopolymers revealing an enthalpic contribution of complex coacervation

313

decrease in exothermicity could be due to a reduction in free protein molecules remaining in the

18

. The heat flow versus time profiles

14 ACS Paragon Plus Environment

47,48

. However, the

Journal of Agricultural and Food Chemistry

314

reaction chamber after injections resulted in reduction in released energy. Hosseini et al. and

315

Hadian et al. reported similar exothermic sequence for BLG interaction with ALG and Persian

316

gum 18,49.

317

Thermodynamic changes during titration were figured out by measuring binding isotherms fitted

318

using a binding model for one binding site and processed by integrating the isotherm peaks and

319

subtraction of heat of dilution of ALG into buffer solution (Figure 7 and 8, lower panel). As a

320

binding site a positively charged amino acid (at the pI) on either BLG or BSA was assumed. The

321

concentration of binding sites was calculated by multiplying the molar concentration of the

322

number of their basic amino acids. Binding stoichiometry (N), affinity constant (Ka), enthalpy

323

(∆H) and entropy contributions (∆S) as thermodynamic parameters were calculated for the

324

interaction of BSA-ALG and BLG-ALG (Table 2).

325

The binding stoichiometry (N) of BLG-ALG was calculated using charge molar ratio to be

326

0.00521 ± 0.0007, suggesting that BLG became saturated with 0.00521 molar charges of ALG

327

while in case of BSA-ALG, 0.00852 molar chargers of ALG were required to bind one molecule

328

of BSA. This equals to the molar ratios of BLG: ALG = 0.784 µM:0.017 µM and BSA: ALG =

329

0.485 µM:0.017 µM.

330

The binding constant (Ka) values point out that the binding of BSA and BLG to ALG can be

331

considered as a high affinity, however in case of BSA it is almost two times bigger than BLG. At

332

pH 4.2, BSA carried highly opposite charge, then more binding sites are available for the BSA-

333

ALG interaction, whereas in BLG-ALG, the Ka value was found to be lower. That shows less

334

number of sites for the interaction of ALG with BLG at the same pH.

335

The binding enthalpies were negative and favourable. Weak complexes are attributed by a

336

negative enthalpy as a result of electrostatic attraction whereas the counter-ion release entropy

337

plays an important role when the interaction became stronger 50. 15 ACS Paragon Plus Environment

Page 16 of 34

Page 17 of 34

Journal of Agricultural and Food Chemistry

338

In soy protein isolate-chitosan coacervation at pH 6.0, the interaction was driven by both enthalpy

339

and entropy, showing that the interaction was a combination of changes in the electrostatic

340

interactions between the charged groups and entropy of the release of counter-ions 51.

341

4.

342

Sodium alginate, an anionic polysaccharide formed complex coacervates with two milk proteins,

343

β-lactoglobulin and bovine serum albumin through electrostatic interactions under specific

344

conditions.

345

Insoluble or soluble complexes between milk proteins ALG can be formed as a function of

346

solution pH and biopolymers mixing ratio. Protein-ALG ratio of 2:1, were selected for further

347

analysis. Formation of soluble coacervates initiated in the range between pHC and pHɸ1 due to

348

binding of anionic ALG to cationic patches on milk protein surfaces. Insoluble complex formation

349

occurred in the range pHɸ1 and pHɸ2 where ALG and proteins were oppositely charged, inducing

350

strong electrostatic attractive interactions.

351

According to ITC results at pH 4.2 for BSA-ALG and BLG-ALG exothermic heat transfer was

352

observed. The binding enthalpies for both systems were negative and favourable. Because of more

353

positive charges on BSA at pH 4.2 the binding constant is almost two times bigger than BLG.

354

Thus, there were more binding sites available for BSA-ALG interaction which is in accordance to

355

zeta potential results at pH 4.2 showing a higher positive charge for BSA comparing to BLG.

356

Results from current research indicated that electrostatic attractive interaction established complex

357

coacervate formation between milk proteins and ALG. This study provides fundamental

358

background knowledge to suggest the implementation of milk proteins-ALG interactions in

359

biomaterial, food, cosmetic and pharmaceutical products.

CONCLUSION

360 361

ACKNOWLEDGMENTS 16 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

362

The authors thank Claudia König for assistance with the TEM measurements. This project was

363

partially financed by the IGS BioNanoTech graduate school (Bundesministerium für Bildung und

364

Forschung (BMBF), Austria).

365 366

REFERENCES

367 368

(1)

369 370

Tolstoguzov, V. Protein-polysaccharide interactions. In FOOD SCIENCE AND TECHNOLOGY-NEW YORK-MARCEL DEKKER; 1997; pp 171–198.

(2)

Jun-xia, X.; Hai-yan, Y.; Jian, Y. Microencapsulation of sweet orange oil by complex

371

coacervation with soybean protein isolate/gum Arabic. Food Chem. 2011, 125 (4), 1267–

372

1272.

373

(3)

Ron, N.; Zimet, P.; Bargarum, J.; Livney, Y. D. Beta-lactoglobulin-polysaccharide

374

complexes as nanovehicles for hydrophobic nutraceuticals in non-fat foods and clear

375

beverages. Int. Dairy J. 2010, 20 (10), 686–693.

376

(4)

377 378

polysaccharide interactions. Soft Matter 2008, 4 (5), 932–942. (5)

379 380

(6)

385

Yang, Y.; Anvari, M.; Pan, C. H.; Chung, D. Characterisation of interactions between fish gelatin and gum arabic in aqueous solutions. Food Chem. 2012, 135 (2), 555–561.

(7)

383 384

Schmitt, C.; Turgeon, S. L. Protein/polysaccharide complexes and coacervates in food systems. Adv. Colloid Interface Sci. 2011, 167 (1–2), 63–70.

381 382

Dickinson, E. Interfacial structure and stability of food emulsions as affected by protein–

Dubin, P.; Bock, J.; Davis, R.; Schulz, D.; Thies, C. Macromolecular complexes in chemistry and biology; 2012.

(8)

Liu, J.; Shim, Y. Y.; Wang, Y.; Reaney, M. J. T. Intermolecular interaction and complex coacervation between bovine serum albumin and gum from whole flaxseed (Linum 17 ACS Paragon Plus Environment

Page 18 of 34

Page 19 of 34

Journal of Agricultural and Food Chemistry

386 387

usitatissimum L.). Food Hydrocoll. 2015, 49, 95–103. (9)

388 389

Weinbreck, F. Complexation of whey proteins with carrageenan. J. Agric. Food Chem. 2004, 52 (11), 3550–3555.

(10)

Hosseini, S. M.; Hosseini, H.; Mohammadifar, M. A.; Mortazavian, A. M.; Mohammadi,

390

A.; Khosravi-Darani, K.; Shojaee-Aliabadi, S.; Dehghan, S.; Khaksar, R. Incorporation of

391

essential oil in alginate microparticles by multiple emulsion/ionic gelatin process. Int. J.

392

Biol. Macromol. 2013, 62, 582–588.

393

(11)

394 395

coacervates. Curr. Opin. Colloid Interface Sci. 2007, 12 (4–5), 166–178. (12)

396 397

(13)

(14)

George, M.; Abraham, T. E. Polyionic hydrocolloids for the intestinal delivery of protein drugs: Alginate and chitosan — a review. J. Control. Release 2006, 114 (1), 1–14.

(15)

402 403

Weinbreck, F.; Nieuwenhuijse, H.; Robijn, G. W.; De Kruif, C. G. Complex Formation of Whey Proteins: Exocellular Polysaccharide EPS B40. Langmuir 2003, 19 (22), 9404–9410.

400 401

Weinbreck, F. CHAPTER 2 Complex Coacervation of Whey Protein and Gum Arabic; 2003; Vol. 4, pp 293–303.

398 399

Turgeon, S. L.; Schmitt, C.; Sanchez, C. Protein-polysaccharide complexes and

Skj, G.; Grasdalen, H.; Larsen, B. Monomer sequence and acetylation pattern in some bacterial alginates. Carbohydr. Res. 1986, 154 (1), 239–250.

(16)

Kim, W.; Chung, H.; Shin, I.; Yam, K.; Chung, D. Characterization of calcium alginate and

404

chitosan-treated calcium alginate gel beads entrapping allyl isothiocyanate. Carbohydr.

405

Polym. 2008, 71 (4), 566–573.

406

(17)

Ducel, V.; Richard, J.; Saulnier, P.; Popineau, Y.; Boury, F. Evidence and characterization

407

of complex coacervates containing plant proteins: Application to the microencapsulation of

408

oil droplets. Colloids Surfaces A Physicochem. Eng. Asp. 2004, 232 (2–3), 239–247.

409

(18)

Hosseini, S. M. H.; Emam-Djomeh, Z.; Razavi, S. H.; Moosavi-Movahedi, A. A.; Saboury, 18 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

410

A. A.; Atri, M. S.; Van der Meeren, P. beta-Lactoglobulin-sodium alginate interaction as

411

affected by polysaccharide depolymerization using high intensity ultrasound. Food

412

Hydrocoll. 2013, 32 (2), 235–244.

413

(19)

Zhao, Y.; Li, F.; Carvajal, M. T.; Harris, M. T. Interactions between bovine serum albumin

414

and alginate: An evaluation of alginate as protein carrier. J. Colloid Interface Sci. 2009, 332

415

(2), 345–353.

416

(20)

Gelamo, E. L.; Silva, C. H. T. P.; Imasato, H.; Tabak, M. Spectroscopic studies on the

417

interaction of bovine ( BSA ) and human ( HSA ) serum albumins with ionic surfactants.

418

Biochim. Biophys. Acta (BBA)-Protein Struct. Mol. Enzymol. 2000, 1594 (1), 2255–2271.

419

(21)

420 421

Galazka, V. B.; Dickinson, E.; Ledward, D. a. Effect of high pressure on the emulsifying behaviour of β-lactoglobulin. Food Hydrocoll. 1996, 2 (10), 213–219.

(22)

Matalanis, A.; Jones, O. G.; McClements, D. J. Structured biopolymer-based delivery

422

systems for encapsulation, protection, and release of lipophilic compounds. Food

423

Hydrocoll. 2011, 8 (25), 1865–1880.

424

(23)

Das, K. .; Kinsella, J. . Effect of heat denaturation on the adsorption of β-lactoglobulin at

425

the oil/water interface and on coalescence stability of emulsions. J. Colloid Interface Sci.

426

1990, 2 (139), 551–560.

427

(24)

428 429

Li, Y.; Huang, Q. Influence of protein self-association on complex coacervation with polysaccharide: A Monte Carlo study. J. Phys. Chem. B 2013, 117 (9), 2615–2624.

(25)

Jones, O. G.; McClements, D. J. Recent progress in biopolymer nanoparticle and

430

microparticle formation by heat-treating electrostatic protein-polysaccharide complexes.

431

Adv. Colloid Interface Sci. 2011, 167 (1–2), 49–62.

432 433

(26)

Stone, A. K.; Nickerson, M. T. Formation and functionality of whey protein isolate-(kappa, iota-, and lambda-type) carrageenan electrostatic complexes. Food Hydrocoll. 2012, 27 19 ACS Paragon Plus Environment

Page 20 of 34

Page 21 of 34

Journal of Agricultural and Food Chemistry

434 435

(2), 271–277. (27)

Gorji, S. G.; Gorji, E. G.; Mohammadifar, M. A. Characterisation of gum tragacanth

436

(Astragalus gossypinus)/sodium caseinate complex coacervation as a function of pH in an

437

aqueous medium. Food Hydrocoll. 2014, 34, 161–168.

438

(28)

439 440

Lee, A. C.; Hong, Y. H. Coacervate formation of alpha-lactalbumin-chitosan and betalactoglobulin-chitosan complexes. Food Res. Int. 2009, 42 (5–6), 733–738.

(29)

Ru, Q.; Wang, Y.; Lee, J.; Ding, Y.; Huang, Q. Turbidity and rheological properties of

441

bovine serum albumin / pectin coacervates : Effect of salt concentration and initial protein /

442

polysaccharide ratio. Carbohydr. Polym. 2012, 88 (3), 838–846.

443

(30)

Jones, O.; Decker, E. A.; McClements, D. J. Thermal analysis of β-lactoglobulin complexes

444

with pectins or carrageenan for production of stable biopolymer particles. Food Hydrocoll.

445

2010, 24 (2–3), 239–248.

446

(31)

Mounsey, J. S.; O’Kennedy, B. T.; Fenelon, M. A.; Brodkorb, A. The effect of heating on

447

β-lactoglobulin-chitosan mixtures as influenced by pH and ionic strength. Food Hydrocoll.

448

2008, 22 (1), 65–73.

449

(32)

Ghorbani Gorji, S.; Ghorbani Gorji, E.; Mohammadifar, M. A.; Zargaraan, A.

450

Complexation of sodium caseinate with gum tragacanth: Effect of various species and

451

rheology of coacervates. Int. J. Biol. Macromol. 2014, 67, 503–511.

452

(33)

Klemmer, K. J.; Waldner, L.; Stone, A.; Low, N. H.; Nickerson, M. T. Complex

453

coacervation of pea protein isolate and alginate polysaccharides. Food Chem. 2012, 130 (3),

454

710–715.

455

(34)

456 457

O’Kennedy, B. T.; Mounsey, J. S.; Murphy, F.; Duggan, E.; Kelly, P. M. Factors affecting the acid gelation of sodium caseinate. Int. Dairy J. 2006, 16 (10), 1132–1141.

(35)

Hosseini, S. M. H. Complex coacervation between β-lactoglobulin and anionic 20 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

458

polysaccharides and its potential application for nanoencapsulation of hydrophobic

459

functional components, Doctoral dissertation, Ghent University, 2013.

460

(36)

Motwani, S. K.; Chopra, S.; Talegaonkar, S.; Kohli, K.; Ahmad, F. J.; Khar, R. K.

461

Chitosan-sodium alginate nanoparticles as submicroscopic reservoirs for ocular delivery:

462

Formulation, optimisation and in vitro characterisation. Eur. J. Pharm. Biopharm. 2008, 68

463

(3), 513–525.

464

(37)

465 466

bovine β -lactoglobulin dimer and octamer. Protein Sci. 2009, 12 (11), 2404–2411. (38)

467 468

Gottschalk, M.; Nilsson, H.; Roos, H.; Halle, B. Protein self-association in solution: The

Shinde, U.; Nagarsenker, M. Characterization of gelatin-sodium alginate complex coacervation system. Indian J. Pharm. Sci. 2009, 71 (3), 313.

(39)

Vinayahan, T.; Williams, P. a; Phillips, G. O. Electrostatic Interaction and Complex

469

Formation between Gum Arabic and Bovine Serum Albumin. Biomacromolecules 2010, 11

470

(12), 3367–3374.

471

(40)

Warrand, J.; Michaud, P.; Picton, L.; Muller, G.; Courtois, B.; Ralainirina, R.; Courtois, J.

472

Large-scale purification of water-soluble polysaccharides from flaxseed mucilage, and

473

isolation of a new anionic polymer. Chromatographia 2003, 58 (5), 331–335.

474

(41)

Salgin, S.; Salgin, U.; Bahadir, S. Zeta potentials and isoelectric points of biomolecules:

475

The effects of ion types and ionic strengths. Int. J. Electrochem. Sci. 2012, 7 (12), 12404–

476

12414.

477

(42)

Harnsilawat, T.; Pongsawatmanit, R.; McClements, D. J. Characterization of β-

478

lactoglobulin–sodium alginate interactions in aqueous solutions: A calorimetry, light

479

scattering, electrophoretic mobility and solubility study. Food Hydrocoll. 2006, 20 (5),

480

577–585.

481

(43)

Aberkane, L.; Jasniewski, J.; Gaiani, C.; Scher, J.; Sanchez, C. Thermodynamic 21 ACS Paragon Plus Environment

Page 22 of 34

Page 23 of 34

Journal of Agricultural and Food Chemistry

482

characterization of acacia gum-beta-lactoglobulin complex coacervation. Langmuir 2010,

483

26 (15), 12523–12533.

484

(44)

Gomez-Estaca, J.; Comunian, T. A.; Montero, P.; Ferro-Furtado, R.; Favaro-Trindade, C. S.

485

Encapsulation of an astaxanthin-containing lipid extract from shrimp waste by complex

486

coacervation using a novel gelatin-cashew gum complex. Food Hydrocoll. 2016, 61, 155–

487

162.

488

(45)

Newton, D.; McMullen, J. Characteristics of medicated and unmedicated microglobules

489

recovered from complex coacervates of gelatin‐acacia. J. Pharm. Sci. 1977, 66 (9), 1327–

490

1330.

491

(46)

Schmitt, C.; Sanchez, C.; Despond, S.; Renard, D.; Thomas, F.; Hardy, J. Effect of protein

492

aggregates on the complex coacervation between β-lactoglobulin and acacia gum at pH 4.2.

493

Food Hydrocoll. 2000, 14 (4), 403–413.

494

(47)

Girard, M.; Sanchez, C.; Laneuville, S. Associative phase separation of β-

495

lactoglobulin/pectin solutions: a kinetic study by small angle static light scattering. Colloids

496

Surfaces B Biointerfaces 2004, 35 (1), 15–22.

497

(48)

Schmitt, C.; Palma Da Silva, T.; Rami-Shojaei, C. B. S.; Frossard, P.; Kolodziejczyk, E.;

498

Leser, M. E. Effect of time on the interfacial and foaming properties of /??-

499

lactoglobulin/acacia gum electrostatic complexes and coacervates at pH 4.2. Langmuir

500

2005, 21 (17), 7786–7795.

501

(49)

Hadian, M.; Hosseini, S. M. H.; Farahnaky, A.; Mesbahi, G. R.; Yousefi, G. H.; Saboury,

502

A. A. Isothermal titration calorimetric and spectroscopic studies of ??-lactoglobulin-water-

503

soluble fraction of Persian gum interaction in aqueous solution. Food Hydrocoll. 2016, 55,

504

108–118.

505

(50)

Ou, Z.; Muthukumar, M. Entropy and enthalpy of polyelectrolyte complexation : Langevin 22 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

506

dynamics simulations Entropy and enthalpy of polyelectrolyte complexation : Langevin

507

dynamics. J. Chem. Phys. 2014, 154902 (2006).

508

(51)

Yuan, Y.; Kong, Z. Y.; Sun, Y. E.; Zeng, Q. Z.; Yang, X. Q. Complex coacervation of soy

509

protein with chitosan: Constructing antioxidant microcapsule for algal oil delivery. LWT -

510

Food Sci. Technol. 2017, 75, 171–179.

511

23 ACS Paragon Plus Environment

Page 24 of 34

Page 25 of 34

Journal of Agricultural and Food Chemistry

TABLES Table 1. Zeta potential (mV), of ALG 0.1-BSA 0.2 (0.3%), ALG 0.1-BLG 0.2 (0.3%), BSA (0.2%) and BLG (0.2%) solutions at critical pH values pH

sample

BSA-ALG

BLG-ALG

BSA

BLG

ALG

pHC: 4.8

-50.76±1.07 a

-48.77±0.65a

-0.21±0.16a

-1.00±0.52a

-19.05±0.85

pHɸ2: 4.2

-39.66±1.60 b

-39.73±1.07b

11.97±2.84b

7.21±1.15b

-32.10±0.77

pHopt: 2.8

9.73±2.47 c

-

30.63±2.37b

13.57±1.66c

-46.24±2.31

pHopt: 2.4

-

22.43±3.40d

-

-

-

pHɸ1: 1.8

6.97±1.75 c

15.47±1.93c

6.98±1.66b

pHɸ1: 1.6

-

-

-

2.01±0.38c

-

* a–d Means with different letters within each column differed significantly (p < 0.05).

Table 2. Thermodynamic parameters of binding between BSA (0.1%) and BLG (0.2%) with ALG (0.1%) at 25°C in 5 mM sodium citrate buffer (pH 4.20) Parameters

N

Ka (M-1)

∆H (cal/mol)

∆S (cal/mol*°C)

BSA-ALG

0.00852±0.002

7.27*103±1.46*103

-6.076*104±1.78*104

-186

BLG-ALG

0.00521 ± 0.0007

3.99*103± 2.13*103

-9.845*104± 2.124*105

-314

Samples

24 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figures

3,0 (a)

pH opt

pHopt

Absorbance at 400nm

2,5

2,0

pH2φ2 pH

1,5 pHpH 1

φ1

1,0

pHpH c C

0,5

0,0

0

1

2

3

4

5

pH

6

7

8

0.04 0.03

ALG 0.1 %-BSA 0.1 % ALG 0.1 %-BSA 0.15 % ALG 0.1 %-BSA 0.2 % ALG 0 %-BSA 0.1 % ALG 0.1 %-BSA 0 %

0.02 0.01 0.00 0

2

4

6

3

Absorbance at 400 nm

(b)

pHpH optopt

2

pH φ2

pH 2

pH 1

1

pHφ1 pHpH c C

0

0

1

2

3

4

pH

5

6

7

8

0.02 0.02

ALG 0.1 %-BLG 0.1 % ALG 0.1 %-BLG 0.15 % ALG 0.1 %-BLG 0.2 % ALG 0 %-BLG 0.1 % ALG 0.1 %-BLG 0 %

0.01 0.01 0.00 0

2

4

6

Figure 1. Phase diagram demonstrating critical pH values as a function of pH for BSA-ALG (a) and BLG-ALG (b)

25 ACS Paragon Plus Environment

Page 26 of 34

Page 27 of 34

Journal of Agricultural and Food Chemistry

25 (a)

Number (%)

20

15

10

5

0 0,1

1

10

100

1000

10000

1000

10000

LogSize size (d.nm) (d.nm) pH 4.8 pH 4.2 pH 2.8 pH 1.8

25 (b)

Number (%)

20

15

10

5

0 0,1

1

10

100

Log size(d.nm) (d.nm) Size pH 4.8 pH 4.2 pH 2.4 pH 1.6

26 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

25 (d)(c) pH 4.8 pH 4.2 pH 2.8 pH 1.8

Number (%)

20

15

10

5

0 0,1

1

10

100

1000

10000

Size (d.nm) Log size (d.nm)

25 (e)(d)

pH 4.8 pH 4.2 pH 2.4 pH 1.6

Number (%)

20

15

10

5

0 0,1

1

10

100

1000

10000

(d.nm) Log sizeSize (d.nm)

Figure 2. Particle size distributions of ALG 0.1 %-BSA 0.2 % (a), ALG 0.1 %-BLG 0.2 % (b) BSA

0.2% (c) and BLG 0.2% (d) solutions at critical pH values

27 ACS Paragon Plus Environment

Page 28 of 34

Page 29 of 34

Journal of Agricultural and Food Chemistry

a

b

d

c

Figure 3. TEM images (7800x) of ALG (0.1%)-BSA (0.2%) dispersions at four pH values: 4.8(a), 4.2(b), 2.8(c) and 1.8(d)

28 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

e

f

g

h

*Red square shows the coacervates which are spherical particles with clear center

Figure 4. light microscopic images (20X) of ALG (0.1%)-BSA (0.2%) dispersions at four pH values: 4.8(e), 4.2(f), 2.4(g) and 1.6(h)

29 ACS Paragon Plus Environment

Page 30 of 34

Page 31 of 34

Journal of Agricultural and Food Chemistry

a

b

c

d

Figure 5. TEM images (7800X) of ALG (0.1%)-BLG (0.2%) dispersions at four pH values: 4.8(a), 4.2(b), 2.4(c) and 1.6(d)

30 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

e

f

g

h

*Red square shows the coacervates which are spherical particles with clear center

Figure 6. Light microscopic images (20X) of ALG (0.1%)-BLG (0.2%) dispersions at four pH values: 4.8(e), 4.2(f), 2.4(g) and 1.6(h)

31 ACS Paragon Plus Environment

Page 32 of 34

Page 33 of 34

Journal of Agricultural and Food Chemistry

Figure 7. Thermograms (upper panel) and binding isotherms with theoretical fits (lower panel) concerning ALG dispersion (0.1%) in the titration of the BLG dispersion (0.2%), each dissolved in sodium citrate buffer solution (5 mM, pH 4.20), at 25oC)

32 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 8. Thermograms (upper panel) and binding isotherms with theoretical fits (lower panel) concerning ALG dispersion (0.1%) in the titration of the BSA dispersion (0.1%), each dissolved in sodium citrate buffer solution (5 mM, pH 4.20), at 25oC)

33 ACS Paragon Plus Environment

Page 34 of 34