Structure and Dynamics of Heteroprotein Coacervates - Langmuir

Jun 29, 2016 - MES hydrate buffer was purchased from Sigma-Aldrich (St. Louis, MO, USA), and all other chemicals were from VWR (Radnor, PA, USA)...
0 downloads 0 Views 2MB Size
Subscriber access provided by - Access paid by the | UCSB Libraries

Article

Structure and dynamic of heteroprotein coacervates Paulo De Sa Peixoto, Guilherme M Tavares, Thomas Croguennec, Aurélie Nicolas, Pascaline Hamon, Claire Roiland, and Saïd Bouhallab Langmuir, Just Accepted Manuscript • DOI: 10.1021/acs.langmuir.6b01015 • Publication Date (Web): 29 Jun 2016 Downloaded from http://pubs.acs.org on June 30, 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.

Langmuir 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 26

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

Langmuir

1 2

Structure and dynamic of heteroprotein coacervates

3

4

5

6

Paulo D.S. Peixoto,† Guilherme M. Tavares,†,‡ Thomas Croguennec,†

7

Aurélie Nicolas,† Pascaline Hamon,† Claire Roiland,§ Saïd Bouhallab †,*

8 9 10 11



STLO, UMR1253, INRA, Agrocampus Ouest, 35000, Rennes, France

12



Laboratory of Research in Milk Products, Universidade Federal de Viçosa, BR-36570 Viçosa,

13

Brazil

14

§

15

74205, France

Institut des Sciences Chimiques de Rennes, UMR CNRS 6226, Université de Rennes 1 - CS

16 17 18 19 20 21 22

*corresponding author: Tel: +33 223485742; email: [email protected]

23

ACS Paragon Plus Environment

1

Langmuir

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 26

24

Abstract

25

Under specific conditions, mixing two oppositely charged proteins induces a liquid-liquid phase

26

separation. The denser phase, or coacervate phase, can be potentially applied as a system to protect

27

or encapsulate different bioactive molecules with a broad range of food and/or medical applications.

28

Optimization of the design and efficiency of such systems require a precise understanding of the

29

structure and the equilibrium of the nano-complexes formed within the coacervate. Here, we report

30

on the nano-complexes and the dynamics of the coacervates formed by two well-known, oppositely

31

charged proteins β-lactoglobulin (pI ≈ 5.2) and lactoferrin (pI ≈ 8.5). Fluorescence recovery after

32

photo bleaching (FRAP) and solid state NMR experiments indicate the co-existence of several

33

nano-complexes as primary units for the coacervation. This is the first (to the best of our

34

knowledge) evidence on the occurrence of an equilibrium between quite instable nano-complexes in

35

the coacervate phase. Combined with in silico docking experiments, these data support the fact that

36

coacervation in the present heteroprotein system depends not only on the structural composition of

37

the coacervates but also on the association rates of the proteins forming the nano-complexes.

38 39 40

Keywords: Heteroprotein coacervation, β-lactoglobulin, lactoferrin, primary units, molecular

41

diffusion, NMR

42

ACS Paragon Plus Environment

2

Page 3 of 26

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

Langmuir

43

1

INTRODUCTION

44

Complex coacervation results from the interaction of two oppositely charged colloids leading to a

45

liquid-liquid phase separation. The denser phase is called coacervates and the other phase is the

46

dilute phase or the equilibrium phase. There are numerous examples of complex coacervation

47

between oppositely charged biopolymers but the complex coacervation involving oppositely

48

charged proteins (heteroprotein coacervation) is rather rare. The mechanism of heteroprotein

49

complex coacervation, not completely elucidated, exhibits some specificity compared to complex

50

coacervation involving polyelectytes. Heteroprotein coacervation is observed in very narrow ranges

51

of pH, ionic strength, protein concentration and molar ratio. However, some common features for

52

heteroprotein coacervation exist even if the number of heteroprotein systems studied up to now is

53

too limited to established a general rules.

54

It is assumed that the formation of well-defined primary complex (building block) constitutes a

55

preliminary step to heteroprotein coacervation.1-4 The driving force for building block formation is

56

mainly electrostatic interactions with an entropy contribution due to counter-ions release. The

57

primary units have to be close to charge neutrality for liquid-liquid phase separation to occur and

58

consequently coacervation is only observed between the pI of the proteins involved in the

59

coacervation process. Protein size compensation is another requirement for heteroprotein

60

coacervation as two proteins of opposite charge but equal magnitude do not form coacervates if

61

they have different molecular size.5 It is also observed that the flexibility of at least one protein

62

favors heteroprotein coacervation.6-8 Protein flexibility could stabilize the building block by an

63

optimal exposition of amino acids involved in the interaction.4

64

Once formed, the primary units separate into a dense phase. Such mechanism suggests that protein

65

molar ratio in the coacervates is constant and corresponds to the protein stoichiometry of the

66

building block. This was observed for most of the systems studied up to now but the results are still

67

unclear for lactoferrin (LF) - β-lactoglobulin (β-LG) system. Under selected conditions, Anema and

68

de Kruif

9

proposed a β-LG/LF molar ratio of three while other authors proposed a molar ratio of

ACS Paragon Plus Environment

3

Langmuir

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 26

69

four.1,2 This difference was first explained by the origin of the proteins used for coacervation

70

experiments. Note that in these studies β-LG/LF molar ratio in the coacervates were determined

71

indirectly. By using the same protein source, we determined by direct quantification of the proteins

72

in the coacervates that the protein molar ratio in the coacervates varied from 4 to 6-8 by modifying

73

the protein molar ratio in the mixture.10 From the complexes in the dilute phase, Flanagan, et al.

74

proposed that the building block for LF - β-LG heteroprotein coacervation was a pentamer LF(β-

75

LG2)2 including two β-LG dimers bound to one LF. However, the structure of the coacervate was

76

analyzed by SANS and the results indicate that the proposed model does not exactly fit with the

77

experimental data.11 These authors proposed that the building block (or primary blocks) may adopt

78

different configurations and that some of them may associate into higher order equilibrium

79

structures. The purpose of the present study is to gain further insights into the internal structure and

80

the dynamic of the fascinating coacervates formed throughout specific interactions between

81

proteins. Here we investigate the molecular interactions and the diffusion properties of the proteins

82

in the highly concentrated LF/β-LG coacervate phase using Nuclear Magnetic Resonance (NMR),

83

Fluorescence Recovery after Photo Bleaching (FRAP) and docking simulations. The two

84

experimental technics allow access to complementary information at different time scales. Our data

85

show that one LF can bind multiple β-LG dimers with different affinities. FRAP and NMR

86

measurements support the coexistence of different molecular species in the coacervates. The

87

presence of various primary units in equilibrium constitutes a specific characteristic of BLG-LF

88

coacervates.

2

89 90

2

EXPERIMENTAL SECTION

91

2.1 Reagents and Solutions

92

Bovine β-LG, a mixture containing genetic variants A and B (pI ≈5.2), was provided by a

93

confidential industrial source. The protein powder was dispersed in deionized water (45 g/L),

94

adjusted to pH 5.2 with 1M HCl and kept at 30°C for 5 min, in order to precipitate non-native forms ACS Paragon Plus Environment

4

Page 5 of 26

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

Langmuir

95

of β-LG. The dispersion was centrifuged at 20,000g at room temperature for 10 min (Heraeus

96

Biofuge Primo, Thermo Scientific, Waltham, MA, USA) and the supernatant containing only native

97

β-LG was adjusted at pH 7.0 with 1 M NaOH, freeze-dried and stored at – 20 °C until use. Bovine

98

lactoferrin (LF, pI ≈8.5) (purity of 90% and iron saturation of 10 - 20 % according to technical

99

specification) was purchased from Fonterra Cooperative Group, New Zealand. MES hydrate buffer

100

was purchased from Sigma-Aldrich (St. Louis, MO, USA) and all other chemicals were from VWR

101

(Radnor, PA, USA).

102

β-LG and LF protein powders were solubilized in a 10 mM MES buffer, pH 5.5, at about 5.0 mM

103

and 0.7 mM respectively and filtered through a 0.2 µm membrane (cat. no. 4612, Pall Corporation,

104

Ann Arbor, MI, USA). The exact protein concentration was determined by absorbance at 280 nm

105

(SAFAS UV MC2, Safas, Monaco) using 0.96 L g-1 cm-1 and 1.47 L g-1 cm-1 as extinction

106

coefficients for β-LG and LF respectively. No sign of protein self-aggregation was detected in the

107

stock solutions that were transparent.

108 109

2.2 Preparation of β-LG-LF coacervates

110

The coacervates were obtained by mixing an appropriate volume of the protein stock solutions to

111

reach β-LG and LF final concentration of 0.6 mM and 0.06 mM respectively. These optimal

112

concentrations were defined in accordance with our previously work.10 The solution was prepared

113

respecting the following order of mixing: MES buffer + β-LG stock solution + LF stock solution.

114

Immediately after mixing the solution evolved in a system containing two liquid phases in

115

equilibrium: (i) a dilute phase and (ii) a dense phase dispersed in the dilute phase, called

116

coacervates. As described by Tavares, et al. 10 the coacervates were separated from the dilute phase

117

by centrifugation (Heraeus Biofuge Primo, Thermo Scientific, Waltham, MA, USA) at 20,000g for

118

10 min. Proteins in the coacervates were quantified by reverse phase chromatography as previously

119

described by Tavares, et al. 10.

120

The coacervate samples studied here were a translucent, slightly pink and viscous liquid containing ACS Paragon Plus Environment

5

Langmuir

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 26

121

around 150 g/kg of LF and 130 g/kg of β-LG, i.e. a β-LG/LF molar ratio about 4, which is in

122

accordance with the ratio observed by other authors.

123 124

2.3 Evaluation of β-LG and LF heterocomplex by rigid docking

125

A rigid docking experiment was conducted to determine the near-native orientation of β-LG (PDB

126

id: 1BEB) and LF (PDB id: 1BLF) using the web server pyDockWeb.12 The software creates

127

10,000 different complexes by changing the relative position of the ligand around the receptor.

128

Then a scoring energy was calculated and the complexes were ranked according. Best poses were

129

assumed to be the ones that displayed the lowest scoring energy. Three docking were performed: i)

130

LF as receptor and β-LG dimer (noted β-LG2) as ligand; ii) The best LF(β-LG2) complex obtained

131

as receptor and another β-LG dimer as ligand; iii) The best LF(β-LG2) complex against himself.

132

The 100 best complexes were clustered by their pairwise Root Mean Square deviation (RMSD) of

133

the Cα atoms with ensemble cluster tool13 embedded in Chimera software14 to reduce and

134

discriminate docking solutions. Consequently we assume that the best docking solutions were the

135

complexes in the most populated clusters. Structures were visualized with VMD.15

136 137

2.4 FRAP: analysis of β-LG binding to LF

138

FRAP analysis was used to characterize the binding between FITC (Fluorescein isothiocyanate)

139

labeled β-LG and LF. In solution where a fluorescent molecule displays a classical diffusion

140

behavior, the radius of a Gaussian bleached spot increases with time after photo-bleaching.16 FRAP

141

recovery curve can be then used to estimate the molecular diffusion constant of the fluorescent

142

molecule. In contrast, in a solution where the fluorescent molecules are strongly bound to immobile

143

obstacles, the radius of a spherical bleached spot does not increase with time.16 In this reaction

144

dominant regime, the time of diffusion of a “free” fluorescent molecule is shorter than the diffusion

145

time of a complex (with bound fluorescent probe). Thus, the large scale fluorescent probe diffusion

146

measured by FRAP is exclusively dictated by the time that the probe remains bound to the obstacle. ACS Paragon Plus Environment

6

Page 7 of 26

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

Langmuir

147

Therefore, FRAP recovery curve informs on the probe-obstacle dissociation rate (κoff) and on the

148

concentration of fluorescent probes bound to the obstacle (Ceq). For reaction dominant regime, the

149

FRAP recovery data can be fitted by the following equation:

150

Eq. 1.

151 152 153

Where frap(t) is the fluorescence intensity at time t and Feq is the concentration of unbound

154

fluorescent molecule.17

155

This equation only fits the experimental data if all different probe-obstacle complexes present the

156

same dissociation rate. However, we found a dependence between the lifetime of the LF(β-LG2)n

157

complexes and the number of β-LG2 bound to LF in the current work. Considering that β-LG2 binds

158

to LF in two different affinity sites,1,

159

equation:17

160

Eq. 2.

10

FRAP data were fitted using a double exponential

161 162 163

The parameters of Eq. 2 are the same as in Eq. 1 and the indices 1 and 2 refer to the complexes with

164

longer and short lifetimes, respectively.

165

Complementarily, it is also possible that one of the two β-LG2 bound to LF presents a too weak

166

affinity to be considered in the reaction dominant regime. In this case, the fraction of β-LG2

167

strongly bound to LF presents a reaction dominant regime, while the other fraction presents a

168

different regime characterized by a significant diffusion time of free β-LG2 between LF-β-LG2

169

complexes (obstacles). In this case, the following equation must be used:17

170

Eq. 3.

171

ACS Paragon Plus Environment

7

Langmuir

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 8 of 26

172 173

The last term in Eq. 3 describes the contribution of the β-LG2 bound to the LF site of highest

174

affinity (identical term of the Eq. 2). In contrast, the first term of the equation describes the

175

contribution of the β-LG2 fraction bound to the LF site of weakest affinity. The equation parameters

176

are the same previously described and τeff is defined as:17

177

Eq. 4.

178 179 180

Where w is the radius of the bleached spot, Deff and Df are the effective diffusion (the hindrance

181

induced by diffusion and binding) and the pure diffusion (the hindrance induced by diffusion only)

182

respectively and k

183

of hindered diffusion for β-LG2 found experimentally18 and theoretically in equivalent dense

184

systems. The best fit was determined according to the methodology described in Supporting

185

Information (Figure S1). Note that the immobility is a relative notion. If the recovery time of the

186

signal is faster than that taken by a macromolecule complex, a so called obstacle, to do a significant

187

displacement (µm scale), the last can be considered as immobile regarding the FRAP experiment.

188

This is not true for NMR because of much smaller time scales and high sensitivity to the rotation

189

motion of quite large complexes.

190

Finally, the LF(β-LG2)n complexes are not totally immobile and their diffusion is significant

191

regarding the time range of the experiment. The diffusion of these complexes themselves was

192

followed and treated as a classical but slow diffusion behavior. If a fraction of β-LG2 forms a stable

193

complex with LF in the time range of the experiment (strongly bound) and another fraction displays

194

a dynamical association (weakly bound) with LF, the Eq. 3 can be used to model the fluorescence

195

recovery curve.

196

Basically, the coacervate phase for the FRAP experiments was prepared as described above. The

on

is the association rate. The value of Df used was 15 µm2/s which is the value

ACS Paragon Plus Environment

8

Page 9 of 26

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

Langmuir

197

proportion of 1% of the bulk β-LG concentration was replaced by the FITC labeled β-LG. The

198

FITC-β-LG was obtained as described by Silva, et al. 19

199

The coacervate phase was placed between glass slide and a coverslip sealed with a small adhesive

200

frame (25 µL capacity). The fluorescence recovery images were obtained using an inverted CLSM

201

(Nikon, Champigny-sur-Marne, France). The observations were realised with a 40 times

202

magnification oil immersed objective at 30 µm from the coverslip. The labeled β-LG in the

203

coacervate phase was excited with 50 mW sapphire laser system at a wavelength of 488 nm and

204

detected from 500 to 530 nm. The fluorescence before photobleaching was recorded using 0.1% of

205

the maximum laser intensity. The photobleaching was realized using 100 % of the laser intensity in

206

an 85 µm² spot and the recovery of fluorescence was followed during 60 min. Obtained images

207

were treated using ImageJ free software.

208 209 210

2.5 NMR: Estimation of hydrodynamic radius and relative abundance of the different β-LG/LF complexes

211

Solid-state NMR experiments were performed on a Bruker Avance III 600 SB (14T) operating at

212

Larmor frequencies of 600.1 MHz for 1H using a 4 mm MAS double resonance probehead. 1H

213

NMR spectra were acquired under static or slow Magic Angle Spinning (MAS at 500 Hz)

214

conditions using a single pulse experiment (t90 1H = 3.6 µs). The recycle delays used to record

215

proton spectra were set to 1 s which was long enough to ensure quantitatively. Moreover, 1H spectra

216

of the sample remained unchanged after 24 h, indicating that the analyzed suspension was stable

217

under MAS as well as in static conditions. Longitudinal (T1) and transverse (T2) relaxation times

218

were measured using a saturation/recovery and spin echo pulse sequences, respectively. 1H spectra

219

were referenced according to Tetramethylsilane (TMS). Spectra deconvolutions were performed

220

using the dmfit softwere as detailed in Supporting Information. To check the repeatability of the

221

protocol, relaxation time has been measured in three different samples.

ACS Paragon Plus Environment

9

Langmuir

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 10 of 26

222

The hydrodynamic radius (Rh) of β-LG2, LF and the formed complexes was estimated from their

223

rotation correlation time (τc), which is strongly correlated to the 1H translation relaxation times, T1

224

and T2. In diluted solution, τc is related to Rh of the molecule as described below:20

225

Eq. 5.

226 227 228

Where η is the solvent viscosity, kb is the Boltzmann constant and T is the temperature.

229

In dense solution, the hydrodynamic interactions between particles are affected by the solvent

230

viscosity (η), the determination of which is not trivial in particular for polydisperse systems.21 The

231

value of η increases with the molecular crowding and the size of the diffusive particles. Theoretical

232

studies showed that the molecular motion of small particles (~1 to 7 nm of radius) displays a quasi-

233

linear relationship with Rh at short time scales (nano to miliseconds), in protein density range

234

equivalent to the ones of the present study (250 to 350 g/L).22,23 Thus, even if it is not possible to

235

exactly determine η (since the composition of the medium is unknown) the quasi-linear relationship

236

between the size of the molecule and τc can be used to roughly estimate a relative Rh for the

237

different species under study. More details on the Rh estimation from relation times are given in

238

Supplementary Information.

239 240

Determination of the relative abundance of the different β -LG/LF complexes from the NMR

241

1

242

The peak width in NMR is very sensitive to the dynamics of the molecular species in solution. If the

243

association/dissociation rate of the complexes is slow regarding NMR time scales (nano to

244

milliseconds), peaks with different widths are observed on a 1D 1H spectrum: complexes with slow

245

dynamics produce broad peaks while their primary units having fast dynamics exhibit narrow peaks.

246

Based on this, the relative abundance of the complexes was directly estimated by measuring the

H spectrum.

ACS Paragon Plus Environment

10

Page 11 of 26

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

247

Langmuir

signal intensity of each peak in a 1D 1H spectrum.

248 249

3

RESULTS

250

3.1 β-LG and LF complexes: rigid Docking

251

The association of one β-LG dimer (β-LG2) and one LF has been evaluated by rigid-body docking

252

using PyDockWeb server. The 100 lowest-energy complexes structures were clustered by pairwise

253

RMSD to reduce docking solutions. Nineteen clusters were obtained and a representative structure

254

by cluster was extracted for further analyses. All of them showed the same LF region (called S for

255

single) interacting strongly with β-LG dimer (Figure 1A). The S site of LF exhibited a very high

256

density of positively charged residues (inset of Figure 1A) suggesting electrostatic interactions as

257

the main driving forces of protein association. Different orientations of β-LG2 were observed

258

supporting various interaction sites on β-LG2 surface. The associated LFβ-LG2 complex was further

259

used as receptor to simulate a second interacting site on LF. The 100 best docking solutions were

260

clustered and highlighted various interaction sites. Two of the major interaction sites were identified

261

in LF (called M for Multiple) representing 48 and 31 % among the best poses (Figure 1B). These M

262

sites are not identified in the first docking experiment, suggesting the involvement of lower binding

263

forces. Furthermore, M sites are localized on the surface in the hinge between the two lobes of LF.

264

This suggests that LF can bind at least one β-LG2 in addition to anther β-LG2 already linked to S

265

site of LF.

ACS Paragon Plus Environment

11

Langmuir

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 26

266 267

Figure 1. Molecular docking between β-LG2 and LF (A), and between β-LG2 and already formed

268

LFβ-LG2 complex (B). (A) Structure of the best complex formed by β-LG2 in red and LF in blue.

269

Other docking solutions of β-LG2 are represented in translucent grey. β-LG2 binds a specific site S,

270

in all docking solutions. The inset shows that the S site on LF exhibits a huge density of positively

271

charged residues (blue) in contrast to the other parts of the protein with positively (blue) and

272

negatively (red) charged residues homogeneously distributed. (B) Illustration of the secondary sites

273

on LF (called site M) where a second β-LG2 binds (with indicated percent) to the LFβ-LG2

274

complex.

275 276

3.2 β-LG binding within the coacervate phase: FRAP analysis

277

Figure 2A shows a bleached spot profile, immediately (0.5 s) and one hour after bleaching. Figure

278

2B shows the corresponding FITC labeled β-LG fluorescence recovery curve. The radius of the

279

bleached spot profile displays only a small increase one hour after bleaching indicating that the

280

diffusion of β-LG2 within the coacervates is dominated by the binding of β-LG2 to immobile

281

obstacles (reaction regime, see experimental section for details). Only 70% of initial β-LG

282

fluorescence intensity was recovered one hour after bleaching. This is a remarkable slow recovery ACS Paragon Plus Environment

12

Page 13 of 26

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

Langmuir

283

even for a system as dense as this one17, 20 indicating that (i) β-LG binds to the obstacles for a long

284

time and (ii) the obstacles are large enough to consider their diffusion negligible after one hour.

285

Since the bleached spot analysis indicates that the system is reaction dominated, the recovery curve

286

has been fitted assuming a model in which the binding of β-LG2 to LF dominate the fluorescence

287

recovery. A model in which one single dissociation time for all complexes is considered did not

288

correctly fit the experimental data (Figure 2B, red line). Proper fits were obtained with models

289

considering either two different dissociation times for the complexes (Figure 2B, green continuous

290

line) or both the diffusion of β-LG and its binding to complexes (Figure 2B, green dotted line). The

291

fitting parameters of the last two cited models are displayed in Table 1.

292

293 Figure 2. FRAP analysis of the FITC labeled β-LG within the coacervate phase. (A): Time evolution of the bleached spot profile from a FRAP experiment using an initial bleached radius of 5.2 µm. (B): Typical FRAP recovery curve (blue) and the fits using the models described in materials and methods section. The red line represents the best fit using the model considering the diffusion of β-LG2 dominated by binding and that all complexes display the same lifetime. The green continuous curve represents the best fit using a model considering the diffusion of β-LG2 dominated by its binding to complexes which display two different lifetimes. The green dotted curve represents the case where both, β-LG2 diffusion and its binding to complexes were considered for the fluorescence recovery. 294 ACS Paragon Plus Environment

13

Langmuir

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 26

Table 1. FRAP fitting parameters

Feq (%)

Reaction dominant 6

Reaction/Diffusion 70*

Ceq1 (%)

45

70*

Ceq2 (%)

49

30

k1 off (s-1)

1.0 x 10

-3

8.6 × 10-3

k2 off (s-1)

1.2 x 10-4

5.2 × 10-4

k1 on (s-1)

-

30

295

15 Df (µm²/s) * In the reaction/diffusion model, Feq and Ceq1 (see equation 3 in experimental section) cannot be

296

distinguished; the value of 70% corresponds to both Feq and Ceq1.

297 298

The dissociation rates k1off and k2off range from 1 to 8 ms-1 and 0.1 to 0.5 ms-1 for the pure reaction

299

and reaction/diffusion models, respectively (Table 1). The dissociation times are slightly lower in

300

the pure reaction model than in the reaction/diffusion model. However, in both models, the first

301

dissociation rate (k1off) was about one order of magnitude higher than the second constant (k2off). In

302

contrast, the relative number of β-LG molecules (Ceq) associated to k1off and k2off was quite

303

different. The number of molecules associated to both rates was around 50/50 in the reaction

304

dominant model. In contrast, more molecules (70%) were associated to the first dissociation rate

305

compared to the second one in the reaction/diffusion model.

306 307 308

3.3 Rotation diffusion coefficients show the presence of three different complexes: NMR analysis

309

Besides the water signal at 4.9 ppm, three signals were distinguished in the 1H spectrum of the

310

coacervate phase (Figure 3A): a large signal with a full width at half maximum (fwhm) of about 50

311

ppm, a second signal with 20 ppm of fwhm and a narrow signal of 2-4 ppm of fwhm (see

312

Supporting Information for details). Protons with the largest fwhm also present the slowest

313

dynamic. The narrow signal displays the typical pattern of water and the others signals are assigned

ACS Paragon Plus Environment

14

Page 15 of 26

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

Langmuir

314

to protons of proteins. The 1H 20 ppm and the 1H 50 ppm corresponds to complexes displaying a

315

relative fast and slow dynamics respectively. Noteworthy, this was further confirmed on the 2D 1H-

316

13

317

dynamics appeared. The CP spectrum displays only the signal of the largest Gaussian with a fwhm

318

of 50 ppm (green dotted lines in Figure 3A). A focus at the center of the 1H spectrum (Figure 3C)

319

shows individual peaks assigned to the amino acids of the proteins composing the second signal and

320

the underneath spectrum of the slowest species described above. Fig. 3D shows that the signal of

321

the slowest species is quite weak after 20 µs using an echo experiment (that measures T2). In

322

contrast, the intensity of the second signal barely changed confirming that their relaxation times,

323

and consequently their dynamics, are extremely different. The deconvolution of the 1H spectrum

324

gave the fraction of protons assigned to the slowest species (Supplementary Information). The peak

325

of the slowest species (largest signal) represents about 52% of the total 1H NMR signal intensity of

326

the proteins, against 48% for the second signal composed of two species (see below): 16% and 32%

327

for the fastest and the moderately slow species, respectively (Table 2).

328

For comparison, Figure 3E and F show the spectra of the solutions of β-LG (Figure 3E) and LF

329

(Figure 3F) at 150 g/L, under the same conditions of pH and temperature. Similar results are

330

obtained when denser solutions (~300 g/L) of β-LG or LF were analyzed.

C cross-polarization (CP) spectrum (Figure 3B) on which only the protons displaying a very slow

ACS Paragon Plus Environment

15

Langmuir

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 26

331 332

Figure 3. NMR signal Assignments. (A) Three main signals detected in 1H spectrum of the

333

coacervate phase: the largest signal is assigned to the association of several LF and β-LG2

334

molecules i.e. (LFβ-LG2)n, the second signal is assigned to β-LG /β-LG2 and LF(β-LG2)2 and the

335

third signal is assigned to water molecules. (B) 2D 1H-13C cross polarization spectrum of the

336

coacervate phase. In the spectrum only the protons having a very slow dynamic display a reasonable

337

intensity. (C) Zoom at the center of the spectrum A focusing on the peaks emanating from β-LG2

338

and LF(β-LG2)2. (D) 1H spectrum using echo filter. In the spectrum between 5 and 20 µs the signal

339

of the largest peak is completely eliminated without affecting significantly the intensity of the other

340

peaks (the water and the second peak corresponding to β-LG2 and LF(β-LG2)2. (E) and (F), 1H

341

spectra of dense β-LG and LF solutions respectively.

342 343

The measurement of the spin relaxation times T1 and T2 gives access to quantitative data about the

344

protein dynamics (notably the protein rotational correlation time as detailed in Supplementary

ACS Paragon Plus Environment

16

Page 17 of 26

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

Langmuir

345

Information). Figure 4 shows the spectra at different delays (from 0 to 3 s) used to measure T1.

346

After deconvolution, two mean protein signals with different dynamics were obtained. The

347

measurement of the T1 of each signal can be determined independently (see experimental section

348

and Figures S2 and S3). For the detected signals, the time evolution echo can be fitted with either

349

mono-exponential (largest peak) or double-exponential (second peak) curves. This indicates that

350

relaxation times are dominated by the global dynamics of the proteins. Based on hydrodynamic

351

radius calculation (Table 2), the largest signal is assigned to complexes composed of several LF and

352

β-LG2 noted (LFβ-LG2)n. From calculated Rh, molecular species associated to the second signal

353

were assigned to β-LG /β-LG2 and LF(β-LG2)2, respectively.

354 355

356 Figure 4. Zoom-in view of NMR signals (insert) for assignment to different complexes in the coacervates. Spectra corresponding to the different delays used to calculate T1 (delays from 0 to 3 s): Green delimited area: large complexes; red delimited area: small molecular entities.

ACS Paragon Plus Environment

17

Langmuir

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 26

Table 2. Abundance (w/w) and dynamics of the complexes in the coacervate phase % Protons

T1 /T2

% β -LG*

Rh (nm)**

β-LG/β /β-LG /β 2

16

2

≈ 33

2

LF(β β-LG2)2

32

17

≈ 33

7

(LFβ β-LG2)n

52

200

≈ 33

30-60

357

* Relative proportion of β-LG molecules in each complex, calculated on the basis of NMR data.

358

** Calculated using the values of T1/T2 ratio (see experimental section and Supplementary

359

Information).

360 361

4

DISCUSSION

362

4.1 Primary units of the coacervates

363

It has been postulated that LF-β-LG coacervates result from the association of heterocomplexes

364

formed by LF and β-LG2.1, 10 Docking simulations showed that each LF molecule is able to bind 2

365

or more β-LG2. The first β-LG2 always bind to the same site of higher affinity on LF surface (site S,

366

Figure 1B). In contrast, the second β-LG2 can bind on different sites of lower affinity, called M

367

sites, located all around LF surface. The theoretical evidence of the existence of two sites of

368

different affinity for β-LG (sites S and M) on each LF confirms the results obtained previously by

369

isothermal calorimetry (ITC)10 and electrostatic modeling.2 The coexistence of these

370

heterocomplexes was already suggested by Flanagan, et al. 2. Their relative abundance depends on

371

β-LG/LF initial molar ratio. In support to docking simulation, FRAP experiments show that β-LG

372

bind to immobile obstacles with two quite different affinities. The immobile obstacles could be

373

either an individual LF monomer or an already formed complex. The fit of FRAP data reveals

374

dissociation rates of about 1 to 8 ms-1 for M sites and 0.1 to 0.5 ms-1 for the S site. Both these values ACS Paragon Plus Environment

18

Page 19 of 26

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

Langmuir

375

correspond to a relative weaker binding strength compared to the ones described for most biological

376

complexes displaying a specific binding (dissociation rate of about 10-3 ms-1 for protein receptors).17

377

Given the strong affinity of LF S-site for β-LG2, the probability of having free LF in the coacervates

378

should be weak under molar excess of β-LG compared to LF. NMR data indicate that β-LG

379

distributes equally to the various species in the coacervate i.e. the free β-LG and the two

380

heterocomplexes LF(β-LG2)2 and (LFβ-LG2 )n (see below). From HPLC quantification, 7.2 mM β-

381

LG were found in the coacervates meaning 2.4 mM in each of the above species. Consequently, the

382

LF concentrations involved in LF(β-LG2)2 and (LFβ-LG2)n heterocomplexes are expected to be 0.6

383

mM and 1.2 mM, respectively. This match perfectly with the quantified LF value of 1.87 mM found

384

in the coacervate phase in which, consequently, less than 5% of LF exists as free entity. Hence,

385

based on protein quantification and ITC experiments10, we suggest that under these conditions, LF

386

molecule cannot bind, in significant proportion, more than two β-LG2 simultaneously.

387 388 389

4.2 Identification of the structures and quantification of the main species in the coacervates

390

Based on our experiments, three types of molecular entities with specific dynamics and Rh seem to

391

be present in LF-β-LG coacervates. The smaller one is assigned to β-LG2 and β-LG monomers (Rh

392

= 2 nm). Considering a dissociation constant Kd of 5.5×10-4 M at working pH of 5.5

393

LG2/β-LG molar ratio is roughly 60/40 for the non-complexed β-LG in the coacervates (2.4 mM).

394

The other entities with Rh of 7 nm and ≈30-60 nm are assigned to heterocomplexes involving LF

395

and β-LG2. We suggest that the complex with 7 nm determined by NMR corresponds to LF(β-

396

LG2)2 while complexes with Rh ranging from ≈ 30 to 60 nm results from the association of several

397

LFβ-LG2 units to form (LFβ-LG2)n, the “skeleton of the coacervates”. In this skeleton, we suggest

398

that the proportion of LF molecules having more than two branched β-LG2 is limited due to steric

399

hindrance leading to the specific and oriented growth of n×LFβ-LG2 into (LFβ-LG2)n

ACS Paragon Plus Environment

10

, the β-

19

Langmuir

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 26

400

heterocomplexes. An association of LF(β-LG2)2 pentameter at the ends of this skeleton structure

401

hinders its further growth leading to a particle definite size. The presence of free LFβ-LG2 which

402

has a Rh close to the one of LF(β-LG2)2 in the coacervates is not excluded. But for a sake of

403

simplification, we assume from NMR data that the main entities in the coacervates are β-LG2, LF(β-

404

LG2)2 and (LFβ-LG2)n with a relative proportions, based on proton assignment, of 17%, 33% and

405

50% respectively (Table 2). Since the number of protons of β-LG represents ∼50% and ∼33% of the

406

total number of protons in LF(β-LG2)2 and (LFβ-LG2)n respectively, we deduced that β-LG

407

molecules in the coacervates distributes almost equally between the three entities. Based on the

408

protein composition in the coacervates determined by NMR, a rapid calculation of the β-LG/LF

409

molar ratio gives ∼4, a value in agreement with the ratio found by protein quantification previously

410

reported.1,10 Combining our results, we propose a schematic molecular composition of the β-LG-LF

411

coacervate phase (Figure 5). This phase exhibits heterogeneous composition with a subtle mix of

412

free β-LG, heterooligomers and heterocomplexes. The number of primary units that forms the

413

heterocomplexes has been chosen considering a Gaussian distribution. They are diluted in a “sea”

414

of β-LG2 and LF(β-LG2)2 complexes. Under the experimental conditions, the number of β-LG2 is

415

close to twice the number of LF(β-LG2)2.

416 417

Figure 5. Two-dimensional representation of the β-LG-LF coacervate phase composition. LF: blue ACS Paragon Plus Environment

20

Page 21 of 26

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

Langmuir

418

full ellipses; β-LG monomers or dimers: red full spheres. Dark circles indicate large

419

heterocomplexes containing several LF and β-LG2 (LFβ-LG2)n. Red circles indicate LF(β-LG2)2

420

complexes. The ratio of each species is derived from NMR data as described in the text.

421 422

4.3 Specific thermodynamic equilibrium governs the stability of the coacervates

423

The LF(β-LG2)2 pentamer was reported to be the primary unit of the LF-β-LG coacervate.11 This

424

suggests that LF(β-LG2)2 entities are very abundant and stable in solution. Consequently, the

425

conditions of formation of LF-β-LG coacervates should be quite resistant to variation of both total

426

protein concentration and β-LG/LF molar ratio. Experimentally, we observed that random

427

aggregates instead of coacervates were formed in mixtures containing β-LG/LF molar ratio as high

428

as 20.10 Increasing β-LG concentration in LF solution should drive the binding equilibrium toward

429

the saturation of the LF binding sites resulting in larger amount of LF(β-LG2)2 and/or LF with more

430

than two bound β-LG2. These multi-branched LF would lead to non-specific associations into

431

random aggregates.

432

The results presented here give an explanation to previous indirect assumptions. We propose that

433

the coacervates result from the coexistence of three metastable entities in dynamic equilibrium: β-

434

LG2, LF(β-LG2)2 and larger complexes (LFβ-LG2)n. Structures with larger size were not detected in

435

the dilute phase according to Flanagan et al.2 and confirmed here by DLS measurements (not

436

shown). Hence, unlike the small molecular species, (LFβ-LG2)n is probably in equilibrium with

437

other species only in the coacervate phase. Any subtle changes of these equilibria by modifying the

438

physicochemical conditions (protein concentration, β-LG/LF molar ratio, ionic strength and pH)

439

could change the relative abundance of each entity. This could explain the reported evolution of the

440

β-LG/LF molar ratio in the coacervation domain and also the transition from coacervates to

441

aggregates for larger changes.10 It could also explain the extreme sensitivity of LF-β-LG

442

coacervation by changing ionic strength and pH when compared to similar liquid-liquid systems

ACS Paragon Plus Environment

21

Langmuir

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 26

443

involving large colloids and polyelectrolytes. Ionic strength and pH conditions affect the

444

equilibrium between electrostatic forces responsible for the long/short range attraction/repulsion in

445

most liquid-liquid systems. The properties of LF-β-LG coacervates are dictated by the balance of

446

the interaction between entities and their relative abundance. Any changes in the physico-chemical

447

parameters have a double impact on the present system: it modifies the self-repulsion/attraction

448

properties of each complex (as for colloids and polyelectrolytes) but also their relative abundance.

449

This explains why the present system is much more sensitive to changes in physicochemical

450

conditions than the ones composed by more stable units (colloids or polyelectrolytes). Complexes

451

with constant stoichiometry were generally reported as primary units for coacervation between

452

oppositely charged polyelectrolytes. Here, we evidence the co-existence of complexes with various

453

stoichiometries in dynamic equilibrium forming the inner structure of heteroprotein coacervates.

454

We therefore confirm the existence of a dynamic equilibrium into LF-β-LG coacervates suggested

455

by Dubin’s group from rheology and neutron scattering experiments.11

456 457

To conclude, this study gives new insights on the structure and dynamic of heteroprotein

458

coacervates. The stability of the heteroprotein coacervates is governed by the presence of non-

459

random nano-complexes in fast equilibrium with smaller entities. The structure of the nano-

460

complexes and their dynamic prevents the formation of large, random aggregates at mesoscale and

461

is responsible of the liquid behavior of the dense phase at macroscopic scale. In the context of drug

462

delivery systems, these data indicate that the key factor governing the formation of the delivery

463

protein vehicle are the association rates of the proteins involved rather than the specific properties

464

of the complexes formed. Thus, the scientific effort to identify physico-chemical conditions for

465

optimizing the formation of delivery protein vehicle should target on a detailed study of protein

466

association dynamics. Also, further studies are still needed to claim generality of our finding on LF-

467

β-LG to other heteroprotein systems.

468

ACS Paragon Plus Environment

22

Page 23 of 26

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

Langmuir

469

Supporting Information

470

The Supporting Information is available free of charge at http://pubs.acs.org. Modeling of FRAP

471

data; Estimation of the hydrodynamic radii from NMR experiments: i- Measure of relaxation times,

472

ii- spectra simulations.

473 474 475

5

REFERENCES

476

(1)

477

Schmitt, C. Heteroprotein complex coacervation: bovine beta-lactoglobulin and lactoferrin.

478

Langmuir 2013, 29, 15614-23.

479

(2)

480

Bovetto, L.; Schmitt, C. Complex equilibria, speciation, and heteroprotein coacervation of

481

lactoferrin and beta-lactoglobulin. Langmuir 2015, 31, 1776-83.

482

(3)

483

2016, 55, 89-99.

484

(4)

485

R.; Schwalbe, H.; Croguennec, T.; Bouhallab, S., et al. Investigation at Residue Level of the Early

486

Steps during the Assembly of Two Proteins into Supramolecular Objects. Biomacromolecules 2011,

487

12, 2200-2210.

488

(5)

489

drive spontaneous self-assembly of oppositely charged globular proteins into microspheres. J. Phys.

490

Chem. B 2010, 114, 4138-44.

491

(6)

492

supramolecular structures resulting from alpha-lactalbumin-lysozyme interaction. Biochemistry

493

2007, 46, 1248-55.

Yan, Y.; Kizilay, E.; Seeman, D.; Flanagan, S.; Dubin, P. L.; Bovetto, L.; Donato, L.;

Flanagan, S. E.; Malanowski, A. J.; Kizilay, E.; Seeman, D.; Dubin, P. L.; Donato-Capel, L.;

Delboni, L. A.; da Silva, F. L. B. On the complexation of whey proteins. Food Hydrocolloid

Salvatore, D.; Duraffourg, N.; Favier, A.; Persson, B. A.; Lund, M.; Delage, M. M.; Silvers,

Desfougeres, Y.; Croguennec, T.; Lechevalier, V.; Bouhallab, S.; Nau, F. Charge and size

Nigen, M.; Croguennec, T.; Renard, D.; Bouhallab, S. Temperature affects the

ACS Paragon Plus Environment

23

Langmuir

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 26

494

(7)

Anema, S. G.; de Kruif, C. G. Coacervates of lysozyme and beta-casein. J. Colloid Interface

495

Sci. 2013, 398, 255-61.

496

(8)

497

Milk after the Binding of Lactoferrin or Lysozyme. J Agr Food Chem 2013, 61, 7142-7149.

498

(9)

499

lactoglobulin. J. Colloid Interface Sci. 2014, 430, 214-20.

500

(10)

501

coacervation between lactoferrin and the two isoforms of β-lactoglobulin. Food Hydrocolloid 2015,

502

48, 238-247.

503

(11)

504

Schmitt, C. Structure of bovine beta-lactoglobulin-lactoferrin coacervates. Soft Matter 2014, 10,

505

7262-8.

506

(12)

507

protein-protein docking using electrostatics and desolvation scoring. Bioinformatics (Oxford,

508

England) 2013, 29, 1698-9.

509

(13)

510

ensemble of NMR-derived protein structures into conformationally related subfamilies. Protein Eng

511

1996, 9, 1063-1065.

512

(14)

513

C.; Ferrin, T. E. UCSF Chimera--a visualization system for exploratory research and analysis. J.

514

Comput. Chem. 2004, 25, 1605-12.

515

(15)

516

1996, 14, 33-8, 27-8.

517

(16)

518

of proteins in complex with inositol lipids serves to coordinate free movement with spatial

519

information. J. Cell Biol. 2009, 184, 297-308.

Anema, S. G.; de Kruif, C. G. Protein Composition of Different Sized Casein Micelles in

Anema, S. G.; de Kruif, C. G. Complex coacervates of lactotransferrin and beta-

Tavares, G. M.; Croguennec, T.; Hamon, P.; Carvalho, A. F.; Bouhallab, S. Selective

Kizilay, E.; Seeman, D.; Yan, Y.; Du, X.; Dubin, P. L.; Donato-Capel, L.; Bovetto, L.;

Jimenez-Garcia, B.; Pons, C.; Fernandez-Recio, J. pyDockWEB: a web server for rigid-body

Kelley, L. A.; Gardner, S. P.; Sutcliffe, M. J. An automated approach for clustering an

Pettersen, E. F.; Goddard, T. D.; Huang, C. C.; Couch, G. S.; Greenblatt, D. M.; Meng, E.

Humphrey, W.; Dalke, A.; Schulten, K. VMD: visual molecular dynamics. J. Mol. Graph.

Hammond, G. R.; Sim, Y.; Lagnado, L.; Irvine, R. F. Reversible binding and rapid diffusion

ACS Paragon Plus Environment

24

Page 25 of 26

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

Langmuir

520

(17)

Sprague, B. L.; McNally, J. G. FRAP analysis of binding: proper and fitting. Trends Cell

521

Biol. 2005, 15, 84-91.

522

(18) Peixoto, P. D.; Bouchoux, A.; Huet, S.; Madec, M. N.; Thomas, D.; Floury, J.; Gésan-Guiziou,

523

G. Diffusion and Partitioning of Macromolecules in Casein Microgels: Evidence for Size-

524

Dependent Attractive Interactions in a Dense Protein System. Langmuir 2015, 31(5), 1755-1765.

525

(19)

526

the influence of the charge and shape of solutes on diffusion. J. Dairy Sci. 2013, 96, 6186-98.

527

(20) Rule, G. S.; Hitchens, T. K. Fundamentals of protein NMR spectroscopy. Focus on structural

528

biology. 2006, vol 5, Springer New Delhi, India.

529

(21)

530

NMR 1993, 3, 335-348.

531

(22)

532

macromolecular motion. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 18457-18462.

533

(23) Cai, L. H.; Panyukov, S.; Rubinstein, M. Mobility of nonsticky nanoparticles in polymer

534

liquids. Macromolecules, 2011, 44(19), 7853-7863.

Silva, J. V.; Peixoto, P. D.; Lortal, S.; Floury, J. Transport phenomena in a model cheese:

Boulat, B.; Bodenhausen, G. Measurement of proton relaxation rates in proteins. J. Biomol.

Ando, T.; Skolnick, J. Crowding and hydrodynamic interactions likely dominate in vivo

535

ACS Paragon Plus Environment

25

Langmuir

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 26 of 26

Peixoto et al. Abstract graphic



ACS Paragon Plus Environment