Conformational Transition of Polyelectrolyte As ... - ACS Publications

Oct 28, 2016 - Glyn O. Phillips Hydrocolloid Research Centre, School of Food and Biological Engineering and. ‡ ... car. Moreover, it was found that ...
0 downloads 13 Views 1MB Size
Subscriber access provided by RYERSON UNIVERSITY

Article

Conformational transition of polyelectrolyte as influenced by electrostatic complexation with protein Yiping Cao, Shugang Li, Yapeng Fang, Katsuyoshi Nishinari, and Glyn O. Phillips Biomacromolecules, Just Accepted Manuscript • DOI: 10.1021/acs.biomac.6b01335 • Publication Date (Web): 28 Oct 2016 Downloaded from http://pubs.acs.org on November 3, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Biomacromolecules is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

1

Conformational transition of polyelectrolyte as

2

influenced by electrostatic complexation with protein

3

Yiping Cao,† Shugang Li, † Yapeng Fang,*, †, ‡ Katsuyoshi Nishinari,†, ‡ and Glyn O. Phillips†

4

5 6



7

University of Technology, Wuhan 430068, China;

8



9

Technology, Wuhan 430068, China.

10

Glyn O. Phillips Hydrocolloid Research Centre, School of Food and Biological Engineering, Hubei

Hubei Collaborative Innovation Centre for Industrial Fermentation, Hubei University of

*Corresponding author: Yapeng Fang, Email: [email protected]; Tel: +86-(0)-27-59750470.

11

12

13

14

KEYWORDS: conformational transition, electrostatic complexation, theoretical modeling,

15

β-lactoglobulin, κ-carrageenan, gelling properties

16

17

1

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 24

18

ABSTRACT: Conformation and conformational transitions are essential for the biological and

19

technological functions of natural polyelectrolytes, e.g., DNA. This study aims to clarify how the

20

conformational transition of natural polyelectrolyte is affected and tuned by electrostatic

21

complexation

22

protein/polyelectrolyte system, β-lactoglobulin (β-lg) and κ-carrageenan (κ-car), was used for the

23

investigation. The effect was found to be determined by the molecular state of β-lg/κ-car electrostatic

24

complexation and the molecular weight of protein. β-lg/κ-car complexation in soluble state had a

25

subtle effect on the coil-to-helix transition of κ-car, while that in insoluble state greatly suppressed it.

26

Based on the McGhee-Hippel theory, a quantitative model was successfully developed to describe

27

the effect of protein/polyelectrolyte electrostatic complexation on the conformational transition of

28

polyelectrolyte. The model can also provide additional information on the change of tertiary structure

29

of β-lg upon electrostatic complexation with κ-car. Moreover, it was found that β-lg or its

30

hydrolysates with a molecular weight larger than 2000 Da hindered the conformational transition of

31

κ-car, while those with a molecular weight lower than 1000 Da promoted it. The observations offer a

32

promising approach to controlling the conformational transition and related properties of

33

polyelectrolytes for technological applications.

34

INTRODUCTION

with

protein

as

encountered

in

many

biological

processes.

A

model

35

Natural polyelectrolytes exhibit diverse conformations in solutions, including random coil, single

36

helix, multiple helix, aggregate and sphere-like structure, etc.1-3 These conformations are involved in

37

many biological processes and in human diseases. For instance, DNA with mutated conformation

38

(e.g., hairpin, triplex, cruciform, and left-handed form) is associated with diseases such as myotonic

39

dystrophy, Huntington’s chorea, fragile X syndrome, and Friedreich’s ataxia.1 In cell biology, 2

ACS Paragon Plus Environment

Page 3 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

40

conformational equilibria of ubiquitin can determine its specific binding with protein or DNA, thus

41

regulating protein degradation, cell signaling/trafficking and DNA damage response.4 Conformation

42

and conformational transitions of polyelectrolytes influence also their bioactivities. Triple helical

43

schizophyllan shows strong inhibition against tumor cells, whereas its single coil counterpart has

44

almost no bioactivity. This behavior is attributed to an increased interaction between the extended

45

chain conformation and the immune system.5 Sulphation or phosphatization is known to greatly

46

improve the inhibitivity of branched α-(1→4)-D-glucan against H-22 tumor cells. The effect is

47

explained by a conformational transition from compact sphere to random coil upon the chemical

48

modification, which confers the polyelectrolyte more flexibility to bind with relevant receptors on

49

the tumor cell surface.6 Conformational transitions of polyelectrolytes are applied in industry for

50

applications such as thickening, gelling, and stabilization.7

51

Electrostatic complexation of polyelectrolyte with protein is also important in biological processes

52

including DNA replication, transcription and repair.8,

53

non-specifically to DNA via electrostatic interaction, and move subsequently along the DNA strand

54

until specific binding sites are identified.10 Protein/polyelectrolyte electrostatic complexation occurs

55

naturally in many living organisms, triggering a variety of biological functions, such as the adhesive

56

property of sandcastle worms provided by the complexation of highly polar proteins11 and the

57

mechanical properties of mammalian cartilages provided by the ternary complexes of aggrecan,

58

hyaluronan and cartilage link proteins12. Protein/polyelectrolyte complexation has found numerous

59

applications

60

separation/purification,13 enzyme stabilization/immobilization,11 drug delivery, and gene therapy14.

61

In the food industry, protein/polyelectrolyte electrostatic complexation was used to design

in

the

biotechnological

and

9

Protein is considered first to bind

pharmaceutical

industries,

including

protein

3

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 24

62

low-calorie and low-starch foods,15 encapsulate flavors and probiotics,16, 17 and stabilize emulsions

63

and foams18.

64

Coupling of protein/polyelectrolyte electrostatic complexation with conformational transition was

65

frequently encountered in biological processes and technological applications. Our previous studies

66

have

67

conformational transition via two interconnected mechanisms, namely, specific ionic binding and

68

chain stiffening.19 It is necessary now to clarify how the conformational transition of polyelectrolyte

69

is controlled by electrostatic complexation with protein. We demonstrate this behavior by reference

70

to a model protein/polyelectrolyte system, β-lactoglobulin (β-lg) and κ-carrageenan (κ-car).

shown

that

protein/polyelectrolyte

complexation

is

influenced

by

polyelectrolyte

71

β-lg is the dominant globular protein found in whey protein.20 It has 162 amino acid residues, and

72

a molecular mass of 18.4 kDa.20, 21 κ-car is a naturally anionic polysaccharide extracted from red

73

seaweed species and is widely used in food, cosmetic, and pharmaceutical products as gelling agents,

74

thickeners, stabilizers, and excipients.7 It is made up of alternating α-(1→3)-D-galactose-4-sulfate

75

and β-(1→4)-3,6-anhydro-D-galactose repeating units.7, 22 The temperature-induced conformational

76

transition of κ-car has been a subject of debate for many years. It is traditionally accepted that κ-car,

77

in the presence of ions (e.g. K+), undergoes a conformational transition from random coil to double

78

helix and further to helical aggregates with lowering temperature.19,

79

elegantly showed that the κ-car chains first undergo a simple coil-to-helix transition, and then the

80

single helices interwine or supercoil into supermolecular strands with a hierarchical chirality

81

amplification, using high-resolution atomic force microscopy.24-26 This represents a significant

82

advancement of the debates on the conformational transition of carrageenans.

83

EXPERIMENTAL SECTION

23

Recently, Schefer et al.

4

ACS Paragon Plus Environment

Page 5 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

84

Biomacromolecules

Materials

85

β-lg was kindly supplied by Davisco, which according to the supplier has a minimum purity of >

86

95%. β-lg has a weight average molar mass of Mw = 19.1 kDa, a polydispersity index of Mw/Mn =1.1,

87

and a radius of gyration of Rg = 6.5 nm, as determined by gel permeation chromatography-multiple

88

angle laser light scattering (GPC-MALLS, Wyatt Technology Corporation, USA) at 25 oC in 0.1 M

89

phosphate buffer (pH 7.0) using a Sepax SRT SEC-150 column (Sepax Technologies, USA). The

90

isoelectric point (IEP) of the β-lg sample measured by Nano-ZS ZetaSizer (Malvern Instruments, UK)

91

is 5.2 (Figure S1), which agrees with literature values.20, 21, 27

92

κ-car was a gift from FMC biopolymer (Gelcarin GP-911NF). It was converted into the sodium

93

salt using ion exchange resin (Amberlite IR-120, Sigma), and then lyophilized. The purified sample

94

contains 6.32% Na, 0.067% K, 0.0027% Mg, and 0.0083% Ca as determined by atomic absorption

95

spectrometry. The purification was to avoid the complicated effects of mixed counterions, allowing a

96

precise molecular characterization of κ-car. The molecular parameters measured by GPC-MALLS at

97

25 oC in 0.1 M NaI using a Shodex OHpak SB-805 separation column (GE Healthcare Co., USA) are:

98

Mw = 467 kDa; Mw/Mn = 1.2; Rg = 85.0 nm, which represent the doubly-associated helical

99

conformation of κ-car without further aggregation.19, 28 The zeta potential of κ-car, measured in the

100

pH range of 3-8, is characteristic of strong polyelectrolytes with nearly a constant value of ca. -60

101

mV (Figure S1), which agrees with most literature reports.29

102

Preparation of β-lg hydrolysates

103

Hydrolysates with different molecular weight were obtained by hydrolyzing β-lg to different

104

extents. 5.0 % β-lg was heated at 85 oC for 5 min, followed by hydrolysis using the enzyme papain

105

(3000 U, Biosharp) at the optimum conditions (45 oC, pH 4.5) for 4 h. The hydrolysate was separated

106

using ultrafiltration tube (MWCO 3000, Biosharp), and the filtrate was loaded into dialysis tube 5

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 24

107

(MWCO 1000, Biosharp) for further separation. After dialysis against Millipore water at 4 oC for 96

108

h, the hydrolysates inside and outside the dialysis tube were collected by freeze-drying and

109

designated as S1 and S2, respectively. A total hydrolysate of β-lg (S3) was obtained by heat

110

treatment at 120 oC in 6.0 M HCl for 12 h, followed by removal of HCl using vacuum evaporation at

111

60 oC. Although S1, S2 and S3 were mixtures of peptide fragments of different length and chemical

112

composition, as a whole, their average amino acid compositions were nearly the same as the mother

113

protein β-lg without hydrolysis. This was confirmed by analysis using an automatic amino acid

114

analyzer (Hitachi, L-8900). Weight average molar mass measured by GPC-MALLS at 25 oC in 0.1

115

M phosphate buffer using a Sepax SRT SEC-150 column is: β-lg (19100 Da) > S1 (2053 Da) > S2

116

( 921 Da). The weight average molar mass of S3 was calculated to be 127 Da, according to the amino

117

acid composition of β-lg.20, 21

118

Solution preparation

119

Stock solutions of 0.06-1.80 wt% β-lg and 0.90 wt% κ-car were prepared by dissolving

120

appropriate amounts of the samples into 50 mM KCl. κ-car solutions were heated at 85 oC for 1 hour

121

under magnetic stirring. β-lg solutions were dissolved at ambient temperature overnight on a roller

122

mixer. β-lg/κ-car mixtures at a fixed κ-car concentration (0.15%) and various β-lg/κ-car mixing

123

ratios (w/w, 0 < r IEP.31 It is because that both κ-car and β-lg are overall

188

negatively charged at this pH, with a zeta potential of -60 mV and -25 mV, respectively (Figure S1).

189

The absence of electrostatic complexation at pH 5.6 is supported by the turbidity measurements in

190

Figure 1c where τ remains close to zero at all mixing ratios at 45 oC.

191

In contrast, at pH 4.7 < IEP, the exothermic peak associated with the coil-to-helix transition of

192

κ-car was significantly reduced with increasing β-lg/κ-car mixing ratio (Figure 1d). The exothermic

193

peak tends to disappear completely at r > 8.00. This indicates that the conformational transition of

194

κ-car is greatly suppressed when mixed with β-lg. This can be attributed to the electrostatic

195

complexation between β-lg and κ-car, which impairs the helix formation of κ-car. Figures 1e plots

196

the enthalpy change (∆H) of the conformational transition of κ-car as a function of r at pH 4.7. ∆H

197

shows a two-step decrease with increasing r, and a turning point is located at r = 0.67. Beyond the

198

turning point (r > 0.67), the decrease in ∆H is apparently accelerated. It means that the

199

conformational transition of κ-car is more seriously impaired beyond the turning point.

200

Interestingly, the turning point in ∆H-r curve coincides with that obtained in τ-r curve at 45 oC

201

(Figure 1f). Turbidity is relatively low at r < 0.67, whereas it increases abruptly at r > 0.67. Similar

202

coincidence was found at other pHs, and the turning point is at r = 0.33, 1.45, 2.70 and 4.00, 10

ACS Paragon Plus Environment

Page 11 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

203

respectively, for pH = 4.4, 4.9, 5.0 and 5.1(see Figure S2). Turbidity was seen as a sensitive measure

204

of electrostatic complexation of protein/polyelectrolyte.11, 31, 32 The abrupt increase in τ at the turning

205

points is attributed to the transition from soluble to insoluble electrostatic complexes.31, 32 When β-lg

206

molecules crosslink different κ-car chains, large aggregates that are insoluble are formed, and thus

207

the turbidity increases.31, 32 Note that further decrease in pH below 4.4 caused extensive precipitation,

208

preventing the measurement of the conformational transition of κ-car. The evolution of electrostatic

209

complexes as a function of pH and mixing ratio has been detailedly studied in our previous works.29,

210

30

211

The molecular state of β-lg/κ-car complexation seems to control the conformational transition of

κ-car (Figure 2).

(a) cooling

(b) cooling

212 213

Figure 2. Schematic representation of the effect of β-lg/κ-car electrostatic complexation on the

214

conformational transition of κ-car upon cooling: (a) β-lg/κ-car complexation in soluble state; (b)

215

β-lg/κ-car complexation in insoluble state. κ-car and β-lg molecules are illustrated in black and red,

216

respectively. During the conformational transition, coiled κ-car chains transform into ordered helical

217

conformations (bold lines) which can be single helix, double helix, or even supermolecular strand.

11

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 24

218

During initial stage of electrostatic complexation (Figure 2a), a small number of β-lg are bound to

219

κ-car, and the complexes remain soluble. β-lg molecules pose limited steric effect and κ-car

220

molecules have sufficient freedom to form helices through pairing. Therefore, the conformational

221

transition of κ-car is influenced to a limited extent by the electrostatic complexation with β-lg. As

222

electrostatic complexation further proceeds (Figure 2b), more and more β-lg are bound to κ-car, and

223

cross-linking between κ-car occurs. This causes the formation of insoluble β-lg/κ-car complexes.

224

κ-car molecules are sterically blocked by β-lg and have limited freedom to restructure to form

225

helices. The conformational transition of κ-car in this instance is greatly suppressed by electrostatic

226

complexation with β-lg. It should be pointed out that although the exact conformations of κ-car are

227

to some extent controversial (i.e., single helix, double helix, or supermolecular helical strand),24-26

228

the proposed mechanism, as schematically shown in Figure 2, applies to different situations.

229

Theoretical modeling

230

As already discussed, the conformational transition of κ-car is greatly suppressed by electrostatic

231

complexation with β-lg, and the origin of this effect is κ-car being physically occupied with β-lg.

232

Assuming that the relative extent of conformational transition of κ-car is proportional to the number

233

of repeating units that are unoccupied with β-lg, a quantitative model can be developed to describe

234

the effect of protein/polyelectrolyte electrostatic complexation on the conformational transition of

235

polyelectrolyte (see Supporting Information), based on the DNA-protein binding theory proposed by

236

McGhee- Hippel:33

237

(1 − φ (r ) )φ (r ) + 1 − φ (r )  

m



m −1

 C p M wP N (1 − φ (r ))  − Kmφ (r ) m r − CP =0 Pr m  M w 

(2)

238

where r is the protein/polyelectrolyte weight ratio; φ(r) is the relative extent of conformational

239

transition of polyelectrolyte; m represents the number of repeating units consecutively covered by 12

ACS Paragon Plus Environment

Page 13 of 24

240

one protein molecule and is the reciprocal of the binding stoichiometry n; K is the binding constant;

241

Cp is the molar concentration of polyelectrolyte; M wP and M wPr are the molecular weight of

242

polyelectrolyte and protein, respectively; N is the total number of repeating units of polyelectrolyte.

1.2

1.03

(a)

(b)

1.00

φ =0.93

0.9

0.97 0.6

φ

φ

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

0.94

φ =0.93

0.3 0.91 0

0.88 0

243

2

4

6

r =β -lg/κ -car

8

10

0

0.9

1.8

2.7

r =β -lg/κ -car

3.6

4.5

244

Figure 3. (a) Relative extent of conformational transition of κ-car (φ ) as a function of β-lg/κ-car

245

mixing ratio r at different pHs: pH=5.6 (■); pH=5.1 (●); pH=5.0 (◆); pH=4.9 (▲); pH=4.7 (×);

246

pH=4.4 (+); (b) Enlarged part of Figure a showing two-step decrease in φ. The solid lines represent

247

the curve fittings to Equation 2. The dot-dash lines at φ=0.93 mark the turning points of φ -r curves.

248

φ(r) could be experimentally measured as the ratio of the enthalpy change of conformational

249

transition of protein/polyelectrolyte mixture (∆H(r)) to that of pure polyelectrolyte (∆H(r=0)). Figure

250

3 plots φ as a function of β-lg/κ-car mixing ratio at different pHs. Except at pH 5.6, which is above

251

the IEP of β-lg, φ at pH = 5.1, 5.0, 4.9, 4.7 and 4.4 all exhibit a two-step decrease with r. The

252

decrease is more pronounced at lower pH. This is due to an increased electrostatic complexation of

253

β-lg/κ-car at lower pH, which exerts a stronger effect on suppressing the conformational transition of

254

κ-car. Intriguingly, the turning points of the two-step decrease at different pHs occur at the same

255

value of φ = 0.93 (Figure 3b). This probably indicates a common critical binding threshold beyond

256

which insoluble β-lg/κ-car complexes start to be formed and the conformational transition of κ-car is

13

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 24

257

seriously inhibited. This threshold corresponds to 7% repeating units of κ-car being bound/occupied

258

with β-lg.

259

Taking N = 572, Cp= 6.42×10-6 mol/L, M wPr = 19100 Da and M wP = 233500 Da for the present

260

system, φ measured by DSC can be well fitted to Equation (2) (Solid lines in Figure 3a). The fittings

261

yielded the thermodynamic binding parameters m and K at different pHs, as listed in Table 1. The

262

parameters are in good agreement with those obtained from ITC measurements at 45 oC (see Figure

263

S3). The ITC results represent the binding of β-lg to κ-car in coil state and just prior to the

264

coil-to-helix transition, and therefore to large extent can approximate their dynamic binding during

265

the subsequent conformational transition. The agreement validates the theoretical model as

266

developed in Equation (2).

267

Table 1. Thermodynamic parameters of the binding of β-lg to κ-car obtained from DSC and ITC. Method

parameter

pH=4.4

pH=4.7

pH=4.9

pH=5.0

pH=5.1

m

25.2

17.1

13.9

13.2

12.2

K

1.7×108

2.3×106

3.1×103

3.5×102

5.1×101

m

25.5

16.6

14.2

12.8

12.3

K

1.9×108

2.1×106

3.0×103

3.9×102

6.2×101

DSC

ITC 268

m=1/n, n is the binding stoichiometry of β-lg to κ-car.

269

With lowering pH from 5.1 to 4.4, K increases from 5.1×101 to 1.7×108 M-1. The increase in K is

270

apparently attributed to an increase in electrostatic attraction. Note that the binding constants with

271

magnitude as high as 108 at lower pHs might represent strong out-of-equilibrium bindings, where

272

precipitation of β-lg/κ-car complexes occurred. m is also increased by lowering pH, but its change

273

around IEP (i.e., pH 5.1 and 5.0) is insignificant. Considering a repeating unit length of 1.03 nm for 14

ACS Paragon Plus Environment

Page 15 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

274

κ-car,34 m = 12.2 at pH 5.1 corresponds to a consecutive binding length of 12.6 nm, which is very

275

close to the diameter of β-lg (2Rg = 13 nm) as measured by GPC-MALLS. It implies that the weak

276

binding between β-lg/κ-car close to IEP is determined by the size of β-lg (most likely in dimeric

277

form).35 In contrast, m = 25.2 at pH 4.4, which corresponds to a consecutive binding length of 30.0

278

nm. This length is more than double the diameter of β-lg. It could be due to an increase in protein

279

surface charge, which increases its effective Debye length. Another possibility is that the strong

280

binding between β-lg/κ-car at lower pH induces a geometric elongation of β-lg.

281

DSC measurements (see Figure S4) show that at pH values (i.e., pH 5.1 and 5.6) close to IEP,

282

there is almost no difference in protein denaturation temperature between pure β-lg and β-lg/κ-car

283

mixture. However, at lower pHs (e.g., pH 4.4), the denaturation temperature of pure β-lg is

284

considerably higher than that in the β-lg/κ-car mixture. The reduction in denaturation temperature

285

indicates a less stable conformation or tertiary structure of β-lg upon electrostatic complexation with

286

κ-car. Furthermore, structural component analysis by attenuated total reflectance Fourier transform

287

infrared spectroscopy (ATR-FTIR) shows that the β-sheet and α-helix content of β-lg are both

288

decreased upon complexation with κ-car at lower pHs (see Figure S5). It was also found that protein

289

structures were altered upon complexation with polyelectrolyte in α-gliadin/gum arabic or

290

β-lg/acacia gum systems.32, 36 The structural changes revealed by DSC and ATR-FTIR may be

291

responsible for the geometric elongation of β-lg, as predicted by the theoretic model in Equation (2).

292

The situation is similar to the binding of proteins onto negatively charged surfaces (e.g., mica, silica)

293

at pHs < IEP or hydrophobic surfaces. The strong electrostatic interaction or hydrophobic interaction

294

upon adsorption onto the surfaces caused somewhat denaturation and spreading/elongation of the

295

proteins along the surfaces. 37-39

15

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 24

296

On the other hand, β-lg was reported to exhibit a number of pH-induced, local and global

297

structural transitions between pH 1.0 and pH 13, including the dimer-to-monomer transition (pH

298

2.5-4.0), the Q-to-N transition (pH 4.5-6.0) and the Tanford transition (> pH 6.5), etc.35 In the pH

299

range investigated (pH 4.4-5.6), β-lg exists predominately as dimers, and undergoes a

300

conformational change from the compact native form (N) at pH 6.0 to the more extended acidic form

301

(Q) at pH 4.5. The hydrodynamic diameter of β-lg molecule was reported to increase from 5.16 nm

302

to 5.44 nm, accordingly.40 Such a pH-induced size expansion (only by 5%) could contribute to, but

303

was not fully responsible for the observed elongation of β-lg upon electrostatic complexation with

304

κ-car (Table 1). Moreover, Taulier and the coworker observed a slight change in β-lg tertiary

305

structure (i.e., increase in volume and compressibility) during the Q-to-N transition, which however

306

was not accompanied by any significant alteration in secondary structure.35 This slight change in

307

tertiary structure seems not to have conspicuous influence on the electrostatic complexation of β-lg

308

with κ-car, as the complexation behaviors at different pHs can be described by a common theoretical

309

model.

310

Effect of protein molecular weight

311

Since protein/polyelectrolyte electrostatic complexation suppresses the conformational transition

312

of polyelectrolyte via the physical hindrance of protein, it is desirable to investigate in details the

313

effect of protein molecular weight or size. For this purpose, β-lg was hydrolyzed into different

314

extents by enzymatic or acidic treatment, and three hydrolysates of different molecular weights were

315

fractionated using ultrafiltration and dialysis: S1 (2053 Da) > S2 (921 Da) > S3 (127 Da). The three

316

fractions have almost the same average amino acid compositions as the mother protein β-lg.

16

ACS Paragon Plus Environment

Page 17 of 24

44

(a)

1.2

Control lysine S3 S2 S1 β-lg

43

42

41

0.6

0.3

40

0 0

317

(b)

0.9

φ

To(oC) h

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

2

4

r

6

8

10

0

2

4

r

6

8

10

318

Figure 4. Onset temperature (To) (a) and relative extent (φ ) (b) of the conformational transition of

319

κ-car, as affected by electrostatic complexation with β-lg or its hydrolysates. κ-car 0.15 %; pH = 4.7.

320

Figure 4 shows how differently the β-lg hydrolysates affect the conformational transition of κ-car

321

upon electrostatic complexation at pH 4.7. The effect of lysine was also compared, as it is one of the

322

most typical and abundant amino acids in β-lg. When the hydrolysate molecular weight is > 2000 Da

323

(e.g., β-lg and S1), the onset temperature (To) and relative extent (φ) of conformational transition of

324

κ-car are markedly reduced with increasing r, suggesting a suppressing effect of the hydrolysates.

325

When the hydrolysate molecular weight is around 1000 Da (e.g., S2), To and φ are nearly constant

326

with increasing r. It indicates that the electrostatic complexation in this range of hydrolysate

327

molecular weight has no effect on the conformational transition of κ-car. Significantly, when the

328

hydrolysate molecular weight is < 1000 Da (e.g., S3 and the amino acid lysine), To increases with

329

increasing r, while φ remains constant. The increase in To suggests that the hydrolysates stabilize the

330

helical conformation of κ-car, which can be formed at higher temperatures.

331

As discussed above, the suppressing effect of β-lg or its hydrolysates with a molecular weight

332

larger than 2000 Da is due to their steric effect, which physically hinders the conformational

333

transition of κ-car upon electrostatic complexation. K and m for S1 binding to κ-car at pH 4.7 were

334

calculated to be 4.1×102 M-1 and 6.8, respectively, based on the curve-fitting using Equation (2). The 17

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 24

335

values are significantly smaller than those of 2.3×106 M-1 and 17.1 for β-lg binding to κ-car at pH 4.7,

336

indicating a decreased binding affinity with decreasing hydrolysate molecular weight. m = 6.8

337

corresponds to a consecutive binding length of 6.8 x 1.03 nm = 7.0 nm, which agrees well with the

338

size of S1 of 7.2 nm (containing 19 amino acids, each being approximately 0.38 nm in length).41

339

The steric effect disappears completely when the molecular weight of β-lg hydrolysates is

340

decreased to around 1000 Da. With further decreasing molecular weight below 1000 Da, the ionic

341

strength effect of the short hydrolysates becomes pronounced, due to the increasing number of amino

342

and carboxyl groups. This facilitates the coil-to-helix transition of κ-car and stabilizes the formed

343

helices during cooling. The promoting effect of β-lg hydrolysates with a molecular weight lower than

344

1000 Da is similar to that of adding electrolytes such as NaCl.22 The addition of NaCl was reported

345

to increase also the onset temperature of the conformational ordering of κ-car, while it did not

346

change noticeably the transition enthalpy at high ionic concentration. The promoting effect was

347

attributed to the screening of intra- or inter-molecular electrostatic repulsion and the alteration of

348

hydration properties of κ-car chains, which favors the formation of ordered helices.42

349

Application to tuning gelation and gel viscoelasticity

350

Since the conformational transition of κ-car underlies its gelation, we tried controlling the gelation

351

and gel viscoelasticity of κ-car using electrostatic complexation with β-lg or its hydrolysates. The

352

gelation and gel viscoelasticity of κ-car in mixtures with β-lg at pH 4.7 are shown in Figure 5.

353

During cooling, the storage modulus (G') of 0.15% κ-car exhibits an abrupt increase around 40.6 oC

354

(Figure 5a), indicating the occurrence of gelation.43 The onset temperature of gelation seems not to

355

be influenced by mixing with β-lg. However, the gel elasticity as characterized by G' is greatly

356

reduced with increasing r. The gelation is completely prevented at r = 10.0. Figure 5b displays the

357

frequency dependence of gel viscoelasticity of β-lg/κ-car mixture at 10 oC. For pure κ-car, both 18

ACS Paragon Plus Environment

Page 19 of 24

358

moduli are independent of frequency with G' > G'', indicative of a typical elastic gel.44 With

359

increasing r, G' and G'' become more and more frequency-dependent and cross over at high

360

frequencies. This is an indication of weak viscoelastic gels or viscous solutions.44 The results

361

demonstrate that by mixing with protein at different ratios electrostatic complexation can be used to

362

tune the gelling behavior of polyelectrolytes, to give viscoelastic properties ranging from elastic gels

363

to viscous solutions. The effect is in line with that on the conformational transition of κ-car, as

364

shown in Figure 3a.

1000 103

1000 103

(a)

r =0.00

10 10

r =2.00 r =6.67

G', G'' (Pa)

G' (Pa)

10 101

0.1 10-1 10

(b)

100 102

100 102

101 10

r =10.0

1010 0.1 10-1

15

20

25

T

365

30

35

40

45

0.01 10-2 10-2 0.01

0.1-1 10

(oC)

10 10

101 10

100 102

f (Hz)

366

Figure 5. (a) Gelation profiles of storage modulus G' versus temperature during cooling for κ-car at

367

different β-lg/κ-car mixing ratios r; (b) Frequency sweep of storage (G', closed symbol) and loss

368

moduli (G'', opened symbol) of β-lg/κ-car mixtures at 10 oC. κ-car 0.15 %; pH=4.7.

1000 103

1000 103

(a)

G', G'' (Pa)

101 10

1010

0.1 10-1 10

369 370

(b)

control

β -lg

100 102

100 102

G' (Pa)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

S1

10 101

S2 S3

10 10

lysine

10-1 0.1

15

20

25

30

35

40

45

10-2 0.01 10-2 0.01

T (oC)

10-1 0.1

10 10

10 101

102 100

f (Hz)

Figure 6. (a) Gelation profiles of storage modulus G' versus temperature during cooling for κ-car in 19

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 24

371

mixtures with β-lg and its hydrolysates; (b) Frequency sweep of storage (G', closed symbol) and loss

372

moduli (G'', opened symbol) of the mixtures at 10 oC. κ-car 0.15 %; r = 10.0; pH = 4.7.

373

Figure 6 demonstrates tuning the gelation and gel viscoelasticity of κ-car by electrostatic

374

complexation with β-lg hydrolysates of different molecular weights. At r =10.0 and pH = 4.7, the

375

presence of β-lg or S1 (> 2000 Da) significantly reduces the gelling ability and viscoelasticity of

376

κ-car (Figure 6a). The corresponding frequency sweep exhibits weak gel type or liquid behavior

377

(Figure 6b). The presence of S2 ( ~ 1000 Da) has almost no effect on the gelation and gel

378

viscoelasticity of κ-car. The gelation profile and frequency sweep of S2/κ-car overlap with those of

379

pure κ-car. In contrast, the low molecular weight hydrolysates, S3 and lysine (< 1000 Da), enhance

380

considerably the onset gelation temperature of κ-car, although the gel elasticity is barely affected

381

(Figure 6b). The results reveal different modulatory effects of protein and its hydrolysates on the

382

gelation and gel properties of polyelectrolyte. The effect of protein hydrolysates is in concert with

383

that on the conformational transition of κ-car, as shown in Figure 4.

384

CONCLUSIONS

385

The effect of protein/polyelectrolyte electrostatic complexation on the conformation transition of

386

polyelectrolyte was investigated both experimentally and theoretically by referring to β-lg/κ-car. The

387

effect was found to be related to the molecular state of electrostatic complexation: I) electrostatic

388

complexation in soluble state had subtle effect on the coil-to-helix transition of κ-car, which is due to

389

the relatively high freedom of κ-car at the initial stage of protein binding; II) electrostatic

390

complexation in insoluble state however greatly suppressed the conformational transition, which is

391

due to the high physical hindrance imposed by β-lg upon extensive protein binding. Based on the

392

McGhee-Hippel theory, a quantitative model was successfully developed to describe the effect of 20

ACS Paragon Plus Environment

Page 21 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

393

protein/polyelectrolyte electrostatic complexation on the conformation transition of polyelectrolyte.

394

Moreover, it was found that the effect was closely associated with the molecular weight of β-lg or its

395

hydrolysates. Larger hydrolysates (> 2000 Da) had an inhibitory effect on the conformational

396

transition of κ-car, whereas shorter hydrolysates (< 1000 Da) tended to promote it. The study

397

demonstrated the possibility of controlling the gelation and gel properties of polyelectrolytes via

398

electrostatic complexation with protein or its hydrolysates.

399

ASSOCIATED CONTENT

400

Supporting Information

401

Methods for measuring protein denaturation temperature (DSC) and secondary structure

402

(ATR-FTIR); Zeta potential of β-lg and κ-car (Figure S1); Enthalpy change and turbidity for

403

β-lg/κ-car mixtures as a function of mixing ratio at different pHs: 4.4, 4.9, 5.0 and 5.1 (Figure S2);

404

ITC (Figure S3); Denaturation temperature of β-lg (Figure S4); Secondary structure of β-lg (Figure

405

S5); Theoretical modeling. This material is available free of charge via the Internet at

406

http://pubs.acs.org.

407

AUTHOR INFORMATION

408

Corresponding Author

409

*Email: [email protected]; Tel: +86-(0)-27-59750470.

410

Notes

411

The authors declare no competing financial interest.

412

ACKNOWLEDGEMENTS

21

ACS Paragon Plus Environment

Biomacromolecules

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 24

413

The research was supported by National Natural Science Foundation of China (31671811,

414

31322043, 31171751), projects from Hubei Provincial Department of Science and Technology

415

(2014BHE004, 2012FFA004) and Department of Education (T201307), program for New Century

416

Excellent Talents in University (NCET-12-0710), and project from Wuhan Science and Technology

417

Bureau (2015070504020218).

418

REFERENCES

419

1. Choi, J. Chem. Soc. Rev. 2011, 40, 5893-5909.

420

2. Tanrikulu, I. C.; Forticaux, A.; Jin, S.; Raines, R. T. Nat. Chem. 2016, DOI:10.1038/nchem.2556.

421

3. Tao, Y.; Yan, F.; Wu, X. Biomacromolecules 2007, 8, 2321-2328.

422

4. Ye, Y.; Blaser, G.; Horrocks, M. H.; Ruedasrama, M. J.; Ibrahim, S.; Orte, A.; Klenerman, D.;

423

Jackson, S. E.; Komander, D. Nature 2012, 492, 266-270.

424

5. Chen, X.; Xu, X.; Zhang, L.; Zeng, F. Carbohydr. Polym. 2009, 78, 581-587.

425

6. Chen, C.; Wu, W.; Xu, X.; Zhang, L.; Liu, Y.; Wang, K. Carbohydr. Polym. 2014, 78, 308-316.

426

7. Prajapati, V. D.; Maheriya, P. M.; Jani, G. K.; Solanki, H. K. Carbohydr. Polym. 2014, 105,

427

97-112.

428

8. Felsenfeld, G.; Groudine, M. Nature 2003, 421, 448-453.

429

9. Stingele, J.; Jentsch, S. Nat. Rev. Mol. Cell Bio. 2015, 16, 455-460.

430

10. Marcovitz, A.; Levy, Y. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 17957-17962.

431

11. Kayitmazer, A. B.; Seeman, D.; Minsky, B. B.; Dubin, P. L.; Xu, Y. Soft Matter 2013, 9,

432

2553-2583.

433

12. Turgeon, S. L.; Schmitt, C.; Sanchez, C. Curr. Opin. Colloid Interface Sci. 2007, 12, 166-178.

434

13. Du, X.; Dubin, P. L.; Hoagland, D. A.; Sun, L. Biomacromolecules 2014, 15, 726-734.

435

14. Elzoghby, A. O.; Samy, W. M.; Elgindy, N. A. J. Control. Release 2012, 161, 38-49.

436

15. Wu, B.; Degner, B.; McClements, D. J. J. Phys. Condens. Matter 2014, 26, 464104.

437

16. Yeo, Y.; Bellas, E.; Firestone, W.; Langer, R.; Kohane, D. S. J. Agr. Food Chem. 2005, 53,

438

7518-7525.

439

17. Bosnea, L. A.; Moschakis, T.; Biliaderis, C. G. Food Bioprocess Tech. 2014, 7, 2767-2781.

440

18. Li, X.; Fang, Y.; Alassaf, S.; Phillips, G. O.; Jiang, F. J. Colloid Interf. Sci. 2012, 388, 103-111. 22

ACS Paragon Plus Environment

Page 23 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Biomacromolecules

441

19. Cao, Y.; Fang, Y.; Nishinari, K.; Phillips, G. O. Sci. Rep. 2016, 6, 23739.

442

20. Kontopidis, G.; Holt, C.; Sawyer, L. J. Dairy Sci. 2004, 87, 785-796.

443

21. Brownlow, S.; Cabral, J. H. M.; Cooper, R.; Flower, D. R.; Yewdall, S. J.; Polikarpov, I.; North,

444

A. C. T.; Sawyer, L. Structure 1997, 5, 481-495.

445

22. Cao, Y.; Wang, L.; Zhang, K.; Fang, Y.; Nishinari, K.; Phillips, G. O. J. Phys. Chem. B 2015,

446

119, 9982-9992.

447

23. Grasdalen, H.; Smidsroed, O. Macromolecules 1981, 14, 229-231.

448

24. Schefer, L.; Adamcik, J.; Diener, M.; Mezzenga, R. Nanoscale 2015, 7, 16182-16188.

449

25. Schefer, L.; Adamcik, J.; Mezzenga, R. Angew. Chem. Int. Ed. 2014, 53, 5376-5379.

450

26. Schefer, L.; Usov, I.; Mezzenga, R. Biomacromolecules 2015, 16, 985-991.

451

27. Engelhardt, K.; Lexis, M.; Gochev, G.; Konnerth, C.; Miller, R.; Willenbacher, N.; Peukert, W.;

452

Braunschweig, B. Langmuir 2013, 29, 11646-11655.

453

28. Grasdalen, H.; Smidsroed, O. Macromolecules 1981, 14, 1842-1845.

454

29. Fang, Y.; Li, L.; Inoue, C.; Lundin, L.; Appelqvist, I. Langmuir 2006, 22, 9532-9537.

455

30. Li, X.; Fang, Y.; Alassaf, S.; Phillips, G. O.; Zhang, Y.; Zhao, M.; Zhang, K.; Jiang, F. Langmuir

456

2012, 28, 10164-10176.

457

31. Weinbreck, F.; De Vries, R. P.; Schrooyen; De Kruif, C. G. Biomacromolecules 2003, 4, 293-303.

458

32. Mekhloufi, G.; Sanchez, C.; Renard, D.; Guillemin, S.; Hardy, J. Langmuir 2005, 21, 386-394.

459

33. McGhee, J. D.; Von Hippel, P. H. J. Mol. Biol. 1974, 86, 469-489.

460

34. Vreeman, H. J.; Snoeren, T. H. M.; Payens, T. A. J. Biopolymers 1980, 19, 1357-1374.

461

35. Taulier, N.; Chalikian, T. V. J. Mol. Biol. 2001, 314, (4), 873-889.

462

36. Chourpa, I.; Ducel, V.; Richard, J.; Dubois, P.; Boury, F. Biomacromolecules 2006, 7,

463

2616-2623.

464

37. Marchin, K. L.; Berrie, C. L. Langmuir 2003, 19, 9883-9888.

465

38. Roach, P.; Farrar, D.; Perry, C. C. J. Am. Chem. Soc. 2005, 127, 8168-8173.

466

39. Bergkvist, M.; Carlsson, J.; Oscarsson, S. J. Biomed. Mater. Res. A 2003, 64, 349-356.

467

40. Timasheff, S. N.; Mescanti, L.; Basch, J. J.; Townend, R. J. Biol. Chem. 1966, 241, 2496-2501.

468

41. Structure-function properties of food proteins. In 1st ed.; Phillips, L. G.; Whitehead, D. M.;

469

Kinsella, J., Eds. Academic Press: California, 1994; p 51.

470

42. Djabourov, M.; Nishinari, K.; Ross-Murphy, S. B., Physical gels from biological and synthetic 23

ACS Paragon Plus Environment

Biomacromolecules

471

polymers. Academic Press: Cambridge University, 2013; p 127-144.

472

43. Madbouly, S. A.; Otaigbe, J. U. Macromolecules 2005, 38, 10178-10184.

473

44. Terech, P.; Weiss, R. G. Chem. Rev. 1997, 97, 3133-3160.

474 475

Table of Contents 1.2

Electrostatic complexation in soluble state

0.9

φ

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 24

Electrostatic complexation in insoluble state m −1  C MP  (1 − φ (r ))φ (r ) + 1 − φ (r )  − Km φ (r ) m r p w − C N (1 − φ (r )) = 0

0.6

 

m

 

 

M wPr

P

m

 

0.3 0.0 0

476

2

4

6

r =β -lg/κ-car

8

10

24

ACS Paragon Plus Environment