Nanocomplexes arising from protein-polysaccharide electrostatic interaction as a promising carrier for nutraceutical compounds

Nanocomplexes arising from protein-polysaccharide electrostatic interaction as a promising carrier for nutraceutical compounds

Accepted Manuscript Nanocomplexes arising from protein-polysaccharide electrostatic interaction as a promising carrier for nutraceutical compounds Sey...

2MB Sizes 0 Downloads 36 Views

Accepted Manuscript Nanocomplexes arising from protein-polysaccharide electrostatic interaction as a promising carrier for nutraceutical compounds Seyed Mohammad Hashem Hosseini, Zahra Emam-Djomeh, Paolo Sabatino, Paul Van der Meeren PII:

S0268-005X(15)00156-3

DOI:

10.1016/j.foodhyd.2015.04.006

Reference:

FOOHYD 2949

To appear in:

Food Hydrocolloids

Received Date: 10 October 2014 Revised Date:

3 March 2015

Accepted Date: 5 April 2015

Please cite this article as: Hosseini, S.M.H., Emam-Djomeh, Z., Sabatino, P., Van der Meeren, P., Nanocomplexes arising from protein-polysaccharide electrostatic interaction as a promising carrier for nutraceutical compounds, Food Hydrocolloids (2015), doi: 10.1016/j.foodhyd.2015.04.006. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT Graphical abstract: Schematic representation of the nutraceuticals entrapment within electrostatically stable nanocomplexes arising from β-lactoglobulin-sodium alginate interaction Title: Nanocomplexes arising from protein-polysaccharide electrostatic interaction as a promising carrier for nutraceutical compounds

RI PT

Seyed Mohammad Hashem Hosseinia,*, Zahra Emam-Djomehb, Paolo Sabatinoc, Paul Van der Meerenc a

M AN U

SC

Department of Food Science and Technology, College of Agriculture, Shiraz University, 71441-65186 Shiraz, Iran b Department of Food Science, Technology and Engineering, Faculty of Agricultural Engineering and Technology, Agricultural Campus of the University of Tehran, 31587-11167 Karadj, Iran, P. O. Box: 4111 c Particle and Interfacial Technology Group, Faculty of Bioscience Engineering, Ghent University, Coupure Links 653, B-9000 Gent, Belgium * Corresponding author. Tel.: +98 71 32286110; fax: +98 71 32286110 E-mail address: [email protected] (S. M. H. Hosseini).

Nutraceutical

Anionic polysaccharide

+

β-lactoglobulin– nutraceutical complex

AC C

EP

β-lactoglobulin

TE D

pKa
Soluble nanocomplexes

ACCEPTED MANUSCRIPT 1

Nanocomplexes arising from protein-polysaccharide electrostatic interaction

2

as a promising carrier for nutraceutical compounds

3

Seyed Mohammad Hashem Hosseinia,b,*, Zahra Emam-Djomehb, Paolo Sabatinoc,

5

Paul Van der Meerenc

RI PT

4

6

7

a

8

University, 71441-65186 Shiraz, Iran

9

b

M AN U

SC

Department of Food Science and Technology, College of Agriculture, Shiraz

Department of Food Science, Technology and Engineering, Faculty of Agricultural

10

Engineering and Technology, Agricultural Campus of the University of Tehran,

11

31587-11167 Karadj, Iran, P. O. Box: 4111

12

c

13

Ghent University, Coupure Links 653, B-9000 Gent, Belgium

16

17

18

TE D

EP

15

AC C

14

Particle and Interfacial Technology Group, Faculty of Bioscience Engineering,

19

* Corresponding author. Tel.: +98 71 32286110; fax: +98 71 32286110

20

E-mail address: [email protected] (S. M. H. Hosseini).

21

1

ACCEPTED MANUSCRIPT ABSTRACT

23

The main purpose of the current work was exploring the potential application of the

24

protein–polysaccharide soluble nanocomplexes as delivery systems for

25

nutraceuticals in liquid foods. In this study, the intrinsic transporting property of β-

26

lactoglobulin (BLG) was utilized to develop nanoscale green delivery systems. The

27

binding analysis using fluorescence spectroscopy suggested that the complexation

28

between BLG and four nutraceutical models including β-carotene, folic acid,

29

curcumin and ergocalciferol occurred under all conditions but varied as a function of

30

pH and nutraceutical type. The 1H-NMR study of hydrophilic ligands binding to BLG

31

provided complementary information on the interactions between protein and water

32

soluble ligands. These findings resulted in designing nanoscopic delivery systems for

33

encapsulation of both hydrophobic and hydrophilic bioactives in clear liquid food

34

products of acidic pH. The stability experiments demonstrated that the nutraceuticals

35

of low solubility in water were successfully entrapped within electrostatically stable

36

nanocomplexes arising from BLG-sodium alginate interactions. The electrophoretic

37

mobility analysis showed that soluble nanocomplexes had good stability against

38

aggregation.

39

Keywords: Complex coacervation; Beta-lactoglobulin; Nanoparticle; Delivery system;

40

Fluorescence spectroscopy; 1H-NMR

AC C

EP

TE D

M AN U

SC

RI PT

22

Chemical compounds studied in this article 41

42

Beta-carotene (PubChem CID: 5280489); Curcumin (PubChem CID: 969516); Ergocalciferol (PubChem CID: 5280793); Folic acid (PubChem CID: 6037)

2

ACCEPTED MANUSCRIPT 43

44

1. Introduction Recently, some classes of the chemical compounds such as minerals (Fe+2, Mg+2), antioxidants (tocopherols, flavonoids, phenolic compounds), carotenoids (β-

46

carotene, lycopene, lutein, zeaxanthin), vitamins (D, B1, B2), fatty acids (omega 3,

47

conjugated linoleic acid), phytosterols (stigmasterol, β-sitosterol, campesterol), fibers

48

(inulin) and prebiotics have been the focus of research. After isolation from a food

49

matrix such compounds are called nutraceuticals: ‘the link between nutrients and

50

pharmaceuticals’, a term coined by Stephen DeFelice in 1979 and defined as ‘food

51

or parts of food that provide medical or health benefits, including the prevention and

52

treatment of disease’ (DeFelice, 1995). The value of these traditional supplements

53

led to their application for food enrichment and fortification and ultimately to develop

54

new functional foods (McClements, Decker, Park, & Weiss, 2009; Livney, 2010).

SC

M AN U

Most of the nutraceuticals show instability against chemical or physical

TE D

55

RI PT

45

degradation and tend to degrade during storage when incorporated into foods.

57

Encapsulation systems, also known as ‘delivery systems’, are typically used to

58

incorporate them into the foods (Shimoni, 2009). So, the development of structured

59

delivery systems for the encapsulation of bioactives is an important area of research

60

for the food industry in order to improve the quality of foods and beverages

61

(Matalanis, Jones, & McClements, 2011). Due to the hydrophobic nature of the most

62

functional compounds, their incorporation into non-fat aqueous foods and beverages

63

(especially clear ones) is challenging (Sagalowicz, & Leser, 2010; Matalanis et al.,

64

2011). In this case, the encapsulated bioactive ingredients must be stabilized in a

65

liquid environment, which is completely different from stabilizing the bioactives in a

66

solid environment (Sagalowicz et al., 2010).

AC C

EP

56

3

ACCEPTED MANUSCRIPT 67

Different types of delivery systems have been used to introduce bioactives into foods including oil in water ordinary and multilayered emulsions (Saberi, Fang, &

69

McClements, 2014; Yoksan, Jirawutthiwongchai, & Arpo, 2010), multiple emulsions

70

(Jiménez-Colmenero, 2013; Santos, Bozza, Thomazini, & Favaro-Trindade, 2015),

71

microemulsions and nanoemulsions (Qian, Decker, Xiao, & McClements, 2012;

72

Rao, & McClements, 2012, Davidov-Pardo, & McClements, 2015), solid lipid

73

nanoparticles (SLNs) (Pandita, Kumar, Poonia, & Lather, 2014), cyclodextrins (Choi,

74

Ruktanonchai, Min, Chun, & Soottitantawat, 2010), amylose (Zabar, Lesmes, Katz,

75

Shimoni, & Bianco-Peled, 2010), liposomes (Takahashi, Inafuku, Miyagi, Oku, Wada,

76

Imura, & Kitamoto, 2007), and micelles (Zimet, Rosenberg, & Livney, 2011).

SC

M AN U

77

RI PT

68

It seems that the main attention in the research field of aqueous delivery systems is to try to physically or chemically ‘complex’ or ‘bind’ the active ingredients

79

(specially hydrophobic ones) to a molecular or supramolecular structure in order to

80

protect it in this way from chemical or physical deterioration (Sagalowicz et al.,

81

2010). Many types of biopolymers are capable of binding lipophilic and hydrophilic

82

molecules and forming molecular complexes. The bioactive molecules may be

83

bound to individual biopolymer molecules (such as globular proteins, amylose and

84

starch derivatives) at one or more active binding sites by either specific or non-

85

specific interactions with different molecular origins (hydrophobic or electrostatic)

86

(Zabar et al., 2010; Liang, & Subirade, 2010), or they may be incorporated within

87

clusters formed by a single type (such as casein micelles) (Shapira, Assaraf, &

88

Livney, 2010; Zimet et al., 2011) or mixed types (such as molecular clusters arising

89

from electrostatic attraction between proteins and ionic polysaccharides) of

90

biopolymers (Zimet, & Livney, 2009; Ron, Zimet, Bargarum, & Livney, 2010).

AC C

EP

TE D

78

4

ACCEPTED MANUSCRIPT 91

The formation of non-covalent electrostatic complexes between proteins and polysaccharides can potentially lead to different functional properties, compared to

93

the two biopolymers taken individually (Schmitt, Sanchez, Desobry-Banon, & Hardy,

94

1998; McClements, 2006). This is generally due to a synergistic combination of the

95

functional features of both the protein and the polysaccharide (Schmitt, Aberkane, &

96

Sanchez, 2009). Mixing proteins and polysaccharides in an aqueous medium leads

97

to two different types of phase separation phenomena including thermodynamic

98

incompatibility (repulsive) and thermodynamic compatibility (attractive) depending on

99

the charged properties of both biopolymers, and hence on the factors influencing

SC

RI PT

92

them, such as the pH and the ionic strength. Other important parameters influencing

101

the complex coacervation between biopolymers include total biopolymer

102

concentration, protein to polysaccharide mixing ratio, molecular weight, charge

103

density, molecular conformation, charge distribution, processing factors such as

104

temperature, pressure, shearing and acidification method (Schmitt et al., 2009;

105

Turgeon, & Laneuville, 2009). During attractive interactions, protein and

106

polysaccharide in the biopolymer rich phase are held together mainly by electrostatic

107

forces and can appear as coacervates, complexes (soluble or insoluble) and gels

108

(Turgeon et al., 2009). Among the formed structures, coacervates can be used as

109

delivery systems for encapsulation purposes. However, interaction between

110

coacervates leads to coalescence and formation of transient multivesicular

111

structures that can coalesce further and eventually completely separate into a dense

112

coacervated phase (Sanchez, Mekhloufi, Schmitt, Renard, Robert, Lehr, Lamprecht,

113

& Hardy, 2002), which limits its application as a delivery system in fortification of

114

clear liquid foods where a uniform structure is a concern. During recent years, some

115

researchers have examined the potential applications of soluble and/or insoluble

AC C

EP

TE D

M AN U

100

5

ACCEPTED MANUSCRIPT 116

complexes arising from protein-polysaccharide interactions in order to encapsulate

117

hydrophilic and lipophilic molecules (Bedie, Turgeon, & Makhlouf, 2008; Zimet et al.,

118

2009; Ron et al., 2010). BLG as a small globular protein contains 162 amino acid residues with one free

RI PT

119

thiol group and two disulfide bonds and has a molecular weight of 18.4 kDa (Fox,

121

2009). It is a member of the lipocalin family of proteins because of its ability to bind

122

small hydrophobic molecules. The quaternary structure (association properties) of

123

the protein varies among monomers, dimers or oligomers depending on the pH,

124

temperature, concentration and ionic strength as a result of a delicate balance

125

among hydrophobic, electrostatic and hydrogen-bond interactions (Sakurai, & Goto,

126

2002; Gottschalk, Nilsson, Roos, & Halle, 2003). At pH 5–8, BLG exists as a dimer,

127

at pH 3–5 the dimers associate to form octamers, and at extreme pH values (<2 or

128

>8) most protein exists as monomers. At pH>9, the molecule is irreversibly

129

denatured (Harnsilawat, Pongsawatmanit, & McClements, 2006).

M AN U

TE D

130

SC

120

BLG folds up into an 8-stranded (A-H) antiparallel β-barrel (Fig. 1). A ninth βstrand (I) flanks the first strand. It is this strand that forms a significant part of the

132

dimer interface in the bovine protein. The conical central cavity (the calyx or β-barrel)

133

which provides the main ligand-binding site is made of β-strands A-D forming one

134

sheet, and β-strands E-H forming a second sheet. It seems that the β-barrel can

135

accommodate linear molecules like palmitic acid and also retinol with the β-ionone

136

ring system inside (Kontopidis, Holt, & Sawyer, 2004; Edwards, Creamer, &

137

Jameson, 2009). There is a 3-turn α-helix on the outer surface of the β-barrel

138

between strands G and H. The loops that connect the β-strands at the closed end of

139

the β-barrel (BC, DE, and FG) are generally quite short, whereas those at the open

140

end are significantly longer and more flexible (Kontopidis et al., 2004; Edwards et al.,

AC C

EP

131

6

ACCEPTED MANUSCRIPT 2009). In particular, the EF loop (residues 85-90) acts as a gate over the binding site.

142

The latch for this gate is the side chain of the Glu89 (Qin, Bewley, Creamer, Baker,

143

Baker, & Jameson, 1998). Indeed, the pH-dependent conformational change of this

144

loop is regulated by the protonation of Glu89 (Ragona, Fogolari, Catalano, Ugolini,

145

Zetta, & Molinari, 2003). At low pH, the gate is in closed position, and binding is

146

inhibited or impossible, whereas at high pH it is open, allowing ligands to penetrate

147

into the calyx (Kontopidis et al., 2004). Computational studies have shown that three

148

potential binding sites are possible for ligand binding to BLG: the calyx, the surface

149

pocket with a hydrophobic lining in a groove between the α-helix and the β-barrel

150

and the outer surface near Trp19-Arg124 (Liang, Tajmir-Riahi, & Subirade, 2008).

M AN U

SC

RI PT

141

Among the large spectrum of protein-polysaccharide pairs which have been

152

studied under the effects of different parameters, biopolymer complexes formation

153

between sodium alginate (Na-ALG) and milk proteins was extensively investigated.

154

Harnsilawat et al. (2006) investigated the characteristics of Na-ALG, β-lactoglobulin

155

(BLG) and their mixtures in aqueous solutions under the effect of pH (3-7) using

156

isothermal titration calorimetry (ITC), dynamic light scattering (DLS), turbidity, zeta-

157

potential and soluble protein measurements. At pH 5, BLG and Na-ALG formed fairly

158

soluble complexes due to the electrostatic attraction between negatively charged

159

carboxyl groups (-COO-) of Na-ALG and positively charged surface patches (resulted

160

from basic amino acids) of BLG. Zhao, Li, Carvajal, and Harris (2009) used both the

161

native bovine serum albumin (BSA) and the heat-denatured BSA to study the effect

162

of protein conformational changes during protein-alginate complex formation. In this

163

work a comparison was performed between the mixtures of Na-ALG with either

164

native BSA or the heat-denatured BSA using zeta-potential analyzer, DLS and

165

turbidimetric analysis in combination with protein conformational studying tools,

AC C

EP

TE D

151

7

ACCEPTED MANUSCRIPT Fourier transform infrared spectroscopy (FT-IR) and size exclusion chromatography.

167

Peinado, Lesmes, Andrés, and McClements (2010) fabricated sub-micrometer

168

biopolymer particles by electrostatic complexation of heat-denatured lactoferrin

169

particles with Na-ALG. Biopolymer particles were formed by mixing the protein

170

particles with Na-ALG at pH 8 and then lowering the pH to promote electrostatic

171

deposition of polysaccharides onto the protein particle surfaces. The effects of pH (2-

172

11) and ionic strength (0-200 mM NaCl) on the properties and the stability of the

173

complexes were studied using turbidity, DLS and electrophoresis measurements.

174

Hosseini et al. (2013a, b), studied the interaction of BLG with either Na-ALG or κ-

175

carrageenan (before and after an ultrasound treatment) using ITC, streaming current

176

detector (SCD), turbidity, dynamic light scattering and electrophoretic mobility

177

measurements. The results showed that the sonication decreased the interaction

178

strength between BLG and treated anionic polysaccharide. Fioramonti, Perez,

179

Aríngoli, Rubiolo, and Santiago (2014) designed and characterized soluble

180

biopolymer complexes formed by self-assembly of Na-ALG and whey protein isolate

181

(WPI). Resultant self-assembled biopolymer complexes were then characterized by

182

the same set of spectroscopic techniques. Higher WPI:Na-ALG ratios produced

183

soluble self-assembled biopolymer particles, where hydrophobic patches of protein

184

aggregates would be more occluded ensuring the protection of potential lipophilic

185

bioactive agents attached inside. It has also been reported that BLG can form water-

186

soluble complexes with docosahexaenoic acid (DHA) and ergocalciferol and protect

187

these lipophilic compounds from degradation by heat and oxidation (Zimet et al.,

188

2009; Ron et al., 2010). Liang et al. (2008) and Forrest, Yada, and Rousseau (2005)

189

concluded that interaction with BLG may strongly influence the stability of phenolic

190

compounds and vitamin D3, respectively and hence their bioavailability in processed

AC C

EP

TE D

M AN U

SC

RI PT

166

8

ACCEPTED MANUSCRIPT 191

foods. It has been shown that BLG has some antioxidant activity, apparently due to

192

its free thiol group (Liu, Chen, & Mao, 2007). Therefore, BLG can be used as a

193

versatile carrier of bioactive molecules in controlled delivery applications. The hypothesis of the current study is that the binding properties of BLG (a

RI PT

194

member of the lipocalin protein family) towards some bioactive compounds can be

196

used to produce green delivery systems. To form a core-shell structure, the

197

produced nanoscale delivery system can be overprotected by deposition of an

198

anionic polysaccharide as a secondary layer (or shell) around the protein core in the

199

pH range between the pKa of the polysaccharide and the isoelectric point of BLG.

200

The first objective of the current study was to determine the binding properties of four

201

nutraceutical models including curcumin (CUR), folic acid (FA), β-carotene (βC) and

202

vitamin D2 (VD) to β-lactoglobulin (BLG) under the effect of pH using

203

spectrofluorometry technique. The second objective was to study the binding

204

between BLG and hydrophilic ligands using 1H-NMR. The third objective was to

205

assess the stabilization efficiencies of nutraceutical compounds in an acidic (pH

206

4.25) clear beverage model using nanoparticles (soluble complexes) produced via

207

electrostatic interactions between BLG and Na-ALG.

208

2. Materials and methods

209

2.1. Materials

M AN U

TE D

EP

AC C

210

SC

195

Guluronate-rich sodium alginate (ALG, composition: 66.26% (w/w) ALG,

211

14.19% (w/w) moisture and 9.55% (w/w) ash) from Laminaria hyperborean was

212

purchased from BDH Co. (Poole, UK). The molecular weight of sodium alginate was

213

200 kDa and the uronate compositions were 37.5% mannuronate and 62.5%

214

guluronate. β-lactoglobulin isolate (BLG, 18.4 kDa, minimum purity 90% (w/w)) from 9

ACCEPTED MANUSCRIPT bovine milk was obtained from Davisco Foods International Inc. (Eden Prairie, MN,

216

USA). Sodium azide (as a preservative), ergocalciferol (vitamin D2) (VD), β-carotene

217

(βC), curcumin (CUR), folic acid (FA) and N-acetyl-L-tryptophanamide (NATA) were

218

obtained from Sigma Chemical Co. (St. Louis, MO, USA). Deuterium oxide (D2O,

219

99.8% D) was obtained from Armar Chemicals (Döttingen, Switzerland). Glacial

220

acetic acid, analytical grade hydrochloric acid, sodium hydroxide, monosodium

221

dihydrogen phosphate, 8-anilinonaphthalene-1-sulfonic acid ammonium salt (ANS)

222

and absolute ethanol were purchased from Acros Organics (Geel, Belgium).

223

Deionized water (18.2 MΩ cm resistivity) from a MilliQ water system

224

(Millipore Corporation, MA, USA) was used for the preparation of all solutions. In this

225

study, all materials were used directly from the sample containers without additional

226

purification taking into account their purity.

227

2.2. Preparation of solutions

SC

M AN U

ALG (0.8% (w/w)) and BLG (0.4% (w/w)) stock solutions were prepared by

TE D

228

RI PT

215

dispersing in deionized water containing 0.03% (w/w) sodium azide. The solutions

230

were then stirred at 250 rpm at ambient temperature for 12 h to ensure complete

231

hydration of the biopolymers in order to use on the following day.

232

2.3. Bioactives binding to BLG

AC C

233

EP

229

Fluorescence spectroscopy was used to study the binding of bioactives to BLG

234

by measuring the binding-induced quenching of the intrinsic fluorescence of BLG

235

tryptophanyl residue (Trp19), which is particularly sensitive to changes of its

236

microenvironment (Wang, Allen, & Swaisgood, 1998). For this experiment, BLG

237

stock solutions were made fresh daily by dissolving in either 10 mM phosphate buffer

238

at pH 7 or 10 mM acetate buffer at pH 4.25 (to determine the effect of pH on the

239

binding properties of ligands). After filtration through a 0.22-µm syringe filter to obtain 10

ACCEPTED MANUSCRIPT aggregate free BLG dispersion, the protein concentrations in the solutions were

241

determined by UV spectroscopy using an Ultrospec 1000 UV-visible

242

spectrophotometer (Pharmacia-Biotech, Biochrom Ltd., Cambridge, England). The

243

samples were measured in a 1 cm quartz cell at 278 nm against buffer solutions as

244

the reference. BLG concentrations were calculated using an extinction coefficient (ε)

245

of 17600 M-1 cm-1 (Liang et al., 2008) and amounted to 2.90 and 3.52 µM at pH 7

246

and 4.25, respectively. The low concentration of the BLG solutions avoided inner

247

filter effects which could occur during the experiment. To facilitate the binding of

248

nutraceutical components to BLG, they were prepared daily by dissolving in absolute

249

ethanol, except with FA which was dissolved in phosphate buffer (pH 7). To prevent

250

degradation, nutraceutical stock solutions of the appropriate concentrations (βC:

251

0.130 and 0.104 mM, FA: 0.213 and 0.199 mM, CUR: 0.201 and 0.255 mM, VD:

252

0.393 and 0.509 mM, at pH 7 and 4.25, respectively) were purged with nitrogen gas,

253

and stored in the dark at 4 ºC. Samples were prepared at room temperature in 2.5 ml

254

plastic tubes covered with aluminum foil, by mixing BLG (2 ml) and different amounts

255

of nutraceutical (0, 2, 5, 9, 14, 20, 27, 35, 44, 54 µl) stock solutions. The protein-

256

nutraceutical solutions were vortexed for 30 s and allowed to equilibrate for 10 min

257

prior to fluorescence measurement. As the ethanol dissipates in the water, most of

258

the nutraceutical components bind to the protein’s binding site(s). The highest

259

resulting ethanol concentration was about 2.7% (v/v), which had no appreciable

260

effect on protein structure. Fluorescence spectra were recorded at room temperature

261

on a Cary Eclipse Fluorescence spectrophotometer (Varian Inc., Walnut Creek, CA,

262

USA) using an excitation wavelength of 287 nm and an emission wavelength of 332

263

nm (Cogan, Kopelman, Mokady, & Shinitzky, 1976). The band slit (spectral

264

resolution) was 5 nm for both excitation and emission. To eliminate the effects of

AC C

EP

TE D

M AN U

SC

RI PT

240

11

ACCEPTED MANUSCRIPT protein dilution by the added ligand solution and the possible tryptophan

266

fluorescence changes induced by ethanol, for each sample, a blank BLG solution

267

containing an identical concentration of ethanol (and/or buffer for FA) was prepared

268

and treated in the same manner as the sample. A second blank containing NATA

269

was also prepared in a manner similar to all samples. NATA is unable to interact with

270

nutraceutical models; however it displays a fluorescence spectrum typical of

271

tryptophan. The decrease in fluorescence intensity of blanks containing NATA is not

272

due to the interaction but it results from the inner filter effect as a consequence of

273

ligand absorbance at 287 nm. NATA was also used to exclude the possibility of

274

unspecific interactions of the nutraceutical model with the protein’s tryptophan

275

indoles, where at increasing nutraceutical concentrations the nutraceutical may

276

absorb light, which would otherwise excite the indole groups, and thus fluorescence

277

would decrease for this reason (Dufour, & Haertlé, 1990) not for binding. The

278

concentration of NATA (0.228 mg/150 ml buffer and 0.265 mg/150 ml buffer, at pH 7

279

and 4.25, respectively) had been selected in the way that it had the same emission

280

at 332 nm as the studied BLG solution. The fluorescence intensity changes of the

281

blanks were subtracted from fluorescence intensity measurements of the

282

ligand/protein complexes for every considered sample. In these experiments, before

283

correction for the blanks, the fluorescence intensity at 332 nm of the original BLG

284

solution was normalized to 1. All experiments were run in duplicate samples put in

285

quartz cuvettes of 1 cm optical path length. After correcting for the blanks, the

286

differences in fluorescence intensity at 332 nm between complex and free protein

287

were used to measure the apparent dissociation constant (K'd) and the apparent

288

mole ratio of ligand to protein at saturation (n). The direct linear plotting method of

289

Eisenthal and Cornish-Bowden (1974), where the corrected fluorescence is plotted

AC C

EP

TE D

M AN U

SC

RI PT

265

12

ACCEPTED MANUSCRIPT directly against ligand concentrations, was used to obtain K'd directly from the

291

median of intersecting regression lines representing individual observations on the

292

abscissa. The n values were obtained directly from the fluorescence titration curve

293

plotted against nutraceutical/protein molar ratio correlating to the saturation point.

294

2.4. NMR study of hydrophilic ligands binding to BLG

RI PT

290

Interactions between BLG and FA and/or ANS were investigated by proton

296

nuclear magnetic resonance (1H-NMR) spectroscopy. Stock solution of ligand of

297

appropriate concentration was prepared by dissolving in deuterium oxide (D2O). To

298

determine the effect of protein on ligand binding, BLG was added to ligand stock

299

solution at different concentrations. All NMR experiments were performed on a

300

BrukerAvance II spectrometer operating at a 1H- frequency of 700.13 MHz. A 5 mm

301

TXI gradient probe with a maximum gradient strength of 57.5 G·cm-1 was used

302

throughout. Temperature was controlled to within ±0.1 °C with a Eurotherm 2000VT

303

controller. Diffusion coefficients were measured by PFG-NMR with a convection

304

compensated double-stimulated-echo experiment (Jerschow, & Muller, 1997) using

305

monopolar smoothened square shaped gradient pulses and a modified phase cycle

306

(Connell, Bowyer, Bone, Davis, Swanson, Nilsson, & Morris, 2009).

EP

TE D

M AN U

SC

295

The echo-decay of the resonance intensity obtained with the double stimulated

308

echo sequence obeys equation (1), from which it is clear that the diffusion coefficient

309

(D) is derived from the echo-decay as a function of the parameter k

AC C

307

[

310

I = I 0 exp − D ( γ G δ s ) 2 ∆ '

311

I = I 0 exp[− D ⋅ k ]

]

(1)

13

ACCEPTED MANUSCRIPT where I is the echo intensity with gradient; I0 is the echo intensity at zero gradient; γ

313

is gyromagnetic ratio; G the maximum gradient amplitude; δ the duration of the

314

gradient pulse and ∆' is the diffusion delay corrected for the finite gradient pulse

315

duration ∆' = ∆ − 0.6021 ⋅ δ . The gradient shape factor s was set to 0.9, to account for

316

the smoothed rectangular gradient shape used in the experiments.

317

2.5. Nanoencapsulation of nutraceutical models

After finding nutraceutical binding characteristics to BLG, a series of solutions

SC

318

RI PT

312

containing constant final BLG concentration (0.1% w/w), and constant final

320

nutraceutical concentration at a molar ratio of 1:1 were prepared by adding

321

nutraceutical compound dissolved in absolute ethanol (except for FA which was

322

dissolved in deionized water) to the protein solutions. The protein solution (at neutral

323

pH) was then stirred for half an hour. Polysaccharide stock solution was added to the

324

BLG-nutraceutical solution at the desired quantity (0.75% w/w to obtain a transparent

325

system containing nanoparticles) (Hosseini et al., 2013) and deionized water was

326

added to obtain a final constant volume as well as constant protein and

327

polysaccharide concentrations. Then the pH was adjusted while stirring to 4.25 using

328

0.4, 0.1 and/or 0.01 M HCl solutions (the post-blending acidification method), and the

329

samples were stirred further for half an hour for equilibration. For each sample, a

330

blank consisting of deionized water which contained an identical concentration of

331

nutraceutical compound dissolved in ethanol (and/or deionized water for folic acid)

332

was prepared and treated in the same manner as the sample. A second blank

333

containing BLG and nutraceutical compound without sodium alginate was also

334

prepared in a manner similar to all samples.

AC C

EP

TE D

M AN U

319

14

ACCEPTED MANUSCRIPT 335 336

2.6. Particle size and electrophoretic mobility analyses Measurements of particle size distribution were carried out using a dynamic light scattering (DLS) instrument (90Plus, Brookhaven Instruments Corp., Vienna,

338

Austria). Analyses were carried out at a scattering angle of 90° at 25 °C. The effective

339

diameter (also called Z-average mean diameter) was obtained by cumulant analysis.

340

The electrophoretic mobility was determined by laser Doppler anemometry with

341

palladium electrodes using a ZetaPals instrument (Brookhaven Instruments Corp.,

342

Vienna, Austria) at fixed light scattering angle of 90° at 25 °C. During both dynamic

343

light scattering and electrophoretic light scattering measurements, the viscosity of

344

the continuous phase was assumed to correspond to pure water.

345

2.7. Statistical analysis

SC

M AN U

346

RI PT

337

Measurements were performed two or three times using freshly prepared samples and analyzed by ANOVA using the MSTATC program (version 2.10, East

348

Lansing, MI, USA). Results were reported as means and standard deviations.

349

Comparison of means was carried out using Duncan’s multiple range tests at a

350

confidence level of 0.05.

351

3. Results and discussion

352

3.1. Binding properties of the nutraceutical compounds to BLG

AC C

EP

TE D

347

353

The raw data was analyzed to measure the apparent dissociation constant (K'd)

354

and the apparent mole ratio of ligand to protein at saturation (n) which are presented

355

in Tables 1 and 2, respectively. A low K'd value indicates high binding affinity. The

356

analysis suggested that binding occurred under all conditions but varied as a

357

function of pH and nutraceutical model. From these observations (n>1 as well as

358

binding dependence on pH and nutraceutical type), it can be concluded that the four

15

ACCEPTED MANUSCRIPT nutraceutical models investigated are bound on proteins on different binding sites

360

and/or by different binding mechanisms. An opposite trend was observed for the two

361

estimated parameters at both pH values, where the weaker observed affinity was

362

associated with the higher n value.

363

RI PT

359

Among the four nutraceutical models investigated, β-carotene had the highest binding affinity and the lowest binding stoichiometry at both pH values. The more

365

hydrophobic character of β-carotene may explain the higher affinity of this molecule

366

for the hydrophobic binding sites of BLG, whereas its large size may be responsible

367

for the lower binding stoichiometry. Frapin, Dufour and Haertlé (1993) reported that

368

the affinity of BLG for saturated fatty acids increases gradually from lauric acid to

369

palmitic acid (as a longer aliphatic chain) due to an increase in the hydrophobic

370

nature of the ligand. The β-carotene-BLG molar binding ratio at pH 7 was 0.48± 0.04

371

which is in good agreement with the results (0.49 ± 0.03) reported by Dufour and

372

Haertlé (1991). The solvent exposure of a part of the retinol isoprenoid chain, when

373

bound to BLG could explain the 1:2 stoichiometry of β-carotene-BLG complexes.

374

This tetraterpen has two β-ionone rings, joined by an isoprenoid chain, which may

375

interact distally with two BLG molecules (Dufour et al., 1991). Another possibility is

376

that this ligand can bind at the dimer interface. According to Wang, Allen and

377

Swaisgood (1998, 1999) the site found at the dimer interface is the highest affinity

378

site as observed in the current study.

AC C

EP

TE D

M AN U

SC

364

379

Folic acid is less hydrophobic than curcumin and ergocalciferol but is bound to

380

a higher extent. This suggests possible interactions that are not only hydrophobic in

381

nature, and/or binding to the other binding sites than the calyx. Liang et al. (2010)

382

proposed that the binding site of resveratrol may be on the outer surface near Trp19-

383

Arg124, while folic acid binds to the surface hydrophobic pocket in a groove between 16

ACCEPTED MANUSCRIPT the α-helix and the β-barrel. The n value of folic acid-BLG complex at pH 7 was

385

1.25± 0.05. This value is in very good agreement with those (1.30 ± 0.03 and 1.17 ±

386

0.04) obtained by Liang et al. (2010) during excitation at 280 and 295 nm,

387

respectively.

388

RI PT

384

The curcumin-BLG ratio at pH 7 was 0.95± 0.06 which is in good agreement with the results (1 and 0.85) reported by Sneharani, Karakkat, Singh and Rao (2010)

390

and by Mohammadi, Bordbar, Divsalar, Mohammadi and Saboury (2009) at pH 7

391

and 6.4, respectively. Based on the obtained results (no significant change in binding

392

stoichiometry) at the lower pH level (4.25) and on those reported in literature

393

(Riihimäki, Vainio, Heikura, Valkonen, Virtanen, & Vuorela, 2008), it seems that

394

curcumin (as a phenolic compound) binds to a site other than the calyx.

M AN U

395

SC

389

The ergocalciferol-BLG molar ratio at pH 7 was 1.31± 0.12. Wang, Allen, and Swaisgood (1997) reported the value of 2.00 ± 0.16 at pH 7. It seems that the

397

increased stoichiometry (binding to the external hydrophobic surface patch) is

398

accompanied by relatively loose binding as evidenced by the decreased observed

399

affinity. According to Forrest et al., (2005), the weakest affinity site is found at the

400

hydrophobic surface patch. A higher than 1 stoichiometry indicated that the other

401

binding sites (crevice next to the alpha helix and the dimer interface) of higher or

402

equal affinity are involved in the binding to BLG. These binding sites may be

403

saturated prior to, or simultaneously with the main binding site. Zimet et al. (2009)

404

reported that 2.67 ± 1.26 moles of DHA were bound per mole of BLG.

405

AC C

EP

TE D

396

The binding capacities of BLG mediated by pH were evident in the present

406

study. Upon lowering the pH, the binding constants changed depending on the

407

nutraceutical nature. Since the EF loop, that acts as a gate in the central cavity, is in

17

ACCEPTED MANUSCRIPT a closed conformation at acidic pH (Kontopidis et al., 2004), the binding

409

stoichiometry of β-carotene was decreased with decreasing pH. Previous NMR

410

studies have shown that palmitic acid starts to be released at a pH lower than 6, and

411

80% of the palmitic acid has already been released at pH 2 (Ragona, Fogolari, Zetta,

412

Perez, Puyol, De Kruif, Lohr, Ruterjans, & Molinari, 2000). The binding affinity of β-

413

carotene at pH 4.25 was increased because hydrophobic interactions are enhanced

414

at pH values around the isoelectric point of BLG (~4.7). Since the EF loop is blocking

415

ligand access, another possibility is that a single β-carotene molecule was tightly

416

bound between the monomers (at the interfaces) found within the octamers as

417

evidenced by the strongest observed affinity.

SC

M AN U

418

RI PT

408

The decrease in the binding stoichiometry of folic acid can be attributed to its lower solubility at pH 4.25 than at pH 7 (~1 and ~5 mg/L, respectively) (Younis,

420

Stamatakis, Callery, & Meyer-Stouta, 2009).

421

TE D

419

The binding stoichiometry of curcumin was not significantly (p<0.05) different at pH 4.25 as compared to pH 7.00. Riihimäki et al. (2008) reported that, contrary to

423

retinol, the release of phenolic compounds was not observed at acidic pH, which

424

suggests that phenolic compounds and their derivatives do not bind to the central

425

calyx. Liang et al. (2008) studied the binding of the natural polyphenolic compound

426

resveratrol to bovine BLG. The observed blue shift of the fluorescence emission

427

maxima and the increase in the emission intensity implied that the environment of

428

the polyphenol bound to BLG is not as hydrophobic as the cavity of BLG, suggesting

429

binding on the surface of the protein. Moreover, the qualitative docking results

430

performed by Riihimäki et al. (2008), showed that phenolic compounds and their

431

derivatives would not bind to the central calyx, supporting the results of this study for

432

the curcumin-BLG complex.

AC C

EP

422

18

ACCEPTED MANUSCRIPT 433

The binding stoichiometry of ergocalciferol was significantly decreased by decreasing pH. At pH 4.25, approximately 0.85 molecule of vitamin D2 were bound

435

per BLG monomer with a relatively weak affinity. Since the EF loop is in a closed

436

position at this pH, the results indicate that loose binding occurred at the external

437

hydrophobic surface patch. Since protein association is increased after lowering the

438

pH to 4.25 (formation of the octamers), it seems that the decreased available surface

439

area inhibited greater ligand access at external hydrophobic surface patches and

440

interfaces. The binding to the external hydrophobic surface patches and the

441

interfaces is accompanied by relatively loose and tight affinity, respectively. The

442

absence of a significant change in binding affinity of ergocalciferol to BLG with

443

decreasing pH is likely due to a combination of both a decrease in binding affinity

444

arising from the contribution of the hydrophobic surface patches in binding and an

445

increase in binding affinity caused by the binding to interfaces resulting in an overall

446

unchanged binding affinity.

447

3.2. NMR results

SC

M AN U

TE D

NMR spectroscopy is a useful technique to study molecular interactions. NMR

EP

448

RI PT

434

provides different measurable parameters that depend on the amount and the

450

strength of the interaction. The former may be quantified by analysis of the diffusion

451

behavior of ligand and protein both separately and upon mixing. The latter gives rise

452

to important changes in the chemical environment and in the rotational as well as

453

translational mobility of the ligand, which is reflected in a change in the chemical

454

shift, the line width and relaxation rate, as well as the molecular diffusion rate

455

constants of the ligand. Indeed, binding of ligands to hydrophobic sites may cause a

456

change in chemical shift to lower values. In addition, bound ligand molecules will

457

have a reduced mobility which provokes a peak broadening. Considering the

AC C

449

19

ACCEPTED MANUSCRIPT relatively limited binding of hydrophobic ligands (ranging from 1/2 to 2/1 molar ratio),

459

the large difference in molar mass and the absence of highly intense peaks in the

460

NMR spectrum of the ligands (as compared to the intense CH2-signal in typical

461

surfactant-like chemicals), NMR detection of the ligand in the presence of a large

462

amount of protein is hampered. To overcome this problem, initial measurements

463

were performed using Anilino Naphthalene Sulphonate (ANS, data not shown). This

464

fluorescent dye is widely used as a surface hydrophobicity probe, based on the fact

465

that its fluorescence intensity is largely increased when present in a rather

466

hydrophobic environment. ANS is known to bind to proteins. In addition, its aromatic

467

nature is responsible for the fact that most ligands NMR peaks are resolved from the

468

protein peaks. 1D proton NMR is a quick method to achieve qualitative information

469

about the degree of interaction. In this type of experiment the variation in chemical

470

shift of the ligand resonances upon sorption as compared to the dissolved molecules

471

can be used to roughly estimate the partitioning coefficient as well as the type of

472

interactions. A more accurate quantification of the interaction of the ligands with BLG

473

is obtained by using Diffusion Ordered SpectroscopY (DOSY-NMR) measurements.

474

Hereby, the observed diffusion coefficient of the ligand in the presence of proteins

475

(Dobs) is the weighted average of non-bound molecules (with diffusion coefficient

476

Dfree) and protein bound molecules (with the same diffusion coefficient as the protein

477

Dpro). If the bound fraction is represented by P, the weighted average may be

478

calculated according to the following equation:

479

480

AC C

EP

TE D

M AN U

SC

RI PT

458

Dobs = PD sorb + (1 − P ) D free

(2)

from which

20

ACCEPTED MANUSCRIPT 481

P=

D free − Dobs

(3)

D free − Dsorb

Hence, the bound protein fraction follows from experimental values of the diffusion

483

coefficient of the ligands in the absence (Dfree) and presence (Dobs) of protein, as well

484

as from the diffusion coefficient of the protein (Dpro).

RI PT

482

Fig. 2 shows the chemical structure of folic acid. In this case a complete

486

assignment is possible as summarized in Fig. 3. From Fig. 4, it is clear that the

487

addition of protein only slightly influences the NMR contribution of the folic acid. The

488

resonances show little chemical shift variation and there are no signs of line

489

broadening. However, the diffusion measurements show that the diffusion behavior

490

of folic acid changes (slower diffusion) upon addition of protein, as visible in Fig. 5.

491

The sorbed amount of folic acid can be estimated according to the previously

492

mentioned equation, where Dfree is (3.19 ± 0.04)10-10 m2s-1, Dobs is (2.96 ± 0.05)10-10

493

m2s-1 and Dpro is (0.75 ± 0.01)10-10 m2s-1. The solubilized fraction is 9.4%, which

494

corresponds on average to 1.1 folic acid molecules per BLG monomer. This result is

495

in good agreement with the results of fluorimetry. The same series of experiments

496

were repeated for the same system but at different pH conditions (pH ≈ 4.25).

497

Unfortunately, in these conditions, the NMR spectrum resulted in a complete loss of

498

the folic acid resonances. As a further consequence, the diffusion experiments were

499

not possible. The very poor aqueous solubility at the acidic pH is the main reason of

500

the loss of folic acid resonance at lower pH conditions.

501

AC C

EP

TE D

M AN U

SC

485

A more accurate analysis of the chemical shift variations is also required to

502

reveal which ligand functional group is more involved in contact with protein and as

503

consequence to establish the nature of the interactions. However, this technique can

504

only be used when the ligand has a sufficiently large solubility (i.e. at least some 21

ACCEPTED MANUSCRIPT 505

mM) in (heavy) water. Hence, NMR could not be used to study the binding of β-

506

carotene, curcumin or ergocalciferol.

507

509

3.3. Nanoencapsulation and colloidal stability of nutraceutical models

RI PT

508

The effects of BLG and Na-ALG on the colloidal stability of nanoencapsulated curcumin (as hydrophobic model) and folic acid (as hydrophilic model with low

511

solubility at acidic pH) at pH 4.25 are shown in Figs. 6 and 7, respectively. Except for

512

the blank sample of curcumin (which was first dissolved in ethanol before being

513

added to the aqueous phase) in deionized water, there was not any significant

514

difference between the samples just after production (Figs. 6.A and 7.A). After 2

515

hours of production, some differences were observed. The curcumin and folic acid,

516

incorporated into deionized water, started to precipitate and separated almost

517

completely after 24 hours of production. The samples containing BLG and BLG+ALG

518

did not show any precipitation after 24 hours (Figs. 6.B and 7.B). However, the

519

turbidity of both samples was slightly different. 48 hours after production, the

520

differences between these two samples were more obvious (Figs. 6.C and 7.C): the

521

blank sample containing the nutraceutical models and BLG showed some

522

precipitation, while the main sample, which contained the nutraceutical models, BLG

523

and ALG, remained completely transparent without any sign of precipitation. This

524

clearly demonstrated the efficacy of soluble nanocomplexes arising from protein-

525

polysaccharide interactions on nanoencapsulation and colloidal stability of

526

nutraceuticals of low solubility in water even 30 days after production without storing

527

in the dark (Fig. 6.D and 7.D). The low solubility of curcumin in aqueous solution

528

significantly limits its application. Recently, it has been shown that polyelectrolyte-

529

coated curcumin nanoparticles (obtained through a layer by layer shell assembly)

AC C

EP

TE D

M AN U

SC

510

22

ACCEPTED MANUSCRIPT are hydrosoluble (Zheng, Zhang, Carbo, Clark, Nathan, & Lvov, 2010). In our study,

531

curcumin binding to BLG increased its colloidal stability in water. The stability of the

532

main sample containing folic acid was lower than those containing curcumin. The

533

nutraceutical models were incorporated in a 1:1 molar ratio into the protein solution.

534

The apparent mole ratio of folic acid to BLG at pH 4.25 was determined to be 0.39 ±

535

0.04 (Table 2). Hence, it is possible that after decreasing the pH to 4.25, the BLG

536

became overloaded and some precipitation occurred. As the apparent mole ratio of

537

curcumin to BLG at pH 4.25 was found to be 0.82 ± 0.08 (Table 2), BLG overloading

538

was relatively prevented in this case. The particle size and electrophoretic mobility

539

(EM) analyses of the samples containing BLG (0.1% w/w) and ALG (0.075 % w/w)

540

with and without nutraceutical model compounds showed that curcumin addition into

541

the soluble complexes of BLG-ALG, slightly increased the particle size from 269 nm

542

to 278 nm and also slightly decreased the EM from -4.52 to -4.30 (10-8 m2/Vs). Folic

543

acid incorporation into soluble complexes resulted from BLG and Na-ALG

544

interactions, increased the particle size from 269 nm to values higher than 1µm

545

(maybe due to the low solubility of folic acid at pH 4.25) and also decreased the EM

546

from -4.52 to -3.95 (10-8 m2/Vs). These experimental values are in good agreement

547

with the qualitative visual observations, which indicated an increased turbidity in the

548

presence of folic acid. A similar decrease in the absolute value of the EM was

549

reported by Zimet et al. (2009) upon DHA addition into BLG-pectin soluble

550

complexes. One should keep in mind that the measured results are intensity-

551

weighted, which means that the larger particles have the larger contribution. If

552

volume- or number- weighted distributions are considered, much smaller average

553

diameters are obtained. For any delivery system, it is essential that the system

554

remains stable throughout the entire life cycle of the product. Furthermore, the

AC C

EP

TE D

M AN U

SC

RI PT

530

23

ACCEPTED MANUSCRIPT biopolymer nanoparticles should not adversely impact the normal shelf-life of the

556

product itself (Matalanis et al., 2011). EM can be obtained from the perturbations of

557

Brownian diffusivity under a pulsating electrical field and is a crucial parameter for

558

predicting the stability of colloidal delivery systems (Cooper, Dubin, Kayitmazer, &

559

Turksen, 2005; Matalanis et al., 2011).

560

RI PT

555

Nanoparticle formation from the protein – polysaccharide electrostatic

interaction could be understood as the shrinkage of the protein-polysaccharide

562

complexes which occurs at low ionic strength. This shrinkage is a result of a

563

decrease in the intramolecular repulsion of like-charged groups of polysaccharide

564

induced by the interaction of the BLG molecules with the carboxyl groups of the Na-

565

ALG led to a decrease in the backbone chain rigidity (Fig. 8) (Tolstoguzov, 2003).

566

This compaction phenomenon was well predicted by Monte Carlo simulations which

567

showed that at low ionic strengths, a polyelectrolyte chain would wrap around an

568

oppositely charged spherical particle (Girard, Turgeon, & Gauthier, 2003). Indeed,

569

mutual neutralization decreases the net charge, hydrophilicity and chain rigidity of

570

the junction zones resulting in a compact conformation of the complex with the

571

hidden junction zones (Tolstoguzov, 2003).

M AN U

TE D

EP

Another important attribute of a suitable delivery system is its stability against

AC C

572

SC

561

573

processing conditions (such as thermal processing and pressure) which can result in

574

protein denaturation and hence affecting its transporting properties. The precise

575

denaturation process is complex and is influenced by factors such as pH, protein

576

concentration, ionic environment, genetic variant and presence of ligands. Both

577

lowering the pH (Relkin, Eynard, & Launay, 1992) and adding calyx-bound ligands

578

(Busti, Gatti, & Delorenzi, 2006) make the protein more resistant to thermal

579

unfolding. Enzymatic proteolysis observations indicate that BLG is less susceptible 24

ACCEPTED MANUSCRIPT 580

to pressure-induced changes at acidic pH than at neutral or basic pH (Edwards et

581

al., 2009).

582

The targeted delivery and controlled release are also very important. Active ingredients release from whey protein gels was investigated by Tomczy ska-Mleko,

584

& Mleko (2014). As an example, it is beneficial for the encapsulated folic acid to be

585

released in the small intestine where most of the absorption of vitamins takes place.

586

The jejunum is the site of maximum absorption of free folates, where absorption

587

occurs by a pH dependent, carrier-mediated system (Kailasapathy, 2008). Therefore,

588

the biopolymers used for encapsulation should be able to protect the folate in the

589

upper gastrointestinal tract (acidic stomach conditions) and release the folate in the

590

alkaline conditions of the small intestine. The BLG structure is relatively compact and

591

stable in aqueous solutions at acidic pH, as demonstrated by its resistivity to

592

proteolysis by pepsin (Mohan Reddy, Kella, & Kinsella, 1988). The compact structure

593

of BLG at acidic pH can be overprotected by using an anionic polysaccharide due to

594

the electrostatic interactions. The BLG-anionic polysaccharide complexes will be

595

dissociated at alkaline pH of the small intestine (due to the charge similarity of both

596

biopolymers). Hydrophobic interactions are mainly responsible for the binding of

597

nutraceuticals to BLG. Since those interactions are entropy driven and depend on

598

the presence of the native structure in BLG, the release process will occur when the

599

protein loses its native structure during digestion in small intestine.

600

4. Conclusion

AC C

EP

TE D

M AN U

SC

RI PT

583

601

In this study, the intrinsic transporting properties of BLG were utilized to

602

develop nano-sized green delivery systems. The binding analysis suggested that the

603

binding occurred under all conditions but varied as a function of pH and nutraceutical 25

ACCEPTED MANUSCRIPT model compound. From those observations, it could be concluded that the four

605

nutraceutical models investigated are bound to proteins on different binding sites

606

and/or by different binding mechanisms. The binding capacities of BLG mediated by

607

pH were evident in the present study. Upon lowering the pH, the binding constants

608

changed depending on the nutraceutical nature. Whereas NMR enables a more

609

direct determination of the binding to proteins, this technique suffers from the fact

610

that it can only be applied for ligands with a sufficiently large solubility (i.e. in the mM

611

range) in the aqueous phase. These findings resulted in designing nanoscopic

612

delivery systems for encapsulation of both hydrophilic and hydrophobic bioactives in

613

liquid food products. The preliminary stability experiments demonstrated the efficacy

614

of soluble complexes arising from protein-polysaccharide interactions on

615

nanoencapsulation and colloidal stability of nutraceuticals of low solubility in water.

616

Nevertheless, further work is required in the future to determine the bioprotection

617

efficiency, morphology of the stable biopolymer nanocomplexes, stability toward

618

processing conditions and controlled release properties.

619

Acknowledgment

SC

M AN U

TE D

EP

620

RI PT

604

The authors are thankful to University of Tehran, Iranian Nanotechnology Initiative Council and Ghent University for financial support.

622

References

623

Bedie, G. K., Turgeon, S. L., & Makhlouf, J. (2008). Formation of native whey protein

AC C

621

624

isolate-low methoxyl pectin complexes as a matrix for hydro-soluble food

625

ingredient entrapment in acidic foods. Food Hydrocolloids, 22, 836–844.

26

ACCEPTED MANUSCRIPT 626

Busti, P., Gatti, C. A., & Delorenzi, N. J. (2006). Binding of alkylsulfonate ligands to

627

bovine β-lactoglobulin: effects on protein thermal unfolding. Food Research

628

International, 39, 503–509. Choi, M. J., Ruktanonchai, U., Min, S. G., Chun, J. Y., & Soottitantawat, A. (2010).

RI PT

629

Physical characteristics of fish oil encapsulated by β-cyclodextrin using an

631

aggregation method or polycaprolactone using an emulsion–diffusion method.

632

Food Chemistry, 119, 1694–1703.

633

SC

630

Cogan, U., Kopelman, M., Mokady, S., & Shinitzky, M. (1976). Binding affinities of retinol and related compounds to retinol binding proteins. European Journal of

635

Biochemistry, 65, 71–78.

636

M AN U

634

Connell, M. A., Bowyer, P. J., Bone, P. A., Davis, A. L., Swanson, A. G., Nilsson, M., & Morris, G. A., (2009). Improving the accuracy of pulsed field gradient NMR

638

diffusion experiments: correction for gradient non-uniformity. Journal of

639

Magnetic Resonance, 198, 121–131.

641

Cooper, C. L., Dubin, P. L. Kayitmazer, A. B., & Turksen, S. (2005). Polyelectrolyte–

EP

640

TE D

637

protein complexes. Current Opinion in Colloid & Interface Science, 10, 52–78. Davidov-Pardo, G., & McClements, D. J. (2015). Nutraceutical delivery systems:

643

resveratrol encapsulation in grape seed oil nanoemulsions formed by

644

spontaneous emulsification. Food Chemistry, 167, 205–212.

645 646

647 648

AC C

642

DeFelice, S. (1995). The nutraceutical revolution: its impact on food industry R&D. Trends in Food Science and Technology, 6, 59–61. Dufour, E., & Haertlé, T. (1990). Alcohol-induced changes of β-lactoglobulin A-retinol binding stoichiometry. Protein Engineering, 4, 185–190. 27

ACCEPTED MANUSCRIPT 649

Dufour, E., & Haertlé, T. (1991). Binding of retinoids and β-carotene to β-

650

lactoglobulin. Influence of protein modifications. Biochimica et Biophvsica Acta,

651

1079, 316–320. Edwards, P. B., Creamer, L. K., & Jameson, G. B. (2009). Structure and stability of

RI PT

652 653

whey proteins. In A. Thompson, M. Boland, & H. Singh (Eds.).Milk proteins:

654

from expression to food. (pp. 163–204). New York: Elsevier.

Eisenthal, R., & Cornish-Bowden, A. (1974).The direct linear plot, a new graphical

SC

655

procedure for estimating enzyme kinetic parameters. Biochemistry Journal,

657

139, 715–720.

658

M AN U

656

Fioramonti, S. A., Perez, A. A., Aríngoli, E. E., Rubiolo, A. C., & Santiago, L. G. (2014). Design and characterization of soluble biopolymer complexes produced

660

by electrostatic self-assembly of a whey protein isolate and sodium alginate.

661

Food Hydrocolloids, 35, 129–136.

662

TE D

659

Forrest, S. A., Yada, R. Y., & Rousseau, D. (2005). Interactions of Vitamin D3 with Bovine β-Lactoglobulin A and β-Casein. Journal of Agricultural and Food

664

Chemistry, 53, 8003–8009.

666

667 668

Fox, P. F. (2009). Milk: an overview. In A. Thompson, M. Boland, & H. Singh (Eds.).

AC C

665

EP

663

Milk proteins: from expression to food. (pp. 1–54). New York: Elsevier. Frapin, D., Dufour, E., & Haertlé, T. (1993). Probing the fatty acid binding site of βlactoglobulins. Journal of Protein Chemistry, 12, 443–449.

669

Girard, M., Turgeon, S. L., & Gauthier, S. F. (2003). Thermodynamic parameters of

670

β-lactoglobulin-pectin complexes assessed by isothermal titration calorimetry.

671

Journal of Agricultural and Food Chemistry, 51, 4450–4455. 28

ACCEPTED MANUSCRIPT 672

Gottschalk, M., Nilsson, H., Roos, H., & Halle, B. (2003). Protein self-association in

673

solution: the bovine β-lactoglobulin dimer and octamer. Protein Science, 12,

674

2404–2411. Harnsilawat, T., Pongsawatmanit, R., & McClements, D. J. (2006). Characterization

RI PT

675

of β-lactoglobulin–sodium alginate interactions in aqueous solutions: a

677

calorimetry, light scattering, electrophoretic mobility and solubility study. Food

678

Hydrocolloids, 20, 577–585.

679

SC

676

Harnsilawat, T., Pongsawatmanit, R., & McClements, D. J. (2006). Characterization of β-lactoglobulin–sodium alginate interactions in aqueous solutions: A

681

calorimetry, light scattering, electrophoretic mobility and solubility study. Food

682

Hydrocolloids, 20, 577–585.

683

M AN U

680

Hosseini, S. M. H., Emam-Djomeh, Z., Razavi, S. H., Moosavi-Movahedi, A. A., Saboury, A. A., Mohammadifar, M. A., Farahnaky, A., Atri, M. S., & Van der

685

Meeren, P. (2013a). Complex coacervation of β-lactoglobulin – κ-Carrageenan

686

aqueous mixtures as affected by polysaccharide sonication. Food Chemistry,

687

141, 215–222.

EP

Hosseini, S. M. H., Emam-Djomeh, Z., Razavi, S. H., Moosavi-Movahedi, A. A.,

AC C

688

TE D

684

689

Saboury, A. A., Atri, M. S., & Van der Meeren, P. (2013b). β-Lactoglobulin-

690

sodium alginate interaction as affected by polysaccharide depolymerization

691

using high intensity ultrasound. Food Hydrocolloids, 32, 235-244.

692

Jerschow, A., & Muller, N. (1997). Suppression of convection artefacts in stimulated-

693

echo diffusion experiments, double-stimulated-echo experiments. Journal of

694

Magnetic Resonance, 125, 372–375.

29

ACCEPTED MANUSCRIPT 695

Jiménez-Colmenero, F. (2013). Potential applications of multiple emulsions in the

696

development of healthy and functional foods. Food Research International, 52,

697

64–74. Kailasapathy, K. (2008). Encapsulation and controlled release of folic acid. In N.

699

Garti (Ed.), Delivery and controlled release of bioactives in foods and

700

nutraceuticals (pp. 331–343). Florida: CRC Press.

Kontopidis, G., Holt, C., & Sawyer, L. (2004). Invited review: beta-lactoglobulin:

SC

701

RI PT

698

binding properties, structure, and function. Journal of Dairy Science, 87, 785–

703

796.

704

M AN U

702

Liang, L., & Subirade, M. (2010). β-Lactoglobulin/folic acid complexes: formation,

705

characterization, and biological implication. The Journal of Physical

706

Chemistry B, 114, 6707–6712.

710

711 712 713

TE D

709

with resveratrol and its biological implications. Biomacromolecules, 9, 50–56. Liu, H. C., Chen, W. L., & Mao, S. J. T. (2007). Antioxidant nature of bovine milk

EP

708

Liang, L., Tajmir-Riahi, H. A., & Subirade, M. (2008). Interaction of β-lactoglobulin

beta-lactoglobulin. Journal of Dairy Science, 90, 547–555.

AC C

707

Livney, Y. D. (2010). Milk proteins as vehicles for bioactives. Current Opinion in Colloid & Interface Science, 15, 73–83. Matalanis, A., Jones, O. G., & McClements, D. J. (2011). Structured biopolymer-

714

based delivery systems for encapsulation, protection, and release of lipophilic

715

compounds. Food Hydrocolloids, 25, 1865–1880.

716 717

McClements, D. J. (2006). Non-covalent interactions between proteins and polysaccharides. Biotechnology Advances, 24, 621–625. 30

ACCEPTED MANUSCRIPT 718

McClements, D. J., Decker, E. A., Park, Y., & Weiss, J. (2009). Structural design

719

principles for delivery of bioactive components in nutraceuticals and functional

720

foods. Critical Reviews in Food Science and Nutrition, 49, 577–606.

721

Mohammadi, F., Bordbar, A-K., Divsalar, A., Mohammadi, K., & Saboury, A. A. (2009). Interaction of curcumin and diacetylcurcumin with the lipocalin member

723

β-lactoglobulin. Protein Journal, 28, 117–123.

724

RI PT

722

Mohan Reddy, I., Kella, N. K. D., & Kinsella, J. E. (1988). Structural and

conformational basis of the resistance of β-lactoglobulin to peptic and

726

chymotryptic digestion. Journal of Agricultural and Food Chemistry, 36, 737–

727

747.

M AN U

728

SC

725

Pandita, D., Kumar, S., Poonia, N., & Lather, V. (2014). Solid lipid nanoparticles enhance oral bioavailability of resveratrol, a natural polyphenol. Food Research

730

International, 62, 1165–1174.

731

TE D

729

Peinado, I., Lesmes, U., Andrés, A., & McClements D. J. (2010). Fabrication and morphological characterization of biopolymer particles formed by electrostatic

733

complexation of heat treated lactoferrin and anionic polysaccharides. Langmuir,

734

26, 9827–9834.

AC C

735

EP

732

Qian, C., Decker, E. A., Xiao, H., & McClements, D. J. (2012). Physical and chemical

736

stability of β-carotene-enriched nanoemulsions: influence of pH, ionic strength,

737

temperature, and emulsifier type. Food Chemistry, 132, 1221–1229.

738

Qin, B. Y., Bewley, M. C., Creamer, L. K., Baker, H. M., Baker, E. N., & Jameson, G.

739

B. (1998). Structural basis of the Tanford transition of bovine β-lactoglobulin.

740

Biochemistry, 37, 14014–14023.

31

ACCEPTED MANUSCRIPT 741

Ragona, L., Fogolari, F., Catalano, M., Ugolini, R., Zetta, L., & Molinari, H. (2003).

742

EF loop conformational change triggers ligand binding in β-lactoglobulins.

743

Journal of Biological Chemistry, 278, 38840–38846. Ragona, L., Fogolari, F., Zetta, L., Perez, D. M., Puyol, P., De Kruif, K., Lohr, F.,

745

Ruterjans, H., & Molinari, H. (2000). Bovine β-lactoglobulin: interaction studies

746

with palmitic acid. Protein Science, 9, 1347–1356.

Rao, J., & McClements, D. J. (2012). Food-grade microemulsions and

SC

747

RI PT

744

nanoemulsions: role of oil phase composition on formation and stability. Food

749

Hydrocolloids, 29, 326–334.

750

M AN U

748

Relkin, P., Eynard, L. & Launay, B. (1992). Thermodynamic parameters of β-

751

lactoglobulin and α-lactalbumin, A DSC study of denaturation by heating.

752

Thermochimica Acta, 204, 111–121.

Riihimäki, L. H., Vainio, M. J., Heikura, J. M., Valkonen, K. H., Virtanen, V. T., &

TE D

753

Vuorela P. M. (2008). Binding of phenolic compounds and their derivatives to

755

bovine and reindeer β-lactoglobulin. Journal of Agricultural and Food

756

Chemistry, 56, 7721–7729. Ron, N., Zimet, P., Bargarum, J., & Livney, Y. D. (2010). Beta-lactoglobulin-

AC C

757

EP

754

758

polysaccharide complexes as nanovehicles for hydrophobic nutraceuticals in

759

non-fat foods and clear beverages. International Dairy Journal, 20, 686–693.

760

Saberi, A. H., Fang, Y., & McClements, D. J. (2014). Stabilization of vitamin E-

761

enriched mini-emulsions: influence of organic and aqueous phase

762

compositions. Colloids and Surfaces A: Physicochemical and Engineering

763

Aspects, 449, 65–73. 32

ACCEPTED MANUSCRIPT 764 765

766

Sagalowicz, L., & Leser, M. E. (2010). Delivery systems for liquid food products. Current Opinion in Colloid & Interface Science, 15, 61–72. Sakurai, K., & Goto, Y. (2002).Manipulating monomer–dimer equilibrium of bovine βlactoglobulin by amino acid substitution.Journal of Biological Chemistry, 277,

768

25735–25740.

769

RI PT

767

Sanchez, C., Mekhloufi, G., Schmitt, C., Renard, D., Robert, P., Lehr, C.-M.,

Lamprecht, A., & Hardy, J. (2002). Self-assembly of β-lactoglobulin and acacia

771

gum in aqueous solvent: structure and phase-ordering kinetics. Langmuir, 18,

772

10323–10333.

M AN U

773

SC

770

Santos, M. G., Bozza, F. T., Thomazini, M., & Favaro-Trindade, C. S. (2015).

774

Microencapsulation of xylitol by double emulsion followed by complex

775

coacervation. Food Chemistry, 171, 32–39.

Schmitt, C., Aberkane, L., & Sanchez, C. (2009). Protein-polysaccharide complexes

TE D

776

and coacervates. In G. O. Phillips, & P. A. Williams (Eds.), Handbook of

778

hydrocolloids (2nd edition, pp. 420–476).Florida: CRC Press.

779

EP

777

Schmitt, C., Sanchez, C., Desobry-Banon, S., & Hardy, J. (1998). Structure and technofunctional properties of protein-polysaccharide complexes: a review.

781

Critical Reviews in Food Science and Nutrition, 38, 689–753.

782

AC C

780

Semo, E., Kesselman, E., Danino, D., & Livney, Y. D. (2007). Casein micelle as a

783

natural nano-capsular vehicle for nutraceuticals. Food Hydrocolloids, 21, 936–

784

942.

33

ACCEPTED MANUSCRIPT 785

Shapira, A., Assaraf, Y. G., & Livney, Y. D. (2010). Beta-casein nanovehicles for oral

786

delivery of chemotherapeutic drugs’.Nanomedicine-Nanotechnology Biology

787

and Medicine, 6, 119–126. Shimoni, E. (2009). Nanotechnology for foods: delivery systems. In G. Barbosa-

RI PT

788

Cánovas, A. Mortimer, D. Lineback, W. Spiess, K. Buckle, & P. Colonna (Eds.),

790

Global Issues in Food Science and Technology (pp. 411–424). New York:

791

Elsevier.

792

SC

789

Sneharani, A. H., Karakkat, J. V., Singh, S. A., & Rao, A. G. (2010). Interaction of curcumin with β-lactoglobulin–stability, spectroscopic analysis, and molecular

794

modeling of the complex. Journal of Agricultural and Food Chemistry, 58,

795

11130–11139.

796

M AN U

793

Takahashi, M., Inafuku, K., Miyagi, T., Oku, H., Wada, K., Imura, T., & Kitamoto, D. (2007). Efficient preparation of liposomes encapsulating food materials using

798

lecithins by a mechanochemical method. Journal of Oleo Science, 56, 35–42.

801

EP

800

Tolstoguzov, V. B. (2003). Some thermodynamic considerations in food formulation. Food Hydrocolloids, 17, 1–23. Tomczy ska-Mleko, M., & Mleko, S. (2014). Whey protein aerated gels as matrices

AC C

799

TE D

797

802

for controlled mineral release in simulated gastric conditions. Food Research

803

International, 62, 91–97.

804

Turgeon, S. L., & Laneuville, S. I. (2009). Protein + polysaccharide coacervates and

805

complexes: from scientific background to their application as functional

806

ingredients in food products. In S. Kasapis, I. T. Norton, & J. B. Ubbink (Eds.),

807

Modern biopolymer science (pp. 327–363). New York: Elsevier. 34

ACCEPTED MANUSCRIPT 808 809

810

Wang, Q., Allen, J. C., & Swaisgood, H. E. (1997). Binding of vitamin D and cholesterol to β-lactoglobulin. Journal of Dairy Science, 80, 1054–1059. Wang, Q., Allen, J. C., & Swaisgood, H. E. (1998). Protein concentration dependence of palmitate binding to β-lactoglobulin A. Journal of Dairy Science,

812

81, 76–81.

813

RI PT

811

Wang, Q., Allen, J. C., &Swaisgood, H. E. (1999). Binding of lipophilic nutrients to βlactoglobulin A prepared by bioselective adsorption. Journal of Dairy Science,

815

82, 257–264.

Yoksan, R., Jirawutthiwongchai, J., & Arpo, K. (2010). Encapsulation of ascorbyl

M AN U

816

SC

814

817

palmitate in chitosan nanoparticles by oil-in-water emulsion and ionic gelation

818

processes. Colloids and Surfaces B, 76, 292–297.

819

Younis, I. R., Stamatakis, M. K.,Callery, P. S., & Meyer-Stouta, P. J. (2009). Influence of pH on the dissolution of folic acid supplements. International

821

Journal of Pharmaceutics, 367, 97–102. Zabar, S., Lesmes, U., Katz, I., Shimoni, E., & Bianco-Peled, H. (2009). Studying

EP

822

TE D

820

different dimensions of amylose–long chain fatty acid complexes: molecular,

824

nano and micro level characteristics. Food Hydrocolloids, 23, 1918–1925.

825

Zhao, Y., Li, F., Carvajal, M. T., & Harris, M. T. (2009). Interactions between bovine

826

serum albumin and alginate: An evaluation of alginate as protein carrier.

827

Journal of Colloid and Interface Science, 332, 345–353.

828

AC C

823

Zheng, Z., Zhang, X., Carbo, D., Clark, C., Nathan, C-A., & Lvov, Y. (2010).

829

Sonication-assisted synthesis of polyelectrolyte-coated curcumin nanoparticles.

830

Langmuir, 26, 7679–7681. 35

ACCEPTED MANUSCRIPT 831

Zimet, P., & Livney, Y. D. (2009). Beta-lactoglobulin and its nanocomplexes with

832

pectin as vehicles for ω-3 polyunsaturated fatty acids. Food Hydrocolloids, 23,

833

1120–1126. Zimet, P., Rosenberg, D., & Livney, Y. D. (2011). Re-assembled casein micelles and

835

casein nanoparticles as nano-vehicles for ω-3 polyunsaturated fatty acids.

836

Food Hydrocolloids, 25, 1270–1276.

AC C

EP

TE D

M AN U

SC

837

RI PT

834

36

ACCEPTED MANUSCRIPT 838

Table captions

839

Table 1: Apparent dissociation constant (K'd, expressed in nM) of BLG and

841

nutraceutical model compounds at pH 7.00 and 4.25.

RI PT

840

842

Table 2: Apparent molar binding ratios (n, expressed in mol ligand per mol of protein

844

monomer) of nutraceutical model compounds to BLG at pH 7.00 and 4.25.

AC C

EP

TE D

M AN U

SC

843

37

ACCEPTED MANUSCRIPT 845

Table 1

846 K'd (nM) β-carotene

Folic acid

Curcumin

Ergocalciferol

7.00

21 ± 3Ba

34 ± 27Aa

201 ± 72Bb

144 ± 28Ab

4.25

15 ± 2Aa

27 ± 4Ab

51 ± 17Ac

847

173 ± 16Ad

SC

848

RI PT

pH

The superscript letters (A, B) mean that the results within the same

850

column without a common letter are significantly different (p< 0.05); the

851

subscript letters (a, b, c, d) mean that the results within the same row

852

without a common letter are significantly different (p< 0.05).

M AN U

849

AC C

EP

TE D

853

38

ACCEPTED MANUSCRIPT 854

Table 2

855 N β-carotene

Folic acid

Curcumin

Ergocalciferol

7.00

0.48 ± 0.04Ba

1.25 ± 0.05Bc

0.95 ± 0.06Ab

1.31 ± 0.12Bc

4.25

0.23 ± 0.03Aa

0.39 ± 0.04Ab

0.82 ± 0.08Ac

RI PT

pH

856 857

0.85 ± 0.06Ac

The superscript letters (A, B) mean that the results within the same

859

column without a common letter are significantly different (p< 0.05); the

860

subscript letters (a, b, c) mean that the results within the same row without

861

a common letter are significantly different (p< 0.05).

M AN U

SC

858

AC C

EP

TE D

862

39

ACCEPTED MANUSCRIPT 863

Figure Captions

864

Fig. 1: 3D illustration of asymmetric dimer of BLG showing the hydrophobic

866

binding sites

867

Fig. 2: Folic acid chemical structure

868

Fig. 3: NMR spectrum of 3 mM folic acid in D2O at 25 °C with relative peak

869

assignment

870

Fig. 4: NMR spectra of 3 mM folic acid in D2O at 25 °C: alone (blue), in

871

presence of 0.25 mM BLG (red) and in presence of BLG at pH 4.25 (green)

872

Fig. 5: DOSY spectra of 3 mM folic acid in D2O at 25 °C: alone (blue), in presence of

873

0.25 mM BLG (red)

874

Fig. 6: Effects of curcumin nanoencapsulation on its colloidal stability at pH 4.25 (I:

875

dissolved curcumin in ethanol added to deionized water; II: dissolved curcumin in

876

ethanol added to BLG dispersion (0.1% w/w); III: dissolved curcumin in ethanol

877

added to BLG-Na-ALG soluble complexes (Na-ALG/BLG weight ratio of 0.75).

878

Fig. 7: Effects of folic acid nanoencapsulation on its colloidal stability at pH 4.25

879

(I: dissolved folic acid in deionized water added to deionized water; II: dissolved

880

folic acid in deionized water added to BLG dispersion (0.1% w/w); III: dissolved

881

folic acid in deionized water added to BLG-Na-ALG soluble complexes (Na-

882

ALG/BLG weight ratio of 0.75).

883

Fig. 8: Increasing in chain flexibility as a result of mutual neutralization

AC C

EP

TE D

M AN U

SC

RI PT

865

40

ACCEPTED MANUSCRIPT

M AN U

SC

RI PT

884

885 886

890 891

892

893

894

Fig. 1

EP

889

AC C

888

TE D

887

895

41

ACCEPTED MANUSCRIPT 896 OH N

HO

7 H 2N

N

H

H

H

O

4 H N

H

HN

HN

H

O

H

H

2

8 H

6

H

5

SC

897

M AN U

898 899

Fig. 2

900

AC C

EP

TE D

901

H

RI PT

N

42

H 3

HO

1 O

ACCEPTED MANUSCRIPT 902

7

6

5

RI PT

8

2(3)-3(2)

SC

4

M AN U

903 904 905

EP

908

Fig. 3

AC C

907

TE D

906

43

1

ACCEPTED MANUSCRIPT

SC

RI PT

909

M AN U

910 911 912

EP

915

Fig. 4

AC C

914

TE D

913

44

ACCEPTED MANUSCRIPT

SC

RI PT

916

M AN U

917 918 919

EP

922

Fig. 5

AC C

921

TE D

920

45

ACCEPTED MANUSCRIPT 923

A III

I

II

I

III

RI PT

II

SC

I

B

Just after production

After 24 hours D

II

III

TE D

I

926 927 928

AC C

925

EP

After 48 hours 924

II

M AN U

C

Fig. 6

929 930 931

46

After 30 days

III

ACCEPTED MANUSCRIPT A III

C

D

II

III

935 936

I

II

After 30 days

TE D EP

934

III

RI PT

After 24 hours

Fig. 7

AC C

933

II

Just after production

After 48 hours 932

I

SC

I

II

M AN U

I

B

47

III

ACCEPTED MANUSCRIPT 937

Junction zone

-

-

+

- + +

-

-

-

-

M AN U

Anionic polysaccharide

+ -

Globular protein

SC

-

RI PT

Chain segment of increased hydrophobicity and flexibility as a result of mutual neutralization

-

Inflexible hydrophilic chain segment

941 942

EP

940

Fig. 8

AC C

939

TE D

938

48

ACCEPTED MANUSCRIPT

Utilizing intrinsic transporting property of BLG to develop green delivery system BLG-bioactives complexation varied as a function of pH and

RI PT

nutraceutical type.

Successful entrapment of bioactives within electrostatically stable

AC C

EP

TE D

M AN U

SC

nanocomplexes

1