Properties of whey protein concentrate powders obtained by spray drying of sweet, salty and acid whey under varying storage conditions

Properties of whey protein concentrate powders obtained by spray drying of sweet, salty and acid whey under varying storage conditions

Accepted Manuscript Properties of whey protein concentrate powders obtained by spray drying of sweet, salty and acid whey under varying storage condit...

2MB Sizes 32 Downloads 213 Views

Accepted Manuscript Properties of whey protein concentrate powders obtained by spray drying of sweet, salty and acid whey under varying storage conditions Manjula Nishanthi, Jayani Chandrapala, Todor Vasiljevic PII:

S0260-8774(17)30278-9

DOI:

10.1016/j.jfoodeng.2017.06.032

Reference:

JFOE 8937

To appear in:

Journal of Food Engineering

Please cite this article as: Manjula Nishanthi, Jayani Chandrapala, Todor Vasiljevic, Properties of whey protein concentrate powders obtained by spray drying of sweet, salty and acid whey under varying storage conditions, Journal of Food Engineering (2017), doi: 10.1016/j.jfoodeng.2017.06.032 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 1

Properties of whey protein concentrate powders obtained by spray drying of sweet,

2

salty and acid whey under varying storage conditions

3

Manjula Nishanthi1, Jayani Chandrapala1,2, Todor Vasiljevic1*

5

6

1

7

University, Werribee campus, Victoria 3030, Australia

8

2

SC

Advanced Food Systems Research Unit, College of Health and Biomedicine, Victoria

M AN U

School of Science, RMIT University, Bundoora, Victoria 3083, Australia

9 10 11

*Corresponding Author:

Tel: +61 3 9919 8062

Email: [email protected]

12

17 18

EP

16

AC C

15

TE D

13 14

RI PT

4

ACCEPTED MANUSCRIPT Abstract

20

Changes of secondary structure and protein interactions of whey protein (WPs) present in

21

native, sweet, acid and salty-WPCs were analyzed following storage at 4, 25 or 45°C and 22

22

or 33% relative humidity for a period of 90-days. WPs aggregated predominantly through

23

covalent crosslinking, achieving maximum at 45 °C and 33% RH. Greater participation of β-

24

lactoglobulin in covalent crosslinking was evident in all WPCs, while that of α-lactalbumin

25

was significantly (p<0.05) high in acid-WPC only. Reaction order of β-LG denaturation in

26

acid and salty-WPCs was approximately 2 while approximately 1 in native and sweet-WPCs.

27

Activation energy was significantly (p<0.05) higher in native and sweet-WPCs, with

28

averages recorded as 97 and 49.8 kJ mol-1, respectively, than that in acid and salty-WPCs with

29

averages of 27.5 and 33.8 kJ mol-1, respectively, mainly attributed to inherently high

30

concentrations of lactic acid and salts in these WPCs.

31

Key word: Whey protein concentrate, protein aggregation, covalent interactions, acid whey,

32

salty whey

SC

M AN U

TE D EP

AC C

33

RI PT

19

ACCEPTED MANUSCRIPT Introduction:

35

Whey protein concentrates (WPCs) are an important protein source used in many food

36

systems due to their excellent functional characteristics such as emulsification, gelling and

37

foaming. Dairy industry has been successful in utilizing the sweet whey waste stream for

38

manufacturing of various WPCs. However, the industry has not been very successful in

39

utilizing acid and salty whey in an equivalent way to sweet whey, mainly due to their

40

composition-induced processing obstacles such as high acidity and salinity, respectively.

41

Moreover, the powders produced from these two streams form clumps and sticky deposits

42

during storage. A recent study by Nishanthi et al. (2017a) showed that the particle surface of

43

salty-WPC was hydrophobic mainly due to high surface fat content, while that of acid-WPC

44

was characterized by both hydrophobic and hydrophilic nature due to presence of elevated

45

protein concentration. Therefore, these two WPCs possess a high potential to be used as

46

functional ingredients in food systems. Thus, a thorough fundamental understanding on the

47

stability of proteins present in these respective WPCs as affected by storage is a necessity in

48

evaluating the potential of using these two particular WPCs.

49

Storage conditions induce changes in protein conformation at a molecular level, which in turn

50

may affect functional properties of the protein. Intermolecular covalent and non-covalent

51

interactions occur between whey proteins (WPs), resulting in protein unfolding and

52

denaturation, eventually leading to reversible and irreversible aggregation. As a result, a

53

proportion of native WPs decreases after 3 months of storage at 40 °C with a concomitant

54

increase in proportion of denatured proteins. These changes were clearly temperature

55

dependent as aforementioned changes were not noticed at 4 °C or 20 °C up to 15 months of

56

storage (Norwood et al., 2016). At elevated storage temperatures, flexibility and mobility of

57

WPs increases, resulting in partial loss of their tertiary structure (Norwood et al., 2016). WPC

58

was found to have increased surface hydrophobicity and less exposed amino acids on the

AC C

EP

TE D

M AN U

SC

RI PT

34

ACCEPTED MANUSCRIPT particle surfaces after storage at 60 °C and 22 % RH for one month (Burgain et al., 2016). In

60

addition to denaturation, WPs can be subjected to lactosylation during storage. While studies

61

on lactosylation of WPC powders are lacking, Le et al. (2012) reported that six out of twelve

62

lysine residues of α-lactalbumin (α-LA) would lactosylate during storage of milk protein

63

concentrates up to 12 weeks at 40 °C and 84 % RH, which indicates that similar changes

64

could be expected in WPC, especially those containing greater lactose content. These

65

structural modifications in WPs could subsequently affect their functional properties. For

66

instance, formation of insoluble aggregates in response to elevated storage temperatures

67

reduced the solubility of WPC (Zhou et al., 2008) affecting several functional characteristics

68

such as emulsification, gelation and foaming.

69

So far, the studies mainly focused on WPC produced from the sweet whey and no study thus

70

far has established physico-chemical properties of WPs during storage of WPCs produced

71

from acid or salty whey, which were identified as potentially effective functional ingredients

72

(Nishanthi et al., 2017a). Therefore, the main aim of the current study was to establish

73

conformational and surface interactional changes of WPs present in sweet, acid and salty

74

WPCs under three different storage temperatures (4, 25, 45 °C) and two relative humidities

75

(RH) (22% and 33%) over a storage period of 90 days. Native-WPC was used as the

76

reference sample.

77

2. Methodology

78

2.1 Sample preparation

79

Four WPCs, native-WPC, sweet-WPC, acid-WPC and salty-WPC were manufactured in our

80

previous study (Nishanthi et al., 2017a) with the composition shown in Table 1 and used

81

during storage trial in the current study. WPCs were first completely dehydrated by placing

AC C

EP

TE D

M AN U

SC

RI PT

59

ACCEPTED MANUSCRIPT appropriate quantity of samples in vacuum desiccators over phosphorous pentoxide at room

83

temperature (25°C) for 4 weeks.

84

2.2 Storage trial

85

Approximately 3 g of each WPC was placed in standard desiccators (with a porcelain disc

86

size of 250 mm) with a controlled RH (22 and 33%), sealed and placed in temperature

87

controlled cabinets at three temperatures (4, 25 and 45°C). The equilibrium RH in each

88

desiccator was maintained using saturated salt slurries containing potassium acetate (22 ±

89

2%) or magnesium chloride (33 ± 1%). At 4, 25 and 45 °C, magnesium chloride maintained

90

RH of 34, 33 and 32%, while potassium acetate maintained RH of 24, 23 and 23%,

91

respectively. Five desiccators were prepared for each temperature/humidity combination as

92

sampling was conducted at 5 time points representing 0, 14, 30, 60 and 90 days of storage. A

93

pre-trial test was conducted to estimate the equilibrium time and identify the first sampling

94

point. Several samples from each WPC were stored in standard desiccators with saturated salt

95

slurries of magnesium chloride and potassium acetate. After one week of storage, one sample

96

from each WPC was tested for the water activity using a water activity meter (Aqualab CX-

97

3T, Labcell, Hampshire, United Kingdom) at 25°C each day. The four types of WPCs reach

98

to the water activity of 0.22 and 0.33 by one week and four days. A stabilised water activity

99

level could be achieved within two weeks. Therefore, the first sampling (0 day) was

100

conducted after allowing 2-weeks equilibrium period from the initiation date of the storage

101

trial. After opening desiccators at a required time point, the analysis was carried out within 3

102

days.

103

2.3 Fourier Transform Infrared Spectroscopy (FTIR)

104

The FTIR spectra of WPCs were obtained in the range of 4000 - 600 cm−1 using the

105

PerkinElmer Frontier FTIR spectrometer (PerkinElmer, MA, USA) as described previously

106

(Nishanthi et al., 2017b). The spectra were vector normalized, smoothened and deconvoluted

AC C

EP

TE D

M AN U

SC

RI PT

82

ACCEPTED MANUSCRIPT with the aid of IR Solution software (Shimadzu Corporation, Kyoto, Japan), in order to

108

recognize the corresponding peaks within the amide I region between 1600 and 1700 cm−1.

109

2.4 Sodium Dodecyl Sulphate Gel Electrophoresis (SDS – PAGE)

110

In order to establish non-covalent and covalent interactions, Sodium Dodecyl Sulphate (SDS)

111

polyacrylamide gel electrophoresis (PAGE) was performed under reducing (SDSR) and non-

112

reducing (SDSNR) conditions using freshly casted gels with 30% (w/v) acrylamide and 10%

113

(w/v) SDS (Nishanthi et al., 2017b). Samples were reduced with addition of β-

114

mercaptoethanol and subsequent heating at 100 °C for 5 min.

115

2.5 Kinetics analysis

116

The rate (1) and the Arrhenius equations (2) were used to calculate the kinetic parameters

117

involved in disappearance of native WPs during storage of WPCs (Oldfield et al., 2005).

118

−  =   (1)

119

Where:

120

C = Protein concentration (g kg-1)

121

t = time (days)

122

kn = rate constant ((g kg-1)(1-n) day-1)

123

n = reaction order

124



125

 =   (2)

126

Where:

127

k0 = pre-exponential term ((g kg-1)(1-n) day-1)

128

Ea = activation energy (kJ mol-1)

129

R = universal Gas Constant (8.314 J mol-1 K-1)

130

T = temperature (K)

131

2.6 Statistical analysis

132

The experiments were organized in a randomized split plot blocked design with 3 main

133

factors: a type of WPC as the main plot, and temperature and RH as subplots. The

M AN U

SC

RI PT

107

EP

AC C



TE D



ACCEPTED MANUSCRIPT replications served as blocks. Replicated samples were sub sampled and analyzed. The results

135

of the kinetic and compositional analysis were statistically analyzed using the General Linear

136

Model (GLM) and ANOVA procedure, respectively. The level of significance was preset at

137

p<0.05.

138

3. Results and discussion:

139

3.1 Secondary structural changes during storage through FTIR

140

Significant and complex intermolecular interactions occur among WPs in response to storage

141

conditions and associated compositional changes of examined WPCs. These changes affect

142

the conformation of secondary structure motifs such as α-helices, β-sheets, β-turns and

143

random coils in these proteins as shown in Figures 1 to 4.

144

At the beginning of the storage trial, WPs present in native-WPC were characterized as β-

145

sheet rich proteins due to two prominent peaks observed in the regions of 1628-1629 cm-1

146

and 1693 cm-1 (Figure 1). However, other secondary structures, such as α-helixes, random

147

peptides and β-turn were largely missing at 4 °C, but appeared with low intensities at 25 and

148

45 °C, irrespective of RH. In addition to the native moieties, β-sheet cross linking (1682 cm-1

149

and 1615 cm-1) observed at the beginning of storage can be attributed to the thermal impacts

150

associated with pasteurization and spray drying applied during processing of this WPC.

151

During initial stages of the storage (14-30 days), secondary conformations did not change

152

substantially. However, after 60 days of storage WPs underwent several major structural

153

changes. These changes were more pronounced at 33% RH as compared to that at 22% RH,

154

but were independent of storage temperature. Firstly, complete distortion of α-helixes

155

occurred under all storage conditions, which in turn increased the intensity of randomly

156

arranged peptides (1644 cm-1). In presence of high amount of moisture (33 % RH),

157

hydrophilic amino acid residues of α-helix form hydrogen bonds with water molecules, thus

158

disturbing its intermolecular interactions. As a result of α-helical stretching, intensity of β-

AC C

EP

TE D

M AN U

SC

RI PT

134

ACCEPTED MANUSCRIPT turns (1660 cm-1) increased attributed to the broadening of β-turn between β-strand I and H in

160

β-lactoglobulin (β-LG). Secondly, widening of peaks attributed to β-sheets, observed

161

prominently at 33% RH, suggests a possible intermolecular cross linking in the presence of

162

elevated moisture content. Concomitantly, intensity of anti-parallel β-sheets (Lefèvre and

163

Subirade, 1999) (1691 cm-1) increased, suggesting a dimerization or polymerization, moving

164

high number of β-sheets to the molecular interior. Thirdly, greater exposure of amino acid

165

(AA) side chain residues (1610 cm-1) (Lefèvre and Subirade, 1999) were observed mainly at

166

33% RH. The random coils formed by stretching α-helixes may be small with high exposed

167

surface area, which subsequently expose a high number of AA residues. Moreover, protein

168

hydration due to greater relative humidity leads to movement of hydrophilic AA residues to

169

the molecular exterior. At the end of the storage (90 days), the peaks representing β-turns and

170

random coils widened and disappeared suggesting a possible aggregation that destructed

171

native monomeric and dimeric forms of proteins. This loss of secondary structures was

172

especially magnified in the samples stored at 45 °C and 33% RH. Further, the exposed side

173

chain residues disappeared under all storage conditions. Interestingly, a greater content of β-

174

sheets remained unaffected, suggesting the preservation of the calyx of β-LG.

175

Sweet-WPC mainly contained intramolecular β-sheets (1630 and 1693 cm-1) and

176

intermolecular β-sheets (1615 cm-1) at the start of the storage (Figure 2). Interlinking of β-

177

sheets observed at the beginning of the storage suggests an impact of thermal treatments

178

associated with the manufacturing process of sweet-WPC (Nishanthi et al., 2017a). Similar to

179

native-WPC, initial storage period (14-30 days) did not change the profile of the secondary

180

structure of sweet-WPC to a great extent for most storage conditions, although random coils,

181

α-helixes and β-turns appeared in the samples stored at 33% RH. These changes can be

182

attributed to the protein hydration, leading acquisition of some native attributes. Furthermore,

183

a greater exposure of side chain AA residues was observed (1611 cm-1) at 22% RH. Similar

AC C

EP

TE D

M AN U

SC

RI PT

159

ACCEPTED MANUSCRIPT to native-WPC, substantial structural changes were noticed after 60 days of storage.

185

Irrespective of storage conditions, β-sheets lost their conformational stability, partially

186

forming cross linked β-sheets (1680 cm-1). Strong modifications were seen in β-sheets

187

indicated by appearance of two peaks (1623 and 1634 cm-1), which can be attributed to the

188

disruption of intramolecular hydrogen bonds, leading to formation of new stronger

189

intermolecular hydrogen bonds (Ngarize et al., 2004). This modification was especially

190

prominent at 25 and 45 °C. Monomer-dimer equilibrium of β-LG molecules (Lefèvre and

191

Subirade, 1999) has shifted towards the dimeric form, resulting in strongly bonded β-strand

192

structures. Dimer formation, in parallel, leads to a high number of deeply buried β-type

193

structures (1691 cm-1). Moreover, dimer formation takes place at expense of intramolecular

194

α-helix, subsequently, resulting in high number of random coils. The α-helix stretching into

195

random coils was high at 25 and 45 °C, while independent of RH. Therefore, the dimerization

196

which was established in SWP after 60 days was a result of storage temperature. By the end

197

of storage, secondary structures other than β-sheets completely disappeared in the samples

198

stored at 25 and 45 °C. Furthermore, aging of sweet-WPC after 60 days induced a

199

conformational change to the calyx of β-LG indicated by a shift in peak at 1691 cm-1 to a

200

peak with low intensity at 1694 cm-1. This change in turn may lead to the exposure of reactive

201

sites for covalent and hydrophobic interactions.

202

In acid-WPC, β-sheets (1620 – 1635 cm-1) were largely absent at the beginning of the storage

203

(Figure 3). Simultaneously, aggregation between anti-parallel β-sheets (1682 cm-1) was

204

prominent at 22 % RH and 45 °C, involving mainly β-strands located on the surface, as

205

opposed to those in deep areas of β-LG (1691 cm-1), which remained intact. Other secondary

206

structures, such as α-helixes, β-turns and random coils were present at a low intensity,

207

regardless of the storage conditions. By 14 days of storage, additional exposure of β-sheets

208

(1621 cm-1) occurred suggesting a partial unfolding of β-LG in response to lactic acid (LA)-

AC C

EP

TE D

M AN U

SC

RI PT

184

ACCEPTED MANUSCRIPT induced WP alterations (Nishanthi et al., 2017b). Unfolding of β-sheets was more prominent

210

at 33% RH as compared to that at 22% RH. Deprotonation of LA occurs rapidly in the

211

presence of moisture due to its low acid dissociation constant (pKa), which in turn decreases

212

the negative surface charge density of the proteins promoting attractive forces, inducing

213

protein aggregation driven by van der Waals and electrostatic interactions (Israelachvili,

214

2011). After 60 days of storage, three significant modifications were noticed. Firstly, α-

215

helixes disappeared completely forming non-native random structures under all storage

216

conditions. Secondly, conformation of β-sheets changed substantially, converting into

217

distinctive two peaks in the region of 1620 – 1638 cm-1, implying a dimerization of β-LG. In

218

addition, under all storage conditions the intensity of deeply located β-sheets (1693 cm-1) was

219

doubled. In presence of LA, β-LG formed dimers, relocating more of β-sheets to the inner

220

side of the molecule. Thirdly, side chain amino acid residues (1610 cm-1) were greatly

221

exposed, especially at 22% RH as a result of β-sheet modifications and α-helix disruption.

222

Towards the end of the storage (90 day), a greater loss of secondary structures was observed,

223

especially in the samples stored at 25 and 45 °C under 22% RH. A greater loss of random

224

coils and dimeric forms of β-LG suggested a severe aggregation.

225

In contrast to other WPCs, secondary structures of salty-WPC (Figure 4) showed high

226

sensitivity to storage conditions during the first days of storage. For instance, proportions of

227

β-sheets (1626, 1636 and 1693 cm-1), random peptides (1648 cm-1) and β-turns (1660 cm-1)

228

clearly declined as temperature and humidity increased. In parallel, a number of deeply

229

located β-sheets (1693 cm-1) descended with increase in temperature, denoting an exposure of

230

the β-LG core. However, the temperature-dependent reactivity of β-LG did not increase the

231

proportion of cross linked β-sheets (1616 and 1680 cm-1). Due to osmotic pressure caused by

232

Na+ in the medium, β-sheets did not interlink, instead unfolded, forming unordered structures,

233

as apparent by intensity of high random coil. A dimeric form of β-LG was also observed at 4

AC C

EP

TE D

M AN U

SC

RI PT

209

ACCEPTED MANUSCRIPT and 25 °C (1620 – 1636 cm-1), as opposed to monomeric β-LG at 45 °C. Other secondary

235

motifs were also greatly absent at 45 °C. During further storage (14-30 days), monomeric β-

236

LG dominated (Lefèvre and Subirade, 1999) under all storage conditions, suggesting that

237

native-dimers dissociated with aging. With increase in temperature, concomitant reduction of

238

non-native monomers and enhanced aggregation driven by crosslinking of β-sheets (1616 cm-

239

1

240

storage, β-sheets reappeared in the region of 1620 – 1638 cm-1, whereas the peak intensity

241

increased with the storage temperature. Interestingly, the region denoted β-sheets (1620 –

242

1636 cm-1) consisted of two peaks at 25 and 45 °C, denoting dimeric forms of β-LG, while it

243

was just a single peak at 4 °C, attributed to monomeric β-LG. Dimeric β-LG formation in

244

turn increased the number of intermolecular β-sheets (1680 cm-1), resulting in substantially

245

greater proportion of deeply located β-sheets (1691 cm-1) and high number of random coils

246

and β-turns. Up to 60 days, the observed structural changes were independent of the

247

humidity. However, towards the end of storage (90 day), humidity influenced the interactions

248

among β-sheets. High occurrence of intermolecular hydrogen bonds prominent at 22% RH

249

suggested a high occurrence of β-LG dimers (1620 – 1636 cm-1), while intramolecular

250

hydrogen bonds predominated at 33 % RH, suggesting more of monomeric form of β-LG.

251

The prolonged exposure to high humidity reduced the protein-protein interactions, instead

252

allowing protein hydration and formation of near-native protein monomers. This assumption

253

is also supported by the occurrence of less cross linked β-sheets (1615 cm-1) at 33% RH.

254

3.2 Changes in protein interactions and aggregation during storage

255

At the beginning of the storage trial, aggregates (Mw≥250 kDa) covalently linked via

256

disulphide bridges were observed in native-WPC (Figure 5 and 6) accompanied with

257

subsequent disappearance of β-LG directly related to temperature, humidity and storage time.

258

Intensity of α-LA bands on the other hand remained approximately consistent during initial

RI PT

234

AC C

EP

TE D

M AN U

SC

) occurred due to temperature-dependent exposure of reactive sites in β-LG. After 60 days of

ACCEPTED MANUSCRIPT 30 days of storage, but started to decline after this period implying that α-LA would be

260

involved in the overall aggregation process at a later stage. It is known that α-LA contains

261

only four disulphide bonds positioned at Cys6-Cys120, Cys28-Cys11, Cys61-Cys77 and Cys73-

262

Cys91 (McSweeney, 2013), but no free thiol groups. In addition, the holo-form of α-LA,

263

which is present abundantly in native-WPC, provides a high thermal stability (Boye and Alli,

264

2000). For these reasons, involvement of α-LA in overall aggregation is limited. Small

265

amount of caseins was present in this WPC, which appeared to be involved in disulphide

266

linking over the storage period, with the greatest extent at 45 °C and 33% RH.

267

Thiol/disulphide exchange reaction usually occurs between Cys160 of β-LG and free thiols of

268

Cys11 and Cys88 of к-casein (Creamer et al., 2004) forming an aggregate with a minimum

269

Mw of ~ 37 kDa (Figure 5). Similar to other proteins, bovine serum albumin (BSA) was also

270

involved in aggregation. However, its involvement was clearly storage time dependent, while

271

other main factors had no effect. BSA forms disulphide crosslinks with β-LG (Matsudomi et

272

al., 1994) due to a free thiol group at Cys34. In addition to disulphide bridging, WPs in native-

273

WPC participated in non-reducible aggregation (Figure 7), likely due to lactosylation of

274

lysine residues in these proteins. Compared to other WPCs, lactosylation was less prominent

275

in this WPC without a clear association with storage temperature and time, which could have

276

been caused by a very low amount of lactose.

277

Sweet-WPC samples exhibited a similar behavior to that of native-WPC presented by clear

278

formation of protein aggregates covalently linked by disulphide bridging over the overall

279

storage period (Figure 5 and 6), again achieving maximum aggregation at 45 °C and 33%

280

RH. In parallel to the aggregate formation, a rate of β-LG disappearance was the greatest

281

under the same conditions (Table 2), suggesting that β-LG was the main protein involved in

282

aggregation. Conversely, only minor involvement of α-LA in aggregation was observed

283

under all storage conditions. No involvement of BSA was evident in sweet-WPC under all

AC C

EP

TE D

M AN U

SC

RI PT

259

ACCEPTED MANUSCRIPT storage conditions, although comparatively a higher concentration of BSA was present in

285

comparison with other WPCs. In contrast, caseins were involved in disulphide crosslinking,

286

but a clear association with storage conditions and time could not be established. In addition

287

to aggregates linked by disulphide bonds, two distinct bands representing non-disulphide

288

linked aggregates appeared with approximate Mw of 71-72 kDa and ≥250 kDa (Figure 6).

289

Compared to native-WPC, these aggregates were prominent in sweet-WPC, showing a

290

gradual increase with rise in the storage temperature, humidity and time. Sweet-WPC

291

contained 2.9% of lactose, which might have led to lactosylation of lysine segments of β-LG.

292

Among these, Lys8 located at N terminal, Lys141 located at C terminal end and Lys138 located

293

in 3 turn α-helix are easily accessible, therefore, likely the primary sites of the reaction.

294

In acid-WPC, disulphide linked aggregates gradually increased throughout the storage with

295

subsequent disappearance of β-LG A, β-LG B and α-LA (Figure 5 and 6). Rapid involvement

296

of β-LG A was observed in aggregation, compared to β-LG B (Oldfield et al., 1998), mainly

297

due to disulphide bridging via Val118 in β-LA A as compared to Ala118 in β-LG B, located in

298

β-strand H, creating a less favorably packed core (McSweeney, 2013). This in turn increases

299

the flexibility of β-LG A and exposes the free thiol group. Compared to other WPCs,

300

involvement of α-LA in protein aggregation was greater in acid-WPC. Less amount of water

301

might have restricted the availability of Ca2+ to the binding site of α-LA, preventing protein

302

stabilization and thus resulting in prevalence of less stable apo-α-LA. The contribution of

303

caseins in disulphide bonding was apparent only at 45 °C and 33% RH conditions, although a

304

significantly higher amount of caseins was present in acid-WPC than in other WPCs

305

(Nishanthi et al., 2017a). A diffuse region was observed in the region between caseins and

306

BSA (Mw~50 kDa) in the samples stored at 45 °C and two RHs under reducing PAGE

307

conditions (Figure 6), suggesting a possible aggregation between caseins and β-LG and/or α-

308

LA via non-reducible covalent bonds. In addition to reducible aggregates, non-reducible

AC C

EP

TE D

M AN U

SC

RI PT

284

ACCEPTED MANUSCRIPT aggregates likely created by lactosylation with Mw>250 kDa were observed in acid-WPC

310

under all storage condition (Figure 6). The extent of aggregation was clearly temperature and

311

RH dependent (Figure 7) and predominantly involved α-LA, as evidenced by the reaction

312

rates (Table 3). Out of twelve lysine residues in α-LA, maximum of 6 residues (Lys24, Lys112,

313

Lys117, Lys127, Lys133 and Lys141) are lactosylated easily (Le et al., 2012).

314

Similar to other WPCs, salty-WPC contained disulphide linked protein aggregates with

315

Mw>250 kDa (Figure 5 and 6). The aggregation increased with the storage time, achieving

316

the maximum at 45 °C and 33% RH. Both β-LG A and β-LG B participated in disulphide

317

bonding, however, β-LG A completely disappeared after 14 days, suggesting a high reaction

318

rate (Oldfield et al., 1998). Involvement of α-LA in disulphide bonding was observed with

319

extent clearly dependent of the storage temperature and time with no clear association with

320

humidity. Thus, thiol/disulphide interchange reactions between β-LG and α-LA could be the

321

main possible aggregation pathway. In absence of a reactive thiol group, thiol/disulphide

322

interchange reactions occur through the disulphide bonds of Cys6-Cys120 and Cys28-Cys111

323

located on the surface of α-LA, which form new disulphide bonds with Cys66 and Cys160

324

located on the surface of β-LG. Compared to other WPCs, salty-WPC contained a high

325

concentration of lactoferrin (LF) (Mw~87 kDa), which did not participate in covalent

326

aggregation under all storage conditions. On the other hand, BSA was greatly involved

327

concomitantly with the rise in storage temperature, humidity and time. Due to a free thiol

328

group at Cys34 and seventeen disulphide bonds in its primary structure, BSA participates in

329

both disulphide bridging and thiol/disulphide interchange reactions with β-LG (Matsudomi et

330

al., 1994) and α-LA (Matsudomi et al., 1993). Similar to sweet-WPC, salty-WPC contained

331

aggregates linked via covalent non-disulphide bonds in the region of BSA-LF with

332

approximate Mw of 71-72 kDa. Furthermore, the amount of large aggregates (Mw≥250 kDa)

AC C

EP

TE D

M AN U

SC

RI PT

309

ACCEPTED MANUSCRIPT formed via lactosylation increased in parallel with the storage temperature, humidity and

334

time.

335

In four WPCs, type of proteins and interactions involved in protein aggregation vary with the

336

type of powder, storage conditions and time. The main protein involved in disulphide and

337

non-disulphide covalent aggregation was β-LG. Therefore, kinetic parameters of β-LG were

338

estimated to thoroughly understand the differences in behavior of β-LG in these WPCs.

339

3.3. Kinetics of β-LG disappearance

340

Kinetic parameters of β-LG disappearance in different WPCs during storage are presented in

341

Table 2. Disappearance of β-LG follows the common reaction kinetic principles. A

342

substantial reaction occurs if an adequate number of β-LG molecules with sufficient

343

activation energy (Eo) collide at an adequate frequency while maintaining specified molecular

344

orientation to form disulphide or other covalent bonds. Rate of disappearance of β-LG (Rβ-LG)

345

in native-WPC and sweet-WPC was directly related to the storage temperature and humidity,

346

being the greatest at 45 °C and 33% RH. Elevated temperature increases the number of

347

reactive β-LG molecules with exposed free thiol group of Cys121. High RH increases water

348

activity thus favors thermal energy transition and molecular motions, facilitating formation of

349

reactive monomers. In contrast, Rβ-LG of salty-WPC and acid-WPC decreases as RH

350

increases. This observation suggests a greater interference of compositional constituents

351

present in these two WPCs on Rβ-LG. As a result of the low pH in presence of high

352

concentration of LA in acid-WPC (Nishanthi et al., 2017a), dimerization of β-LG and

353

formation of octamers (Townend and Timasheff, 1960) occurs as evidenced by the FTIR

354

spectra and SDSNR PAGE patterns. Due to the large structure of octamer, number of reactive

355

sites exposed to the surface is low, leading to low Rβ-LG. Further, the osmotic environment

356

with the presence of LA and Ca2+ retain water molecules, preventing protein hydration. As a

357

result, in low RH, protein-protein interactions dominate, limiting the mobility of WPs. Ca2+

AC C

EP

TE D

M AN U

SC

RI PT

333

ACCEPTED MANUSCRIPT on the other hand, forms salt bridges between carboxyl groups of amino acids (Zhu and

359

Damodaran, 1994), restricting the collision in required orientation for permanent covalent

360

bonding. Furthermore, anionic lactate actively increases the negative charge at the dimer

361

interface, stabilizing the β-LG dimers (Mercadante et al., 2012) and thereby, prevent the

362

formation of reactive monomers. Salty-WPC, is characteristically high in Na+ (approximately

363

0.5M), which exerts an osmotic pressure of ~0.005 bars, nearly 3 times greater compared to

364

that of sweet-WPC. Under these conditions, Na+ ions are preferentially excluded from the

365

protein surface which affects the binding of water to the protein (Arakawa and Timasheff,

366

1982). In addition, Na+ itself can interact with proteins through salt bridges and charge

367

shielding, subsequently inducing conformational changes and restricting the mobility of the

368

proteins. Influence of compositional variations of acid- and salty-WPCs on β-LG reaction

369

kinetics is further evident by the reaction order and Eo. Reaction orders of acid- and salty-

370

WPCs were approximately closer to 2 under all storage conditions, whereas, that in native-

371

and sweet-WPCs was approximately closer to 1. Furthermore, native-WPC required the

372

greatest Eo compared to other WPCs, implying that β-LG in native-WPC are still in their

373

native form, due to absence of severe thermal treatments during its manufacturing process.

374

Moreover, native-WPC consisted of 84% proteins on dry basis, implying the lack of

375

influence of other whey components on the β-LG structure. Compared to native- and sweet-

376

WPC, Eo was low in acid- and salty-WPCs. Compositional variations and heat treatments

377

applied in manufacturing processes of acid- and salty-WPCs partially denature the β-LG,

378

lowering the Eo.

379

In addition to β-LG, α-LA was also involved in protein aggregation in acid-WPC (Table 3),

380

with a higher rate of disappearance (Rα-LA) than β-LG. This observation contradicts the

381

assumption that in Ca2+ rich environment of acid-WPC α-LA should demonstrate the least

382

reactivity due to the high structural stability attained with the Ca2+ bound holo-form.

AC C

EP

TE D

M AN U

SC

RI PT

358

ACCEPTED MANUSCRIPT However, low Eo established for α-LA in the current study suggests structural instabilities.

384

Perhaps folding intermediates of α-LA with a low affinity for Ca2+ (McSweeney, 2013) are

385

more common than the native form in acid-WPC. Rate of disappearance of α-LA (Rα-LA)

386

varied as 4 < 25 < 45 °C and 33 < 22 %. At high temperature, unfolding rate of α-LA

387

increased, shifting the equilibrium of α-N ⇌ α-U towards the right side (Oldfield et al.,

388

1998), in which state α-LA becomes more prone to engage in thiol/disulphide interchange

389

reactions with reactive β-LG.

390

4.0 Conclusion

391

Origin, composition and storage conditions clearly affected the properties of WPCs

392

examined. Protein aggregation occurred predominantly through covalent crosslinking under

393

all storage temperatures, while higher temperatures accelerated aggregation. Humidity had a

394

variable effect on the protein aggregation, mainly based on the type of WPC. Disulphide and

395

non-disulphide covalent crosslinking of β-LG was the main aggregation pathway in all

396

WPCs, whereas, involvement of α-LA, BSA and caseins depended on the type of WPC. In

397

parallel to the protein aggregation, disappearance of β-LG followed a first-order reaction in

398

native-WPC and sweet-WPC, while it was closely a second-order reaction in acid-WPC and

399

salty-WPC. Compositional variations associated with origin of acid-WPC and salty-WPC

400

facilitates the β-LG loss during the storage. According to the result of the current study, 4 °C

401

can be considered as the optimal storage temperature for all types of WPCs, while

402

determination of optimal relative humidity requires further understanding about effects of

403

storage on physical properties and functional characteristics of these powders.

404

5.0 Acknowledgement

405

The authors acknowledge the financial support granted by the Victoria University

406

Postgraduate Research Funding Scheme. Further gratitude extended to the Greek Yoghurt

AC C

EP

TE D

M AN U

SC

RI PT

383

ACCEPTED MANUSCRIPT and Cheese Manufacturing Companies which contributed with providing whey stream

408

samples. Special acknowledgement is extended to the Dairy Innovation Australia Limited

409

(DIAL) for providing whey processing and spray drying facilities.

410

6.0 Reference

411

Arakawa, T., Timasheff, S.N., (1982). Preferential interactions of proteins with salts in

412

concentrated solutions. Biochemistry 21(25), 6545-6552.

413

Boye, J., Alli, I., (2000). Thermal denaturation of mixtures of α-lactalbumin and β-

414

lactoglobulin: a differential scanning calorimetric study. Food research international 33(8),

415

673-682.

416

Burgain, J., El Zein, R., Scher, J., Petit, J., Norwood, E.-A., Francius, G., Gaiani, C., (2016).

417

Local modifications of whey protein isolate powder surface during high temperature storage.

418

Journal of Food Engineering 178, 39-46.

419

Creamer, L.K., Bienvenue, A., Nilsson, H., Paulsson, M., van Wanroij, M., Lowe, E.K.,

420

Anema, S.G., Boland, M.J., Jiménez-Flores, R., (2004). Heat-induced redistribution of

421

disulfide bonds in milk proteins. 1. Bovine β-lactoglobulin. Journal of Agricultural and Food

422

Chemistry 52(25), 7660-7668.

423

Israelachvili, J.N., (2011). Intermolecular and surface forces: revised third edition. Academic

424

press.

425

Le, T.T., Deeth, H.C., Bhandari, B., Alewood, P.F., Holland, J.W., (2012). A proteomic

426

approach to detect lactosylation and other chemical changes in stored milk protein

427

concentrate. Food Chemistry 132(1), 655-662.

428

Lefèvre, T., Subirade, M., (1999). Structural and interaction properties of β‐Lactoglobulin as

429

studied by FTIR spectroscopy. International journal of food science & technology 34(5‐6),

430

419-428.

AC C

EP

TE D

M AN U

SC

RI PT

407

ACCEPTED MANUSCRIPT Matsudomi, Oshita, T., Kobayashi, K., (1994). Synergistic interaction between β-

432

lactoglobulin and bovine serum albumin in heat-induced gelation. Journal of Dairy Science

433

77(6), 1487-1493.

434

Matsudomi, N., Oshita, T., Kobayashi, K., Kinsella, J.E., (1993). . alpha.-Lactalbumin

435

enhances the gelation properties of bovine serum albumin. Journal of Agricultural and Food

436

Chemistry 41(7), 1053-1057.

437

McSweeney, P.L.H., Fox P.F, (2013). Advanced Dairy Chemistry. Spinger, London.

438

Mercadante, D., Melton, L.D., Norris, G.E., Loo, T.S., Williams, M.A., Dobson, R.C.,

439

Jameson, G.B., (2012). Bovine β-lactoglobulin is dimeric under imitative physiological

440

conditions: dissociation equilibrium and rate constants over the pH range of 2.5–7.5.

441

Biophysical journal 103(2), 303-312.

442

Ngarize, S., Herman, H., Adams, A., Howell, N., (2004). Comparison of changes in the

443

secondary structure of unheated, heated, and high-pressure-treated β-lactoglobulin and

444

ovalbumin proteins using Fourier transform Raman spectroscopy and self-deconvolution.

445

Journal of Agricultural and Food Chemistry 52(21), 6470-6477.

446

Nishanthi, M., Chandrapala, J., Vasiljevic, T., (2017a). Compositional and structural

447

properties of whey proteins of sweet, acid and salty whey concentrates and their respective

448

spray dried powders. International Dairy Journal.

449

Nishanthi, M., Vasiljevic, T., Chandrapala, J., (2017b). Properties of whey proteins obtained

450

from different whey streams. International Dairy Journal 66, 76-83.

451

Norwood, E.-A., Chevallier, M., Le Floch-Fouéré, C., Schuck, P., Jeantet, R., Croguennec,

452

T., (2016). Heat-induced aggregation properties of whey proteins as affected by storage

453

conditions of whey protein isolate powders. Food and Bioprocess Technology 9(6), 993-

454

1001.

AC C

EP

TE D

M AN U

SC

RI PT

431

ACCEPTED MANUSCRIPT Oldfield, D., Taylor, M., Singh, H., (2005). Effect of preheating and other process parameters

456

on whey protein reactions during skim milk powder manufacture. International Dairy Journal

457

15(5), 501-511.

458

Oldfield, D.J., Singh, H., Taylor, M.W., Pearce, K.N., (1998). Kinetics of denaturation and

459

aggregation of whey proteins in skim milk heated in an ultra-high temperature (UHT) pilot

460

plant. International Dairy Journal 8(4), 311-318.

461

Townend, R., Timasheff, S.N., (1960). Molecular Interactions in β-Lactoglobulin. III. Light

462

Scattering Investigation of the Stoichiometry of the Association between pH 3.7 and 5.22.

463

Journal of the American Chemical Society 82(12), 3168-3174.

464

Zhou, P., Liu, X., Labuza, T.P., (2008). Effects of moisture-induced whey protein

465

aggregation on protein conformation, the state of water molecules, and the microstructure and

466

texture of high-protein-containing matrix. Journal of Agricultural and Food Chemistry

467

56(12), 4534-4540.

468

Zhu, H., Damodaran, S., (1994). Effects of calcium and magnesium ions on aggregation of

469

whey protein isolate and its effect on foaming properties. Journal of Agricultural and Food

470

Chemistry 42(4), 856-862.

AC C

EP

TE D

M AN U

SC

RI PT

455

ACCEPTED MANUSCRIPT Table 1: Composition of four WPCs Sweet-WPC

Acid-WPC

Salty-WPC

Moisture (%)

5.5 ± 0.7ab

4.1 ± 0.2ac

5.7 ± 0.8b

3.9 ± 0.08c

Total protein (%)

84.1 ± 0.6a

61.8 ± 0.1b

67 ± 0.3c

10.9 ± 0.2d

Lactose (%)

0.1 ± 0.02a

2.9 ± 0.03b

0.001 ± 0.0001a

0.1 ± 0.03b

Ca (%)

0.04 ± 0.001a

0.02 ± 0.003b

Na (%)

0.01 ± 0.004a

0.4 ± 0.02a

0.03 ± 0.01b

1.1 ± 0.05b

6.5 ± 0.03a

6.4 ± 0.02b

4.2 ± 0.02c

5.5 ± 0.01d

M AN U

Lactic acid (%)

pH

0.7 ± 0.02c

2.4 ± 0.3c

0.2 ± 0.02c

0.07 ± 0.001b

0.14 ± 0.01c

0.08 ± 0.001c

EP

TE D

Values are means of at least 4 independent measurements (n ≥ 4); the results are presented as means ± standard deviation (SD). Different lowercase superscripts in the same row depict the significant difference between means for each spray dried whey powder.

AC C

472 473 474

RI PT

Native-WPC

SC

471

ACCEPTED MANUSCRIPT Table 2: Reaction rate, reaction order, rate constant (Kn) activation energy (Eo) and pre- and

476

pre-exponential factor (K0) of β-LG disappearance in different WPCs during storage

33

22 Sweet 33

22

Order of reaction

Kn

Eo

Ink0

((g Kg-1)(1-n) day-1)

(kJ mol-1)

((g Kg-1)(1-n) day-1)

4

0.09e

1.4a

7.0 x 10-5 a

25

0.18f

1.0b

1.2 x 10-3 b

45

0.29g

0.7c

1.5 x 10-2 c

4

0.11h

1.3d

1.4 x 10-4 d

25

0.15i

1.2e

4.7 x 10-4 e

45

0.39j

0.6f

3.9 x 10-2 f

4

0.13a

1.2g

2.8 x 10-4 g

25

0.22n

45

0.24o

4

0.16p

25

0.19l

45

0.24q

4 25 45

Salty

4 33

25

22 Acid 33

SEM a-x

0.9h

3.2 x 10-3 h

0.8i

6.9 x 10-3 i

1.1j

5.4 x 10-4 j

0.9k

2.6 x 10-3 k

0.8l

5.3 x 10-3 l

0.14b

1.2m

5.8 x 10-4 m

0.16k

1.1n

6.8 x 10-4 n

0.20l

1.0o

2.2 x 10-3 o

0.07d

1.7p

3.3 x 10-5 p

0.08m

1.6q

5.9 x 10-5 q

0.11h

1.4r

2.1 x 10-4 r

AC C

45

RI PT

Native

Reaction rate for β-LG (µg / day)

96.3a

32.2a

97.6b

32.9b

58.3c

17.3c

41.2d

10.5d

23.1e

2.4e

31.9f

3.4f

13.2g

3.2g

54.5h

10.2h

0.1

0.04

SC

22

Temperature (°C)

M AN U

RH (%)

EP

WPC

TE D

475

4

0.12a

1.4s

1.3 x 10-4 s

25

0.13a

1.4t

1.9 x 10-4 p

45

0.14b

1.3u

2.8 x 10-4 t

4

0.06c

1.9v

1.7 x 10-6 u

25

0.06d

1.9w

4.5 x 10-6 v

45

0.10e

1.6x

3.8 x 10-5 w

0.001

0.0

1.4 x 10-3

Different lowercase superscripts in the same column depicts the significant differences between means of kinetic parameters. Results are expressed as means of two independent replications. SEM = Standard error of the mean.

477

ACCEPTED MANUSCRIPT 478

Table 3: Reaction rate, reaction order, rate constant (Kn) activation energy (E0) and pre-

479

exponential factor (K0) of α-LA disappearance in acid-WPCs during storage Reaction rate for α-LA (µg / day)

Order of reaction

Kn

E0

Ink0

((g Kg-1)(1-n) day-1)

(kJ mol-1)

((g Kg-1)(1-n) day-1)

4

0.15a

1.6a

9.4 x 10-5 a

25

0.19b

1.5b

1.9 x 10-4 b

45

0.20c

1.5c

3.9 x 10-4 c

4

0.13d

1.7d

6.4 x 10-5 d

25

0.17e

1.7e

8.7 x 10-5 e

45

0.18f

1.6f

1.7 x 10-4 f

0.001

0.0

7.9 x 10-4

Acid 33

SEM a-f

RI PT

22

Temperature (0C)

25.3 ± 0

1.7 ± 0

17.4 ± 0.1

2.2 ± 0.04

SC

RH (%)

M AN U

Whey

Different lowercase superscripts in the same column depicts the significant differences between means of kinetic parameters. Results are expressed as means of two independent replications. SEM = Standard error of the mean. 480 481

AC C

EP

TE D

482

ACCEPTED MANUSCRIPT List of Figures

484

Figure 1: Vector normalized, smoothened and deconvoluted FTIR spectra of the Amide I

485

region (1600 – 1700 cm-1) for NWP over storage period of 0 (solid- black), 14 (long dashed-

486

black), 30 (short dashed- black), 60 (solid- grey) and 90 day (dashed- grey) under different

487

storage temperatures (4, 25 and 45 0C) and RH (22 and 33 %)

488

Figure 2: Vector normalized, smoothened and deconvoluted FTIR spectra of the Amide I

489

region (1600 – 1700 cm-1) for SWP over storage period of 0 (solid- black), 14 (long dashed-

490

black), 30 (short dashed- black), 60 (solid- grey) and 90 day (dashed- grey) under different

491

storage temperatures (4, 25 and 45 0C) and RH (22 and 33 %)

492

Figure 3: Vector normalized, smoothened and deconvoluted FTIR spectra of the Amide I

493

region (1600 – 1700 cm-1) for AWP over storage period of 0 (solid- black), 14 (long dashed-

494

black), 30 (short dashed- black), 60 (solid- grey) and 90 day (dashed- grey) under different

495

storage temperatures (4, 25 and 45 0C) and RH (22 and 33 %)

496

Figure 4: Vector normalized, smoothened and deconvoluted FTIR spectra of the Amide I

497

region (1600 – 1700 cm-1) for StWP over storage period of 0 (solid- black), 14 (long dashed-

498

black), 30 (short dashed- black), 60 (solid- grey) and 90 day (dashed- grey) under different

499

storage temperatures (4, 25 and 45 0C) and RH (22 and 33 %)

500

Figure 5: SDS Non-reduced PAGE patterns of whey protein powders at 0, 30 and 90 days of

501

storage. L1, L2, L3, L4 represent NWP, SWP, AWP and StWP under 22% RH, while L5, L6,

502

L7, L8 represent NWP, SWP, AWP and StWP under 33% RH, respectively. X1 represent

503

covalently linked aggregates with Mw ≥250 kDa. X2 represents medium size protein

504

aggregates with Mw~35 – 67 kDa. X3 represent β-LG A.

505

Figure 6: SDS Reduced PAGE patterns of whey protein powders at 0, 30 and 90 days of

506

storage. L1, L2, L3, L4 represent NWP, SWP, AWP and StWP under 22% RH, while L5, L6,

AC C

EP

TE D

M AN U

SC

RI PT

483

ACCEPTED MANUSCRIPT L7, L8 represent NWP, SWP, AWP and StWP under 33% RH, respectively. NDCA represent

508

non-disulphide covalently linked aggregates.

509

Figure 7: Changes in absolute quantity (µg) of non-reducible aggregates during storage of

510

NWP (solid line), SWP (dotted line), AWP (dashed line) and StWP (dashed-dotted line) at

511

storage temperatures of 4, 25 and 450C and relative humidity of 22 and 33%, as obtained

512

from image analysis of reducing SDS-PAGE.

AC C

EP

TE D

M AN U

SC

RI PT

507

514 515

Figure 1:

AC C

513

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

517 518 519 520

Figure 2:

AC C

EP

TE D

M AN U

SC

RI PT

516

ACCEPTED MANUSCRIPT

522 523 524 525

Figure 3:

AC C

EP

TE D

M AN U

SC

RI PT

521

ACCEPTED MANUSCRIPT

527 528 529

Figure 4:

AC C

EP

TE D

M AN U

SC

RI PT

526

530 531 532

Figure 5.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

533 534 535

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 6:

ACCEPTED MANUSCRIPT

538

Figure 7:

AC C

537

EP

TE D

M AN U

SC

RI PT

536

ACCEPTED MANUSCRIPT Highlights Properties of various WPCs were examined under different storage conditions.



β-LG was primarily involved in covalent cross linking in all WPCs.



Participation of α-LA in covalent crosslinking was evident primarily in acid-WPC



High storage temperature accelerated covalent linking between whey proteins



Reaction order of β-LG denaturation in acid and salty-WPCs was approximately 2

AC C

EP

TE D

M AN U

SC

while it was approximately 1 in native and sweet-WPCs

RI PT