Template for Electronic Submission to ACS...

74
Gastrophysics: Statistical thermodynamics of biomolecular denaturation and gelation from the Kirkwood-Buff theory towards the understanding of tofu Seishi Shimizu 1 *, Richard Stenner 1† , Nobuyuki Matubayasi 2,3 1 York Structural Biology Laboratory, Department of Chemistry, University of York, Heslington, York YO10 5DD, United Kingdom 2 Division of Chemical Engineering, Graduate School of Engineering Science, Osaka University, Toyonaka, Osaka 560-8531, Japan 3 Elements Strategy Initiative for Catalysts and Batteries, Kyoto University, Katsura, Kyoto 615-8520, Japan Present address: School of Physics, University of Bristol, Bristol BS8 1TL, United Kingdom. 1 1 2 3 4 5 6 8 9 10 11 12 13 14 15

Transcript of Template for Electronic Submission to ACS...

Gastrophysics: Statistical thermodynamics of

biomolecular denaturation and gelation from the

Kirkwood-Buff theory towards the understanding of

tofu

Seishi Shimizu1*, Richard Stenner1†, Nobuyuki Matubayasi2,3

1York Structural Biology Laboratory, Department of Chemistry, University of York, Heslington,

York YO10 5DD, United Kingdom

2Division of Chemical Engineering, Graduate School of Engineering Science, Osaka University,

Toyonaka, Osaka 560-8531, Japan

3Elements Strategy Initiative for Catalysts and Batteries, Kyoto University, Katsura, Kyoto 615-

8520, Japan

†Present address: School of Physics, University of Bristol, Bristol BS8 1TL, United Kingdom.

KEYWORDS: statistical thermodynamics; Kirkwood-Buff theory; hydration; cosolvents;

gelation; tofu

AUTHOR INFORMATION

1

1

2

3

4

5

7

8

9

10

11

12

13

14

15

16

17

18

Corresponding Author:

Seishi Shimizu

York Structural Biology Laboratory, Department of Chemistry, University of York, Heslington,

York YO10 5DD, United Kingdom

Tel: +44 1904 328281, Fax: +44 1904 328281, Email: [email protected]

ABBREVIATIONS

KB, Kirkwood-Buff.

ABSTRACT

Sugars, alcohols, or salts, when added to food, affects the heat denaturation of proteins and the

sol-gel transition of macromolecules. Such an effect of cosolvents has long been known and

exploited; yet understanding how they work at a molecular level has been a matter of scientific

debate for decades, because of the lack of a definitive theory which can provide a microscopic

explanation. Here we show that a rigorous statistical thermodynamic theory, the Kirkwood-Buff

(KB) theory, provides not only a long-awaited microscopic explanation but also a clear guideline

on how to analyze experimental data. KB theory synthesizes the classical Wyman-Tanford

formula and partial molar volume, and enables the determination of biomolecule-water and

biomolecule-cosolvent interactions solely from experimental data. Nothing beyond the materials

in introductory physical chemistry or chemical thermodynamics textbooks is necessary to follow

the derivations presented in this review.

2

19

20

21

22

23

24

25

26

27

2829

30

31

32

33

34

35

36

37

38

39

40

41

1. Statistical thermodynamics for food science: why necessary?

Our aim is to convince the readers that statistical thermodynamics is indeed a useful tool for

food science. This is especially true, when we try to understand what is really happening at a

molecular scale. Molecular-based understanding is central to food science, because it attempts to

elucidate the texture and taste of food based upon its microscopic behavior, i.e., the structure and

interaction of the constituent molecules (de Man, 1999; Walstra, 2003; Belitz, Grosch, &

Schieberle, 2009). Statistical thermodynamics, then, is indispensable, because it is the only

branch of science which can provide a link between the microscopic and macroscopic worlds

(Hill, 1956; Ben-Naim, 2006).

Applying statistical thermodynamics to complex systems such as food is far from being

straightforward. Most commonly, two strategies have been adopted: (i) computer simulation

(Barker & Grimson, 1989; Euston, Ur-Rehman, & Costello, 2007; Fundo, Quitas, & Silva, 2015)

and (ii) development of simple models (van der Sman, 2016; van der Sman et al., 2013).

Simulations (such as molecular dynamics and Monte Carlo) implement statistical

thermodynamics numerically. Simple model-based approaches are drawn chiefly from the

models of polymers, surfaces, and colloids. The crux of the both approaches lies in the elegance

of approximations, aimed at grasping the essence of molecular structure and interactions out of

the overwhelming complexity of food systems.

In contrast to the above, we take an alternative approach: (iii) rigorous theory as a tool to

extract molecular-level information from thermodynamic data. This approach is distinct from (i)

3

42

43

44

45

46

47

48

49

50

51

52

53

54

55

56

57

58

59

60

61

62

63

64

and (ii) in that certainty, credibility and clarity of interpretation are guaranteed by the rigorous

nature of the theory, because the theory comes directly from the Laws of Physics (Shimizu,

2004; Shimizu & Boon, 2004; Shimizu & Matubayasi, 2014c; Shimizu, 2015; Stenner et al.,

2016; Shimizu & Abbott, 2016).

The aim of this review is to demonstrate how useful statistical thermodynamics is. We will

derive all the necessary formulae from scratch. The derivation is quite straightforward; no

background knowledge is required beyond introductory chemical thermodynamics, such as

Gibbs-Duhem and Clausius-Clapeyron equations (Atkins & de Paula, 2014).

2. Thermodynamics without statistical mechanics is prone to confusion

Our proposal to apply rigorous statistical mechanics to food science does not mean in any way

that we are advocating the abolition of the current thermodynamic and calorimetric approaches.

On the contrary, statistical mechanics fulfils the full potential of thermodynamic analysis, by

bringing in an unprecedented interpretive clarity at a molecular level. (Such a combination of

thermodynamics and statistical mechanics is commonly called statistical thermodynamics; our

standpoint is to pursue food(-related) science within the framework of statistical

thermodynamics.) What have instead been abolished are confusion and ambiguity caused by the

lack of an explicit molecular basis (Shimizu, 2004; Shimizu & Boon, 2004; Shimizu &

Matubayasi, 2014a).

4

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

Question: Consider the addition of extra molecular component(s) − such as sugars, salts,

amino acid derivatives, or macromolecules − to food. Such an addition of cosolvents affects

gelation, solubility, denaturation, and aggregation. How do cosolvents modulate such equilibria?

(Note that such extra components are referred to in many different names, such as cosolvents,

cosolutes, additives, or solutes; “cosolvents” will be used throughout this paper.)

A thermodynamic answer: Consider a transition α → β of the solute (referred to as species u

), such as sol →gel, solute in pure phase →solute dissolved in solvent, folded→unfolded, or

monomers→aggregate, and the accompanying the standard Gibbs free energy Δ μu¿. (Throughout

this paper, Δ signifies the change that accompanies a transition α → β.) The addition of

cosolvents (species 2) into water (species 1¿ changes the water activity, and therefore the

chemical potential of water μ1. How Δ μu¿ changes with μ1 can be expressed as a competition

between the change in number of water and cosolvent molecules bound to the biomolecules,

Δ Nu 1 and Δ Nu 2 (Figure 1) that accompany the transition:

−( ∂Δ μu¿

∂ μ1 )T , P , nu →0=Δ Nu 1−

n1

n2Δ Nu 2 (1)

where n1 and n2 are bulk concentrations of water and cosolvent (Wyman, 1948; 1964; Casassa &

Eisenberg, 1964; Tanford, 1968, 1969, 1970; Schellman, 1987; Timasheff, 1998, 2002a, 2002b;

Parsegian, Rand & Rau, 1995, 2000). For example, Δ Nu 1 in the context of protein unfolding

signifies Δ Nui=N uiu −Nui

f , which corresponds to the difference in the number of bound water

between the unfolded state (u) and the folded state (f).

5

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

106

107

108

Eq. (1) is commonly referred to as the Wyman-Tanford formula, whose interpretation owes to

the seminal contributions by Wyman, Eisenberg, Tanford, Schellman, Timasheff, and Parsegian

(Wyman, 1948; 1964; Casassa & Eisenberg, 1964; Tanford, 1968, 1969, 1970; Schellman, 1987;

Timasheff, 1998, 2002a, 2002b; Parsegian, Rand & Rau, 1995, 2000). Eq. (1) can readily be,

and has indeed been, applied to interpret thermodynamic data (Baier, Decker & McClements,

2004; Miyawaki & Tatsuno, 2010; Miyawaki, Dozen & Nomura, 2013; Miyawaki, Omote &

Matsuhira, 2016). All one should do is to plot Δ μu¿ against μ1 (or equivalently against RT ln a1

where a1 is the water activity (Atkins & de Paula, 2014)) in order to obtain Δ Nu 1−n1

n2Δ Nu 2

(Figure 2). The food science applications of this formula include the effects of sugars, salts and

alcohols on the thermal denaturation of proteins, as well as on gelation (Baier, Decker &

McClements, 2004; Miyawaki & Tatsuno, 2010; Miyawaki, Dozen & Nomura, 2013; Miyawaki,

Omote & Matsuhira, 2016).

Further simplification: Which is the dominant contribution to the equilibrium shift, the

change of water binding (Δ Nu1) or cosolvent binding (Δ Nu 2)? The Wyman-Tanford formula

(Eq. (1)) on its own does not provide a definitive answer to this question. Yet dialysis

measurements show that sugars, polyols, and “kosmotropic” salts are preferentially excluded

from biomolecular surfaces (Figure 1(b)) (Timasheff, 1998, 2002a, 2002b). Consequently they

are not bound to biomolecules; hence the change of the bound number is zero, i.e., Δ Nu 2=0

(Parsegian, Rand & Rau, 1995, 2000). This renders Eq. (1) a powerful tool by simplifying it to

−( ∂Δ μu¿

∂ μ1 )T , P , nu →0=Δ Nu 1 (2)

6

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

The change in number of bound water molecules Δ Nu 1, which accompany folding, gelation,

solubilisation, aggregation, can therefore be measured directly from the Wyman-Tanford plot

(Figure 2) (Parsegian, Rand & Rau, 1995, 2000).

Controversy: Eq. (2) was the focus of intense controversy, because of the unrealistically large

numbers of water molecules have been estimated to be released from protein-ligand binding and

allosteric transitions (Timasheff, 1998, 2002a). In the course of controversy, the following

doubts have been raised about Eq. (1) (Timasheff, 1998, 2002a; Parsegian, Rand & Rau, 2000):

Controversy 1: Whether Δ Nu 1 and Δ Nu 2 really have a precise microscopic meaning.

Controversy 2: Whether Δ Nu1 and Δ Nu 2 are independently determinable in principle.

Controversy 3: Whether preferential exclusion really signify the lack of binding.

To appreciate the depth and extent of the above controversy, one should appreciate that the

cosolvent effect on biomolecular equilibria (i.e., ( ∂ Δ μu¿

∂ μ1 )T , P ,nu → 0) have been interpreted from a

number of different “perspectives” (Parsegian, 2000; Timasheff, 1998, 2002a), such as

1. Preferential solvation theory (Timasheff, 1998; 2002a) and the solvent exchange model

(Schellman, 1987), i.e., Eq. (1).

2. Osmotic stress method (Parsegian, 1995, 2000), i.e., Eq. (2).

3. Macromolecular crowding, which attributes the origin of preferential exclusion of

cosolvents to their large molecular size (Zhou et al., 2008; Davis-Searles et al., 2001;

Sapir and Harries, 2016), i.e., Eq. (1) with the assumed dominance of −n1

n2Δ N u2.

7

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

151

4. The “making”/“breaking” of “water structure” by cosolvents, which enhances/weakens

the hydrogen bond network of water around the hydrophobic group, thereby leading to

the cosolvent-induced protein stabilization/denaturation (Frank & Franks, 1968), i.e., Eq.

(1) with the assumed dominance of Δ Nu 1.

The interrelationship between these different perspectives was the deeper historical reason for

the above controversy. In addition,

5. Partial molar volume measurements can also probe biomolecular hydration (Chalikian,

2003), whose estimations are very different from the osmotic stress method (Timasheff,

1998, 2002a)

What is the relationship between all these different perspectives?

The way out. Thermodynamics, by definition, cannot give any definitive answer to any of the

above questions, because thermodynamics itself cannot give any microscopic interpretation of

thermodynamic quantities. Only statistical thermodynamics (i.e., the combination of

thermodynamics and statistical mechanics which can link macroscopic measurements to

intermolecular interactions) can provide an answer (Shimizu, 2004; Shimizu & Boon, 2004;

Shimizu & Matubayasi, 2014a). This will be explained from scratch in Section 3

3. Clarification comes from statistical thermodynamics

Here we re-derive the Wyman-Tanford formula from statistical thermodynamics. Such a re-

derivation clarifies what Δ Nu 1 and Δ Nu 2 really mean at a molecular level without any room for

ambiguity. This molecular-level clarification has indeed resolved the controversy which arose

8

152

153

154

155

156

157

158

159

160

161

162

163

164

165

166

167

168

169

170

171

172

173

174

from the lack of a molecular basis (Shimizu, 2004; Shimizu & Boon, 2004; Shimizu &

Matubayasi, 2014a).

To introduce the statistical thermodynamic theory of the cosolvent effect, there are three

possible approaches. The first is the pioneering matrix inversion approach of Ben-Naim (1977,

2006), generalized later by Smith (2008), which rendered the classical Kirkwood-Buff theory

(Kirkwood & Buff, 1951) a powerful and usable tool in solution chemistry (Shimizu, 2004;

Shulgin & Ruckenstein, 2006; Pierce et al., 2008; Harries, & Rösgen, 2008; Ploetz & Smith,

2013). This approach, however, is mathematically lengthy. The second is our direct derivation

from the first principles of statistical thermodynamics, which is much simpler yet requires some

familiarity with the statistical thermodynamics (Shimizu & Matubayasi, 2014b). The third is a

thermodynamic derivation pioneered by Hall (1971), which has been employed by one of us to

resolve the osmotic stress controversy (Shimizu, 2004; Shimizu & Boon, 2004). This approach

has the advantage of requiring only the introductory knowledge found in most undergraduate

textbooks (Newman, 1994; Shimizu, 2004). One minor disadvantage of this approach, however,

is that there is a small gap in derivation, which can be understood intuitively but must be

accepted without proof. Yet the first, second, and third approaches derive the same equations,

hence we adopt the third for its intuitive appeal (Shimizu, 2004; Shimizu, Booth & Abbott, 2013;

Shimizu & Matubayasi, 2014a, 2014b).

From here, let us focus on one conformational state (either folded or unfolded state of a

protein, or sol or gel state of biomolecular assemblies). The changes signified by Δ will be

introduced later. Consider a three component solution consisting of solute (i=u), water (i=1),

9

175

176

177

178

179

180

181

182

183

184

185

186

187

188

189

190

191

192

193

194

195

196

197

and cosolvent (i=2) molecules. For the purpose of deriving the Wyman-Tanford equation, here

we consider a solute molecule at infinite dilution, which means we ignore solute-solute

interactions. Extension to concentrated solutes is straightforward for two component systems,

and is extremely cumbersome in three component systems. The theory is applicable without any

modification to any principal solvent (i=1) (Hall, 1971; Newman, 1994; Shimizu, 2004; Shimizu

et al., 2013; Shimizu & Matubayasi, 2014a, 2014b); i = 1 does not need to be water (Shimizu &

Abbott, 2016).

Our central interest here is the effect of cosolvents on the solvation of a solute molecule. To

focus exclusively on the interaction between the solute and water, we conceptually fix the centre-

of-mass position of the solute molecule, which means that we adopt Ben-Naim’s standard state

(Ben-Naim, 2006). We then divide the solution into two parts: the first part (called the “solute’s

vicinity”) contains a fixed solute molecule, the other part (called the “bulk”) is far away from the

solute (Figure 3) (Hall, 1971; Shimizu, 2004; Shimizu et al., 2013; Shimizu & Matubayasi,

2014a, 2014b).

Simply put, what we want to describe here is how the solute molecule changes the

concentration of the solvent species around it. If the cosolvent interacts favourably with the

solute, increased cosolvent concentration is observed in solute’s vicinity compared to the bulk

(Figure 1a). If, on the other hand, the cosolvent interacts unfavourably to the solute, decreased

cosolvent concentration follows compared to the bulk phase (Figure 1b). Thus the key of the KB

theory, in a nutshell, is the concentration difference between solute’s vicinity and the bulk

10

198

199

200

201

202

203

204

205

206

207

208

209

210

211

212

213

214

215

216

217

218

219

solution (Hall, 1971; Newman, 1994; Shimizu, 2004; Shimizu et al., 2013; Shimizu &

Matubayasi, 2014a, 2014b).

Our goal, therefore, is to express theoretically the vicinity-bulk concentration difference. To

achieve this, the Gibbs–Duhem equations from introductory chemical thermodynamics textbooks

will be demonstrated to be powerful. Under the constant temperature (i.e., dT=0) the Gibbs-

Duhem equation for the bulk solution is

N 1 d μ1+N2 d μ2=V dP (3)

where N i and μi respectively express the number and chemical potential of species i, V is the

volume, P is the pressure. Likewise, the Gibbs-Duhem equation for the vicinity is

N1v d μ1+N 2

v d μ2+N uv d μu

¿=V v dP (4)

where “v” indicates the number and volume of the vicinity, and “¿” indicates that the solute’s

centre of mass is fixed. Note that μi and P are the same in the local and bulk systems due to

phase equilibrium condition (Hall, 1971; Newman, 1994; Shimizu, 2004; Shimizu et al., 2013;

Shimizu & Matubayasi, 2014a, 2014b).

Since the local-bulk concentration difference is our goal, we introduce the vicinity and bulk

concentrations, niv=N i

v /V v and ni=N i /V , which transform Eqs. (3) and (4) respectively into

n1 d μ1+n2d μ2=dP (5)

n1v d μ1+n2

v d μ2+nuv d μu

¿=dP (6)

Note the existence of V v in the definition of niv for the introduction of vicinity (local)

concentration. The Gibbs-Duhem equations can now be expressed in terms of the vicinity-bulk

concentration difference, niv−ni, by subtracting Eq. (5) from Eq. (6):

11

220

221

222

223

224

225

226

227

228

229

230

231

232

233

234

235

236

237

238

239

240

241

242

(n1v−n1) d μ1+(n2

v−n2) d μ2+nuv d μu

¿=0 (7)

Now we introduce the key concept, excess solvation numbers, N ui, defined as the vicinity-bulk

concentration change per solute molecule

Nui=ni

v−ni

nuv (8)

with the help of which Eq. (7) can be rewritten as

−d μu¿=Nu1 d μ1+Nu 2 d μ2 (9)

Eq. (9) is the fundamental relationship in the Kirkwood-Buff theory, which links the hydration

free energy to the excess solvation numbers (Hall, 1971; Newman, 1994; Shimizu, 2004;

Shimizu et al., 2013; Shimizu & Matubayasi, 2014a, 2014b). Here, it is important to note the

value of N u 1 and N u2 does not depend on the choice of V v as long as it is sufficiently large (Hall,

1971).

Now the Wyman-Tanford formula can be obtained straightforwardly from Eq. (9). Firstly,

differentiating Eq. (9) with respect to μ1 yields

−( ∂μu¿

∂μ1 )T , P,nu → 0=Nu 1+N u2( ∂ μ2

∂ μ1 )T , P , nu → 0(10)

Combining Eq. (10) with ( ∂ μ2

∂ μ1)T , P , nu →0

=−n1

n2, which comes straightforwardly from Eq. (5) under

the isobaric condition dP=0, we finally obtain

−( ∂μu¿

∂μ1 )T , P,nu → 0=Nu 1−

n1

n2Nu 2 (11)

which is essentially the Wyman-Tanford formula. More precisely, Eq. (11) is for one

conformational state of the biomolecule; Eq. (1), which refers to the change α → β, can

12

243

244

245

246

247

248

249

250

251

252

253

254

255

256

257

258

259

260

261

262

straightforwardly be derived by subtracting Eq. (11) for the state α from that for β (Hall, 1971;

Shimizu, 2004; Shimizu et al., 2013; Shimizu & Matubayasi, 2014a, 2014b). Eq. (11) can also be

derived from other approaches (Smith, 2004, 2008; Shulgin & Ruckenstein, 2006; Pearce et al.,

2008).

So far we have derived everything based purely on chemical thermodynamics through the use

of the Gibbs-Duhem equations. The same results can be obtained statistical thermodynamically.

The only, yet crucial, additional insight that statistical mechanics can bring is the rigorous

microscopic definition of the excess solvation numbers in terms of the radial distribution

function,gij (r ) (Figure 4),

Nui=4 π ni∫0

[ gij (r )−1 ] r2 dr (12)

where r is the distance between the centers of the molecular species i and j. (Note that gij (r ) can

be measurable experimentally by X-ray and neutron scattering experiments.)

Now we show that the resolution to Controversy 1 in Section 2 comes from Eq. (12). As

shown in Figure 4, N u 1 and Nu 2 have been defined microscopically in terms of the radial

distribution functions (Shimizu, 2004; Shimizu & Boon, 2004; Shimizu & Matubayasi, 2014a);

N ui represents a net excess of the molecules of species i around the solute molecules compared to

the bulk solution. Figure 4 shows that there are positive and negative contributions to N ui (Figure

4, “Contact”). The attraction between the solute and the species i gives rise to a peak of gij (r ),

which contributes positively to N ui, which has been appreciated by the previous models. What

was not clearly realised before our KB-based theory was the negative contribution to Nui arising

13

263

264

265

266

267

268

269

270

271

272

273

274

275

276

277

278

279

280

281

282

283

284

from the excluded volume effect, namely the contribution from the impossible molecular overlap

configurations (Figure 4, “Excluded volume”). Since such configurations with small distance r

between solute and the species i do not exist, gij (r )=0; such r contributes as density depletion

compared to the bulk, and contributes negatively to N ui. Such a negative contribution forced a

reconsideration of the interpretation of the osmotic effect, which will be discussed in the

subsequent sections.

This clarification leads also to the resolution to Controversies 2 and 3, which will be explained

in Section 4.

4. Statistical thermodynamics (the Kirkwood-Buff theory) completes and fulfils the

Wyman-Tanford formula

What is the point of re-deriving the classical Wyman-Tanford formula yet again in a different

way as in Section 3? The re-derivation unlocks a full interpretative potential at a microscopic

level, which is a marked departure from the age-old assumptions, limitations and consequent

confusions that had long plagued the Wyman-Tanford formula (Shimizu, 2004; Shimizu &

Boon, 2004; Shimizu & Matubayasi, 2014a).

First we show that Nu 1 and N u2 can both be determined independently from experimental data,

thereby resolving Controversy 2 in Section 2. To determine the two unknowns (Nu 1 and N u 2),

two independent equations are necessary; we need an additional relationship that supplements

the Wyman-Tanford formula. Fortunately, the partial molar volume V u can fit the bill

14

285

286

287

288

289

290

291

292

293

294

295

296

297

298

299

300

301

302

303

304

305

306

307

V u ≡( ∂ μu

∂ P )N1 , N2 , T , nu→ 0

=−V 1 Nu 1−V 2 Nu 2+RT κT (13)

where κT is the isothermal compressibility, which only makes a negligible contribution. N u 1 and

N u2 can be determined by solving the simultaneous equations (Eq. (11) and (13)). (Here, μu is the

chemical potential of the solute, whereas μu¿ is the chemical potential for a solute whose centre of

mass is fixed in position; they are related by d μu=d μu¿+RT d ln nu (Ben-Naim, 2006; Shimizu &

Matubayasi, 2014b)). Likewise, the changes accompanying the conformational change α → β,

namely Δ Nu 1 and Δ Nu 2, can be determined by supplementing the Wyman-Tanford formula (Eq.

(1)) with the following relationship for the partial molar volume change ΔV u arising directly

from Eq. (13):

ΔV u=−V 1 Δ Nu 1−V 2 Δ N u2 (14)

Both Δ Nu 1 and Δ Nu 2 are determinable by solving the simultaneous equations consisting of Eqs.

(1) and (14). This has been made possible only by the help of statistical thermodynamics

(Shimizu, 2004; Shimizu & Boon, 2004; Shimizu et al., 2006; Shimizu & Matubayasi, 2014a).

This has led to an understanding as to why the osmotic stress method had grossly overestimated

hydration by neglecting the exclusion of osmolytes from biomolecular surfaces. For the details of

this resolution, see Shimizu (2004) and Shimizu & Matubayasi (2014a).

Thus the simultaneous equations (Eqs. (1) and (14)) not only enabled a determination of both

Δ Nu 1 and Δ Nu 2 but also a synthesis of the osmotic (the Tanford-Wyman formula) and

volumetric measurements, which had hitherto been conducted separately and independently.

What we propose here is to carry out the Wyman-Tanford and volumetric experiments

simultaneously to obtain a full picture, the simultaneous determination of biomolecule-water and

15

308

309

310

311

312

313

314

315

316

317

318

319

320

321

322

323

324

325

326

327

328

329

biomolecule-cosolvent interactions (Gekko, 1989; Chanasattru, Decker & McClements, 2008).

This will further develop fruitful applications of volume and density measurements in food

science.

Statistical thermodynamics thus brought a molecular-based clarification to the Wyman-

Tanford formula, the foundation of the cosolvent effect on food biomolecular folding,

aggregation, and gelation. What remains to be explained is the statistical thermodynamic

resolution to Controversy 3, which can be best demonstrated through the analysis of

experimental data in Section 7.

5. The Wyman-Tanford formula can be simplified when cosolvents are dilute

With an additional insight from the KB theory, the Wyman-Tanford formula can be simplified

further. Instead of the excess numbers (Eq. (12)), it is more convenient to use the following KB

integral Gui (not to be confused with the Gibbs free energy!)

Gui=4π∫0

[ gij (r )−1 ] r2 dr (15)

which measures the interaction per bulk density. Using Eq. (15), Eqs. (1) and (14) can be

rewritten as

−( ∂Δ μu¿

∂ μ1 )T , P , nu →0=n1 ( ΔGu1−ΔGu2 ) (16)

ΔV u=−n1 V 1 ΔGu1−n2V 2 ΔGu 2 (17)

since from Eqs. (12) and (14) the relation ni ΔGui=Δ Nui can be obtained. Eq. (16) has a clearer

meaning: the cosolvent effect on Δ μu¿ is due to the competition between solute-water and solute-

16

330

331

332

333

334

335

336

337

338

339

340

341

342

343

344

345

346

347

348

349

350

351

cosolvent interactions expressed in terms of the KB integrals (Shimizu, 2004; Shimizu & Boon,

2004; Shimizu & Matubayasi, 2014a).

When the cosolvent concentration is dilute, Eqs. (16) and (17) can be simplified even further.

Leaving a more rigorous derivation to the literature (Shimizu et al., 2013), we present here an

intuitive approach. When the cosolvent is dilute, Raoult’s Law, d μ1≃d ( RT ln x1 ) (in its

differential form), holds true, where x1 is the mole fraction of water. Since x1=n1

n1+n2≃ 1

1+n2/n1,

ln x1≃−n2

n10 , where n1

0 is the concentration (density) of pure water. Using this, Eq. (16) can be

simplified at n2 →0 as follows:

1RT ( ∂ Δ μu

¿

∂ n2 )T , P , nu→ 0 ;n2 → 0=ΔGu 1−ΔGu2 (18)

Eq. (17) can be simplified also by noting n10= 1

V 10 (Shimizu & Boon, 2004; Shimizu &

Matubayasi, 2014c):

( Δ V u )n2 →0=−ΔGu1 (19)

To summarise, when the cosolvents are dilute, we can use Eqs. (18) and (19) to facilitate the

analysis of the experimental data significantly (Shimizu, 2004; Shimizu & Boon, 2004; Shimizu

& Matubayasi, 2014a).

6. The Clausius-Clapeyron equations fill the remaining theory-experiment gap

17

352

353

354

355

356

357

358

359

360

361

362

363

364

365

366

367

368

369

370

371

We have thus established a rigorous theory and a method to analyse experimental data to

reveal how biomolecule-water and biomolecule-cosolvent interactions competitively contribute

to the cosolvent-induced changes on protein denaturation and aggregation, solubilisation, and

gelation. The general theory (applicable to any cosolvent concentration) requires ( ∂ Δ μu¿

∂ μ1 )T , P ,nu → 0

and ΔV u (Eqs. (1) and (14)), and the special theory valid for dilute cosolvents (n2 → 0) requires

( ∂ Δ μu¿

∂n2 )T , P,nu → 0 ;n2 →0 and ( Δ V u )n2 →0 (Eqs. (18) and (19)). The goal here is to link these quantities

to the convention of measurements undertaken in food chemistry.

Let us take the gel → sol transition as an example (see Appendix A for why the theory can be

used for gel → sol transition); the theory below can straightforwardly be applied to other

transitions. What is important here is that as long as the thermodynamic quantities relating to a

transition is available, we can calculate the KB integrals. Let us summarise the conclusions first,

before justification. What is readily measurable is the gel →sol transition temperature (T g → s),

and how it is dependent on cosolvent concentration (n2), from which ( ∂ Δ μu

∂ μ1)T , P ,nu → 0

and

( ∂ Δ μu

∂n2)T , P,nu → 0 ;n2 →0

can be calculated by the following formulae:

( ∂ Δ μu

∂ μ1)T , P ,nu → 0

=Δ Sg→ s

δ T g → s

δ μ1(20)

( ∂ Δ μu

∂ n2)T , P ,nu → 0 ;n2 →0

=ΔSg→ s

δ T g→ s

δ n2

(21)

18

372

373

374

375

376

377

378

379

380

381

382

383

384

385

386

387

388

389

where Δ Sg → s is the entropy change which accompanies gel → sol transition, which can be

obtained directly from calorimetry by Δ Sg → s=Δ H g → s /T g →s, where Δ H g → s refers to the

enthalpy change. Likewise, the volume change ΔV g → s can either be obtained directly from

dilatometry, or by the hydrostatic pressure P dependence of T g → s in the following manner

(Gekko & Koga, 1983; Gekko & Kasuya, 1985; Gekko et al., 1987; Shimizu, 2011b):

( ∂ Δ μu¿

∂ P )T , N1 ;n2 →0

=( ΔV g → s )n2 →0=ΔSg→ sδ T g→ s

δP(22)

Deriving Eqs. (20) -(22) is straightforward. In the following, the outline of derivation is

presented, while a more detailed discussion can be found in Appendix B. Let us focus here on

Eqs. (20) and (22), because Eq. (21) has already been derived in our previous paper (Shimizu &

Matubayasi, 2014c). Suppose that, along the sol-gel coexistence, the gel→sol transition

temperature rises from T g → s to T g → s+δ T g → s, when the chemical potential increases from μ1 to

μ1+δ μ1 or when the pressure is changed from P to P+δP. (Here we have used T g→ s for

temperature only to emphasise that we are dealing with the transition temperature.) Along the

sol-gel coexistence,

μu( s)=μu

(g ) (23)

holds true. Note that μu(α )( = s or g) in this equation is the total chemical potential, not the

pseudochemical potential, so that the superscript * is not attached. It is remarked furthermore

that μu( g )is the chemical potential per biopolymer in the gel state, not per aggregate particle. Using

Eq. (B4) for the changes along the sol-gel coexistence, δ T g→ s, δ μ1 and δP, we have

0=d μu(s )−d μu

( g)=−ΔSg→ s δ T g → s+( ∂ ∆ μu

∂ μ1)T=T g→ s, P

δ μ1+ΔV g→ s δP (24)

Hence δP=0 leads to Eq. (20). Under n2 = 0, Eq. (22) follows (Shimizu & Matubayasi, 2014c).

19

390

391

392

393

394

395

396

397

398

399

400

401

402

403

404

405

406

407

408

409

410

Thus we have made it possible to analyze transition temperature measurements in terms of the

KB theory.

7. The Kirkwood-Buff theory can readily be applied to food science: tofu as an example

Here we demonstrate that the KB theory can indeed be applied to food systems quite

straightforwardly. (For other, already published examples, see: Shimizu & Matubayasi, 2014c;

Shimizu, 2015; Stenner et al., 2016). As an example for this review, we have chosen tofu, which

is an important class of food gels, which has not been analyzed previously through the use of

statistical thermodynamics. The gap in the available experimental data, which will be revealed

along the process of data analysis, will hopefully be filled, in order to achieve a full, molecular-

based understanding of food gelation.

7.1 Heat denaturation of soy protein isolates in the presence of salts

Heat-induced denaturation of soy proteins is the first step in tofu-making, prior to gelation. This

process has been known to be strongly affected by the presence of salts (Kohyama & Nishinari,

1993; Banerjee & Bhattacharya, 2012; Nishinari et al., 2014). How at a molecular level do salts

affect heat denaturation? Here we demonstrate that a straightforward analysis of the experimental

data, which stems from the KB theory, can tell a correct mechanism from other.

20

411

412

413

414

415

416

417

418

419

420

421

422

423

424

425

426

427

428

429

430

431

432

Why can some salts enhance the thermal denaturation of soy protein whereas the others can

prevent it? There have been two different hypotheses proposed so far

(i) Hydration/water structure changes hypothesis stems from the classical view that the

hydrophobic effect is the driving force of protein folding and that the hydrophobic effect is

caused by the enhanced hydrogen bonding network of water (called the “iceberg”) around the

hydrophobic solute (Frank & Evans, 1945). Cosolvents either promote (“structure makers”) or

interrupt (“structure breakers”) the iceberg indirectly (Frank & Franks, 1968). This hypothesis

has been used to account for i) the effects of various anions (SO42-, Cl-, Br-, SCN-) on the gelling

and rheological behaviour of soy proteins (Babajimopoulos, Damodaran, Rizvi & Kinsella,

1983), and ii) the interactions between soy proteins and cations, such as Ca2+ (Kohyama, Sano &

Doi, 1995; Puppo & Anon, 1998; Sakakibara & Noguchi, 1977; Zhao, Li, Qin & Chen, 2016)

and Na+ (Anon, Lamballerie & Speroni, 2011; Puppo & Anon, 1998).

(ii) Preferential solvation hypothesis, namely the competition between protein-water and

protein-cosolvent interactions. If protein-cosolvent interaction is stronger than protein-water

(Figure 1(a)), then, as we increase cosolvent concentration, equilibrium shifts towards the

increased exposure of protein surface, to which cosolvents can be bound; this will promote

protein denaturation. On the contrary, if protein-cosolvent interaction is unfavourable compared

to protein-water (Figure 1(b)), the increase in cosolvent concentration shift the equilibrium

towards compactness. This hypothesis has been employed to account the effects of Ca2+ on soy

protein isolates (Canabady-Rochelle, Sanchez, Mellema & Banon, 2009; Scilingo & Anon, 1996;

Yuan, Velev, Chen, Campbell, Kaler & Lenhoff, 2002).

21

433

434

435

436

437

438

439

440

441

442

443

444

445

446

447

448

449

450

451

452

453

454

455

Which of the above two hypotheses is correct? The KB theory can not only answer this

question but also eliminate the ambiguity through the individual determination of protein-water

and protein-cosolvent interactions.

The KB-analysis of the protein denaturation data is straightforward. We consider the transition

from the native state (denoted as n) to the denatured state (d). Here we focus on the dilute salt

limit; extension to higher salt concentrations can be done (which requires more thermodynamic

data – See section 7.2). At this limit, all we need are: δ T n →d

δ n2 (how the denaturation temperature

T n →d depends on cosolvent concentration n2), δ T n →d

δ P (dependence of T n→d on pressure P), and

Δ Sn→ d (the entropy change that accompanies denaturation at zero cosolvent concentration).

Combining Eqs. (18), (19), (21) and (22), a straightforward algebra yields

ΔGu 1=−Δ Sn → d

δ Tn → d

δP=−( ΔV n→d )n2 →0 (25)

ΔGu 2=ΔGu 1−Δ Hn → d

R T n→ d2

δ T n →d

δ n2(26)

In carrying out the data analysis based on Eqs. (25) and (26), care must be taken for a correct

unit conversion. Table 1 shows the data analysis all the way from the raw experimental data to

the KB integrals. What is crucial to the calculation of ΔGu 2 is the large magnitude of the gradient

δ T n→d

δ n2, which causes T n →d at low salt concentrations (15 mM, or even at 5mM) significantly

different from those in the absence of salts, which are large Note that ( Δ V n →d )n2 →0 for soy

proteins have not been measured experimentally to the best of our knowledge. We therefore have

22

456

457

458

459

460

461

462

463

464

465

466

467

468

469

470

471

472

473

474

475

to employ the empirical formula to estimate ( Δ V n →d )n2 →0 (Chalikian & Filfil, 2003)). The

empirical formula by Chalikian and Filfil (2003) may seem upon first inspection to contain many

parameters. However, thanks to their extensive correlational analysis using experimental

volumetric and structural data, the required input parameters have been reduced to two:

molecular weight of the protein, and the degree of unfolding. The molecular weight data used in

the analysis have been summarized in Tables 1 and 2. There is no other way than to assume the

degree of unfolding as being 1 (fully unfolded), which may be debatable. Fortunately, the

contribution from ( Δ V n →d )n2 →0 is estimated to be much smaller in magnitude than

Δ Hn → d

R T n→ d2

δ T n →d

δ n2 (Shimizu, 2011a); so the error in ( Δ V n →d )n2 →0 does not affect the analysis. The

resultant ΔGu 1 and ΔGu 2 are able to judge whether the above hypotheses are valid.

The water structure hypothesis, in the language of Kirkwood-Buff theory, translates to the

dominance of ΔGu 1over ΔGu 2(|∆Gu 1|≫|∆ Gu 2|) i.e. change in the water structure/hydration is

the driving force behind protein denaturation. However, for salt-soy protein systems (Table 1),

the opposite is true: the changes in hydration/water structure indeed contribute negligibly to the

denaturation free energy change (|∆Gu 1|≪|∆ Gu 2|), hence cannot account for the salt-induced

changes in the thermal stability. Even though this conclusion has been drawn, relying upon the

estimated ∆ Gu 1, it is consistent with other proteins (Shimizu et al., 2006; Shimizu, 2011a;

Shimizu & Matubayasi, 2014a).

The preferential solvation hypothesis is consistent with the KB theory, which provides a

rigorous link between the competitive solvation ΔGu 2−ΔGu 1 to the cosolvent-induced

23

476

477

478

479

480

481

482

483

484

485

486

487

488

489

490

491

492

493

494

495

496

497

equilibrium changes. This conclusion was reached purely theoretically. Nevertheless, what is

important and insightful here is that the KB analysis can clarify what “preferential solvation”

really means. The concept of preferential solvation is inseparable from the Wyman-Tanford

formula (Eq (1)), in which Δ Nu2 and Δ Nu 1 (or ΔGu2 and ΔGu 1) are not independent; hence only

the difference, Δ Nu 1−n1

n2Δ N u2 (or ΔGu 2−ΔGu 1) could be extracted from experimental data

(Timasheff, 1998, 2002a). The KB theory, on the contrary, has enabled the indivisual calculation

of ΔGu2 and ΔGu 1 (Shimizu, 2004; Shimizu & Boon, 2004; Shimizu & Matubayasi, 2014a). For

salts (Table 1), the dominance of ΔGu 2 over ΔGu 1 (i.e. |∆Gu2|≫|∆ Gu 1|) shows that the widely

varying magnitude of the protein-salt interaction is the cause for the strongly cosolvent-

dependent thermal denaturation temperature changes. The large, negative ΔGu 2 shows in

conjunction with Eq. (15) that the salts are more excluded from the denatured state than from the

native state (Shimizu & Boon, 2004; Shimizu et al., 2006; Shimizu 2011a; Shimizu &

Matubayasi, 2014a).

To summarise, a simple KB-based analysis can not only show that water structure hypothesis

cannot account for the cosolvent effect on thermal denaturation but also identify that protein-

cosolvent interaction is the true cause. The latter identification could not be done by the Wyman-

Tanford formula alone, which could not separate proein-cosolvent interaction from protein-water

interaction; the separation between the two could be achieved only by the KB theory (Shimizu,

2004; Shimizu & Boon, 2004; Shimizu & Matubayasi, 2014a). The effect of salts at higher

concentrations (Kohyama & Nishinari, 1993; Banerjee & Bhattacharya, 2012; Nishinari et al.,

2014) can be analyzed using Eqs. (1) and (14), instead of Eqs. (18) and (19).

24

498

499

500

501

502

503

504

505

506

507

508

509

510

511

512

513

514

515

516

517

518

519

520

7.2. Effects of polyols on the thermal denaturation of soy proteins

Here we turn to the glycerol and ethylene glycol, which have been known to act weakly on

protein stability. The reason why we turn to such model cosolvents (which are not used for

setting tofu) is because of the important insights into the mechanism of cosolvent-induced

change of gelation, as has been revealed by Gekko et al. (1999) through the study of such

cosolvents.

Sorbitol and glycerol have been known to stabilize the native structures of proteins. Their

stabilization effect is indeed weaker than MgCl2 and CaCl2 from Section 7.1, as well as than

sucrose, trehalose and polyethelene glycol (Shimizu & Boon, 2004; Shimizu, 2011a; Shimizu &

Matubayasi, 2014a). Table 2 shows that sorbitol and glycerol indeed acts as weak stabilizers for

glycinin and β-conglycinin (Gekko et al., 1999).

Even for weak stabilizers like sorbitol and glycerol, the water structure dominance hypothesis (

|∆Gu2|≪|∆ Gu 1|) does not apply (Table 2). Yet the relationship |∆Gu 2|≫|∆ Gu1|, valid for many

stronger osmolytes and salts (Shimizu, 2011a; Shimizu & Matubayasi, 2014a), does not hold true

for sorbitol and glycerol. Indeed, ∆ Gu 2 for glycerol and sorbitol are much smaller in magnitude

than salts, even in the same order of magnitude as ∆ Gu 1, suggesting that ∆ Gu 1 and ∆ Gu 2 both

contribute competitively to the cosolvent-induced change of denaturation. The exclusion of

sorbitol and glycerol from protein surfaces is therefore only very mild.

25

521

522

523

524

525

526

527

528

529

530

531

532

533

534

535

536

537

538

539

540

541

542

The exclusion of ethylene glycol from the protein surface is even weaker, as can be shown by

its much smaller magnitude of ΔGu2 (Table 2). Even though ethylene glycol is generally known

as weak osmolyte, it acts as a weak denaturant for soy proteins, lowering the denaturation

temperature (Gekko et al., 1999). This is because ΔGu 2 is so small in magnitude that ΔGu 1 wins

over (Table 2).

Based upon the appreciation of a weak exclusion of glycerol and an even weaker exclusion of

ethylene glycol, let us move onto how these cosolvents act on tofu gelation.

7.3. Effects of polyols on the gelation of tofu

Setting tofu, namely the gelation of soy bean proteins, requires the addition of gelling agents.

Three common gelling agents are MgCl2, CaSO4 and Glucono-δ-lactone (Kohyama & Nishinari,

1993; Banerjee & Bhattacharya, 2012; Nishinari et al., 2014) Unfortunately, to the best of our

knowledge, systematic experimental data on tofu gelation in the presence of these cosolvents,

which can readily be applied to the analysis based upon the KB theory, have not been found in

the literature. Instead, the most systematic thermodynamic study has been done by Gekko and

coworkers, who have measured the effect of model cosolvents, i.e., polyols, on the gelation of

soy protein extract, together with the pressure dependence (Gekko, 1993; Gekko, Li, & Makino,

1999). Hence, in this section, we study the effect of polyols on tofu gelation, in order to

illustrate the applicability of our theory.

26

543

544

545

546

547

548

549

550

551

552

553

554

555

556

557

558

559

560

561

562

563

564

565

What can our statistical thermodynamic analysis say about the mechanism on the enhancement

of tofu gelation by cosolvents? Three different hypotheses have been proposed: (i) the

enhancement of the water structure around the protein, and the concurrent change in protein

hydration, induced by the cosolvent (Babajimopoulos et al., 1983); (ii) exclusion of cosolvents

from protein surfaces in the sol phase (Canabady-Rochelle et al., 2009; Scilingo & Anon, 1996;

Yuan et al., 2002); (iii) binding between cosolvents and proteins in the gel phase (Arii &

Takenaka, 2013, 2014). Which of the above hypotheses is/are likely to be correct? To answer

this question, we shall again analyze δ T g→ s

δ n2 ,

δ T g→ s

δ P and Δ Sg → s obtained from experiments,

which are not limited to very dilute concentrations of cosolvents. This is why we have to solve

the simultaneous equations (Eqs. (1) and (14)), in combination to Eq. (20), we obtain:

ΔGu 1=−n2V 2

V 1Δ Sg→ s

δ T g → s

δ μ1−ΔV u

(27)

ΔGu 2=V 1 Δ Sg → s

δ T g→ s

δ μ1−ΔV u (28)

where n1 V 1+n2V 2=1 at nu →0 has been used in the derivation. Now we evaluate both KB

integrals. The key quantity δ T g→ s

δ μ1 is calculated from the Wyman-Tanford plot, which is shown

in Figure 5. Density data at T g → s has been used to calculate n1, n2, V 1 and V 2. (Gelation data in

the presence of sorbitol could not be analysed because of the lack of density and osmotic data).

ΔV u was available only at n2=0, and assumed to remain constant irrespective of n2; the small

magnitude of ΔV u makes this a good approximation.

27

566

567

568

569

570

571

572

573

574

575

576

577

578

579

580

581

582

583

584

585

Figure 6 shows that for both glycerol and ethylene glycol the dominant contribution is ΔGu 2,

and that the hydration change ΔGu 1 negligibly small relative to ΔGu 2. This shows straightway

that Water structure hypothesis (i) is unlikely, because of the negligibly small contribution

from the hydration change ΔGu 1. But more importantly, |∆Gu2|≫|∆ Gu 1| for gelation is in stark

contrast to how the same cosolvents acted on soy protein denaturation in Section 7.2; glycerol

was weakly excluded from soy proteins, and even weaker exclusion was observed for ethylene

glycol. What can this stark difference between denaturation and gelation teach us about how

glycerol and ethylene glycol interact with sol and gel?

To answer this question, let us start by understanding the origin of the large, positive and

dominant contribution is ΔGu2 for glycerol-tofu (Figure 6), which is defined as

ΔGu 2=G u2(s)−Gu2

(g ) (29)

where Gu2(g) and Gu2

(s) respectively represent the KB integrals for gel and sol states (Stenner et al.,

2016). Note that sol state consists of denatured soy proteins. According to Eq. (29), for ΔGu 2 to

be large and positive, Gu 2(s)≫Gu 2

(g). A strong attraction between glycerol and the denatured soy

protein would make Gu2(s) more positive. But this is not the case; we already know from Table 2

that glycerol is excluded weakly from protein surfaces. Hence the only way that Gu 2(s)≫Gu2

(g) can

take place is to make Gu2(g) larger negative. This takes place when glycerol is excluded more from

the gel phase than from the sol phase.

The large, negative and dominant contribution ΔGu 2 for ethylene glycol-tofu can be attributed

likewise to Gu 2(s)≪Gu 2

(g). This can be achieved by (1) a very strong exclusion of ethylene glycol

28

586

587

588

589

590

591

592

593

594

595

596

597

598

599

600

601

602

603

604

605

606

607

from the denatured (sol) proteins or (2) the binding of ethylene glycol to the gel. We know that

(1) is in contradiction to our observation in Section 7.2 that ethylene glycol is excluded only very

weakly from proteins. Hence ethylene glycol should bind the gel phase more strongly than the

sol phase.

Taken all together, the KB theory reveals that ethylene glycol, which is the smaller cosolvent,

is embedded into the gel phase, whereas glycerol, which is larger, tends to be excluded from the

gel network.

Thus gel melting temperature data were sufficient to judge the validity of the classical

hypotheses, when analyzed in the framework of statistical thermodynamic theory. Statistical

thermodynamics can therefore make a significant contribution to food science through its power

of definitive clarification at a molecular scale. Towards this goal, systematic soy protein gelation

data in the presence of the common gelling agents are badly needed.

7.4. What the Kirkwood-Buff theory cannot do in its present form

Analysing the thermodynamic data on protein denaturation and gel melting was sufficient to

dispel some of the age-long hypotheses on the mechanism on the cosolvent effect on protein

denaturation and gelation. This was enabled because the KB theory is rigorous, which stems

directly from the basic principles of statistical thermodynamics (Kirkwood & Buff, 1951; Hall,

1971; Ben-Naim, 1977; Newman, 1994; Shimizu, 2004; Shimizu & Matubayasi, 2014a).

29

608

609

610

611

612

613

614

615

616

617

618

619

620

621

622

623

624

625

626

627

628

629

630

Note, however, that the basis of interpretation, the KB integrals, are statistical averages that refer

to the overall deviation of the radial distribution from their bulk values. Hence it is often difficult

to identify the specific atomic groups or protein residues that are responsible for cosolvent-

protein binding or the exclusion of cosolvents (Shimizu, 2004; Shimizu & Boon, 2004; Shimizu,

2011a, 2011b). For example, the KB theory has identified the binding (or embedment) of polyols

as the important driving force of gelation. We expect (subject to the availability of experimental

data) that the same would be the case for salts; yet the KB theory on its own cannot comment, for

example, on the electrostatic effect of the bound ions on crosslinks. To address such questions at

an atomistic basis, carrying out simulations would be necessary. Simulations, on the other hand,

cannot have the ability of giving the definitive answer as the KB theory has due to the model

approximations involved (force field, system size, and statistical procedures); yet the accuracy of

simulations can be validated by a comparison to the KB integrals obtained directly from

experiments (Pierce et al., 2008; Ploetz & Smith, 2013). Hence combining models and the

rigorous theory would prove fruitful in revealing the mechanism of food processes at atomistic

detail.

8. Conclusion

Statistical thermodynamics is useful in understanding how cosolvents affect protein heat

denaturation, biomolecular gelation, and aggregation. When analyzing thermodynamic data,

statistical thermodynamics provides an unambiguous method to calculate both the biomolecule-

water and biomolecule-cosolvent interactions. To do so, all one needs to know is how transition

temperature depends on water activity (or cosolvent density) and pressure, together with the bulk

30

631

632

633

634

635

636

637

638

639

640

641

642

643

644

645

646

647

648

649

650

651

652

653

density data of water-cosolvent mixture (Shimizu, 2004; Shimizu & Boon, 2004; Shulgin &

Ruckenstein, 2006; Pierce et al., 2008; Harries, & Rösgen, 2008; Ploetz & Smith, 2013; Shimizu

& Matubayasi, 2014a).

Before the statistical thermodynamic refurbishment, the study of the cosolvent effects was

plagued with confusion and controversies. This is because thermodynamics on its own cannot

provide an unambiguous link between thermodynamic data and molecular behaviour. Statistical

thermodynamics therefore provides a clear interpretive tool, fruitful for the analysis of

thermodynamic data on food (Shimizu, 2004; Shimizu & Boon, 2004; Shimizu & Matubayasi,

2014a).

What statistical thermodynamics has revealed is somewhat different from the previous

thermodynamic expectations. The classical view, that cosolvents induce changes in biomolecular

hydration, which drives the shift in denaturation, gelation, and aggregation, is not supported by

statistical thermodynamics. Biomolecule-cosolvent interaction (whether attractive or net

exclusive) is instead the dominant driving force (Shimizu & Matubayasi, 2014c; Shimizu, 2015;

Stenner et al., 2016).

Statistical thermodynamics thus provides a useful guideline, clearly showing what kind of

measurements should be undertaken for understanding the cosolvent action at a molecular level.

An attempt in this review to apply the KB theory to soy proteins revealed that there still are

crucial gaps in the available thermodynamic data, which are waiting to be filled.

31

654

655

656

657

658

659

660

661

662

663

664

665

666

667

668

669

670

671

672

673

674

675

676

ACKNOWLEDGMENTS

We thank Katsuyoshi Nishinari and Steven Abbott for stimulating discussions, Kaja Harton for a

careful reading of the manuscript, and Tom Nicol for technical assistance. This work was

supported by the Grants-in-Aid for Scientific Research (Nos. 15K13550 and 26240045) from the

Japan Society for the Promotion of Science, and by the Elements Strategy Initiative for Catalysts

and Batteries from the Ministry of Education, Culture, Sports, Science, and Technology, and by

Computational Materials Science Initiative, Theoretical and Computational Chemistry Initiative,

and HPCI System Research Project (Project IDs: hp160013, hp160019, and hp160214) of the

Next-Generation Supercomputing Project.

32

677

678

679

680

681

682

683

684

685

686

Appendix A

Here we discuss further how to treat the gel state within the framework of biomolecules in

dilution, following Shimizu & Matubayasi (2014c). The crux of the matter is that the protein

molecules can make linked network structure in the gel state, yet are dilute enough so that

protein-protein interactions other than the formation of link can be ignored. Such an approach

has been adopted, albeit implicitly, in the thermodynamic study on the effect solvation on

binding and aggregation (Shimizu & Matubayasi, 2014c). Hence we consider dilute biopolymer

aggregate (of aggregation number m) as the gel state, for which (Eq. (11)) can be rewritten as

−( ∂( μu¿ ( g ,aggregate )

m )∂ μ1

)T , P, nu→ 0

=n1(Gu1( g , aggregate )

m−

Gu 2( g , aggregate )

m ) (A1)

where μu¿(g , aggregate) is the pseudochemical potential of the aggregate of the biopolymers at the gel

state. This quantity, introduced per aggregate particle, can be converted to the chemical potential

per biopolymer through the division by the aggregation number.

Here we show that Eq. (B1) can be simplified further in terms of the KB integral per

biopolymer. To do so, we have to exploit the following: (1) The biopolymer-biopolymer

interactions except for the small region in which the interaction takes place can be ignored; (2)

The biopolymer is much larger in size than the water and cosolvent molecules. It follows from

(1) and (2) that, since biopolymers are large and dilute, we can safely assume that each water and

cosolvent molecules, which are not in the bulk, interact only with its nearest biopolymer.

33

687

688

689

690

691

692

693

694

695

696

697

698

699

700

701

702

703

704

705

706

Let us translate this into the language of the radial distribution function (RDF). The RDF

between the gel aggregate g and the species i at the position r can be expressed as

gui(g ,aggregate )(r )−1=∑

κ[ gκi

(g )(r )−1 ] (A2)

where gκi(g)(r ) is RDF between the κ th biopolymer in the gel and the species i. The basis of Eq.

(A2) is gκi( g ) (r )=1 for all the biopolymers except for the one nearest to the solvent. This means

that the gel aggregate-solvent KB integral

Gui(g ,aggregate)=∑

κGκi

(g)=mGui( g)

(A3)

can be expressed in terms of the KB integral per biopolymer in the gel state, Gui(g). We have thus

generalized the Wyman-Tanford formula for sol-gel transition.

Appendix B

Here we present a detailed derivation of Eqs. (20), (21) from Eq. (24), and of Eq. (22).

Derivation of Eq. (24). Using basic calculus, the total differential d μu(α) at the state α can be

expressed in terms of the change of the variables (μ1 ,T , P) in the following manner:

d μu(α )=( ∂ μu

( α )

∂ μ1)

T , Pd μ1+( ∂μu

(α )

∂T )μ1 , P

dT+( ∂ μu(α )

∂ P )T , μ1

dP (B1)

where the last two partial differentials have the physical interpretations respectively as −Su(α ) and

V u( α ), namely the partial molar entropy and volume of u at the state α . Hence for the sol (α=s)

and gel (α=g) states, Eq. (B1) can be rewritten as

d μu( s)=( ∂ μu

(s )

∂ μ1)

T , Pd μ1−Su

( s) dT+V u(s)dP (B2)

34

707

708

709

710

711

712

713

714

715

716

717

718

719

720

721

722

723

724

725

726

d μu( g)=( ∂ μu

(g )

∂ μ1)

T , Pd μ1−Su

( g ) dT +V u(g)dP (B3)

Subtracting Eq. (B3) from Eq. (B2) yields

d μu( s)−d μu

( g )=( ∂∆ μu

∂ μ1)T , P

d μ1−ΔSg→ s dT+ ΔV g → s dP (B4)

in which ∆ μu=μu( s)−μu

( g ) has been introduced. Eq. (24) can be obtained as the special case of Eq.

(B4), i.e., d μu( s)=d μu

( g ) because of phase equilibrium (Eq. (23)), and around the transition

temperature, T=T s → g.

Derivation of Eq. (20). Here we consider how sol→gel equilibrium is affected by μ1 when P

is kept constant, as has been required by the Wyman-Tanford formula. Hence we consider Eq.

(24) under the isobaric condition, dP=0. This leads to

0=−Δ Sg → sd T g → s+( ∂ ∆ μu

∂ μ1)T =Tg →s , P

d μ1 (B5)

from which Eq. (20) can be derived straightforwardly through algebraic manipulation. Note that

we have used δ T g→ s and δ μ1 instead of d T g → s and d μ1, in order to emphasise that experimental

values have been used to calculate δ T g → s

δ μ1.

Derivation of Eq. (21). Eq. (21) is a special case of Eq. (20) at the dilute cosolvent

concentration limit, n2 → 0. Just as in the derivation of Eq. (18), we use Raoult’s Law, which

yields

d μ1≃d ( RT ln x1 )=−1n1

0 d n2 (B6)

35

727

728

729

730

731

732

733

734

735

736

737

738

739

740

741

742

743

744

745

Combining Eq. (B6) with the l.h.s. of Eq. (20), one obtains,

( ∂ Δ μu

∂ μ1)T , P,nu → 0 ;n2 →0

=−n10( ∂ Δ μu

∂ n2)T , P,nu → 0 ;n2 →0

(B7)

and with the r.h.s. of Eq. (20), one obtains,

Δ Sg → s

δ T g → s

δ μ1=−n1

0 Δ Sg→ s

δ T g → s

δ n2(B8)

Combining Eqs. (20), (B7) and (B8) yields Eq. (21).

Derivation of Eq. (22). Now we consider the change of pressure when the cosolvent is absent.

Going back to the total differential d μu(α ) at the state α , note that the variables here are (T , P),

because there is no change in the solvent composition. Hence Eq. (B1) under this condition

should be rewritten as

d μu(α )=( ∂ μu

( α )

∂T )μ1 , P

dT+( ∂ μu(α )

∂ P )T , μ1

dP (B9)

Repeating the same argument, one arrive at the following two-component equivalent of Eq. (24):

0=−Δ Sg→ sd T g → s+ΔV g→s dP (B10)

Eq. (22) can be derived straightforwardly from Eq. (B10).

36

746

747

748

749

750

751

752

753

754

755

756

757

758

759

760

761

762

REFERENCES

Atkins, P., & de Paula, J. (2014). Atkins’ Physical Chemistry. (10th ed.). Oxford University

Press: Oxford. pp 680 – 682.

Añón, M. C., de Lamballerie, M., & Speroni, F. (2011). Influence of NaCl concentration and

high pressure treatment on thermal denaturation of soybean proteins. Innovative Food Science

and Emerging Technologies, 12, 443–450.

Arii, Y., & Takenaka, Y. (2014). Magnesium chloride concentration-dependent formation of

tofu-like precipitates with different physicochemical properties. Bioscience, Biotechnology, and

Biochemistry, 77, 928-933.

Arii, Y., & Takenaka, Y. (2014). Initiation of protein association in tofu formation by metal

ions. Bioscience, Biotechnology, and Biochemistry, 78, 86–91.

Azfal, W., Mohammadi, A. H., & Richon, D. (2009). Solutions from (273.15 to 363.15) K:

Experimental measurements and correlations. Journal of Chemical and Engineering Data, 54,

1254–1261.

Baier, S. K., Decker, E. A., & McClements, D. J. (2004). Impact of glycerol on thermostability

and heat-induced gelation of bovine serum albumin. Food Hydrocolloids, 18, 91–100.

Babajimopoulos, M., Damodaran, S., Rizvi, S. S. H., & Kinsella, J. E. (1983). Effects of

various anions on the rheological and gelling behavior of soy proteins: Thermodynamic

observations. Journal of Agricultural and Food Chemistry, 31, 1270-1275.

37

763

764

765

766

767

768

769

770

771

772

773

774

775

776

777

778

779

780

781

Banerjee, S., & Bhattacharya, S. (2012). Food gels: Gelling process and new applications.

Critical Reviews in Food Science and Nutrition, 52, 334–346.

Barker, G. C., & Grimson, M. J. (1989). Food colloid science and the art of computer

simulation. Food Hydrocolloids, 3, 345-363.

Belitz, H.-D., Grosch, W., & Schieberle, P. (2009). Food Chemistry. Springer, Berlin.

Ben-Naim, A. (1977). Inversion of the Kirkwood–Buff theory of solutions: Application to the

water–ethanol system. Journal of Chemical Physics, 67, 4884-4890.

Ben-Naim, A. (2006). Molecular Theory of Solutions. Oxford University Press, New York.

Booth, J. J., Abbott, S., & Shimizu, S. (2012). Mechanism of hydrophobic drug solubilization

by small molecule hydrotropes. Journal of Physical Chemistry B, 116, 14915-14921.

Booth, J. J., Omar, M., Abbott, S., & Shimizu, S. (2016). Hydrotrope accumulation around the

drug: The driving force for solubilization and minimum hydrotrope concentration for

nicotinamide and urea. Physical Chemistry Chemical Physics, 17, 8028-8037.

Canabady-Rochelle, L. S., Sanchez, C., Mellema, M., & Banon, S. (2009). Study of calcium-

soy protein interactions by isothermal titration calorimetry and pH cycle. Journal of Agricultural

and Food Chemistry, 57, 5939-5947.

Casassa, E. F., & Eisenberg, H. (1964). Thermodynamic analysis of multicomponent solutions.

Advances in Protein Chemistry, 19, 287-394.

Chalikian T. V. (2003). Volumetric properties of proteins. Annual Review of Biophysics and

Biomolecular Structure, 32, 207-235.

38

782

783

784

785

786

787

788

789

790

791

792

793

794

795

796

797

798

799

800

801

Chalikian, T.V., and Filfil, R. (2003). How large are the volume changes accompanying

protein transitions and binding? Biophysical Chemistry,104, 489-499.

Chanasattru, W., Decker, E. A., & McClements, D. J. (2008). Impact of cosolvents (polyols)

on globular protein functionality: Ultrasonic velocity, density, surface tension and solubility

study. Food Hydrocolloids, 22,1475–1484.

Davis-Searles, P.R., Saunders, A.J., Erie, D.A., Winzor, D.J., & Pielak, G.J. (2001).

Interpreting the effects of small uncharged solutes on protein-folding equilibria. Annuual

Reviews of Biophysics and Biomolecular Structucture, 30, 271-306.

de Man, J. (1999). Principles of Food Chemistry. Aspen Publications, Gaithersburg, Maryland.

Euston, S. R, Ur-Rehman, S., & Costello, G. (2007). Denaturation and aggregation of β-

lactoglobulin—a preliminary molecular dynamics study. Food Hydrocolloids, 21, 1081–1091.

Frank, H. S., & Evans, M. W. (1945). Free volume and entropy in condensed systems III.

Entropy in binary liquid mixtures; partial molal entropy in dilute solutions; structure and

thermodynamics in aqueous electrolytes. Journal of Chemical Physics, 13, 507-532.

Frank. H. S., & Franks, F. (1968). Structural approach to the solvent power of water for

hydrocarbons; Urea as a structure breaker. Journal of Chemical Physics, 48, 4746-4757.

Fundo, J. F., Quintas, M. A. F., & Silva, C. L. M. (2015). Molecular dynamics and structure in

physical properties and stability of food systems. Food Engineering Reviews, 7, 384-392.

Gekko K. (1993). The sol-gel transition of food macromolecules under high pressure. In Food

Hydrocolloids: Structures, Properties, and Functions. 259-264.

39

802

803

804

805

806

807

808

809

810

811

812

813

814

815

816

817

818

819

820

821

Gekko, K., & Kasuya, K. (1985). Effect of pressure on the sol-gel transition of carrageenans.

International Journal of Biological Macromolecules,7, 299-306

Gekko, K., & Koga, S. (1983). Increased thermal stability of collagen in the presence of sugars

and polyols. Journal of Biochemistry, 94, 199-205.

Gekko, K. (1989). Hydration—structure—function relationships of polysaccharides and

proteins. Food Hydrocolloids, 3, 289–299.

Gekko, K., Li, X., & Makino, S. (1999). Competing effect of polyols on the thermal stability

and gelation of soy protein. Bioscience, Biotechnology and Biochemistry, 63, 2208-2211.

Gekko, K., Mugishima, H., & Koga, S. (1987). Effects of sugars and polyols on the sol-gel

transition of κ-carrageenan: Calorimetric study. International Journal of Biological

Macromolecles, 9, 146-152.

Glycerine Producers' Association (1963). Physical properties of glycerine and its solutions.

New York: Glycerine Producers' Association.

Hall, D.G. (1971). Kirkwood-Buff theory of solutions. An alternative derivation of part of it

and some applications. Transactions of the Faraday Society, 67, 2516-2524.

Harries, D., & Rösgen, J. (2008). A practical guide on how osmolytes modulate

macromolecular properties. Methods in Cell Biology, 84, 679–735.

Hill, T.L. (1956) Statistical Mechanics: Principles and Selected Applications. McGraw-Hill,

New York.

40

822

823

824

825

826

827

828

829

830

831

832

833

834

835

836

837

838

839

840

Horstmann, S., Gardeler, H., Wilken, M., Fischer, K., & Gmehling, J. (2009). Isothermal

vapor-liquid equilibrium and excess enthalpy data for the binary systems water + 1,2-ethanediol

and propene + acetophenone. Journal of Chemical and Engineering Data 49, 1508-1511.

Kauzmann, W. (1959). Some factors in the interpretation of protein denaturation. Advances in

Protein Chemistry, 14, 1-63

Kirkwood, J. G., & Buff, F. P. (1951). The statistical mechanical theory of solution. Journal of

Chemical Physics, 19, 774-777.

Kohyama, K., & Nishinari, K. (1993). Rheological studies on the gelation process of soybean

75 and 11s proteins in the presence of glucono-δ-lactone. Journal of Agricultural and Food

Chemistry, 41, 8-14.

Kohyama, K., Sano, Y., & Doi, E. (1995). Rheological characteristics and gelation mechanism

of tofu (soybean curd). Journal of Agricultural and Food Chemistry, 43, 1808-1812.

Liu, C., Teng, Z., Yang, X.-Q., Tang, C.-H., Jiang, Y., Qu, J.-N., Qi, J.-R., Zhu, J.-H. (2010).

Characterisation of soybean glycinin and β-conglycinin fractionated by using MgCl2 instead of

CaCl2. International Journal of Food Science and Technology, 45, 155-162.

Miyawaki, O., & Tatsuno, M. (2010). Thermodynamic analysis of alcohol effect on thermal

stability of proteins. Journal of Bioscience and Bioengineering, 111, 198-203.

Miyawaki, O., Dozen, M., & Nomura, K. (2013). Thermodynamic analysis of osmolyte effect

on thermal stability of ribonuclease A in terms of water activity. Biophysical chemistry,185, 19-

24.

41

841

842

843

844

845

846

847

848

849

850

851

852

853

854

855

856

857

858

859

860

Miyawaki, O., Omote, C., & Matsuhira, K. (2015). Thermodynamic analysis of sol–gel

transition of gelatin in terms of water activity in various solutions. Biopolymers, 103, 685–691.

Newman, K. E. (1994). Kirkwood–Buff solution theory: derivation and applications Chemical

Society Reviews, 23, 31-40.

Nishinari, K., & Fang, K. (2016). Sucrose release from polysaccharide gels. Food & Function,

in press.

Nishinari, K., Fang, K., Guo, S., & Phillips, G.O. (2014). Soy proteins: A review on

composition, aggregation and emulsification. Food Hydrocolloids, 39, 301–318.

Parsegian, V.A., Rand, R.P., & Rau, D.C. (1995). Macromolecules and water: Probing with

osmotic stress. Methods in Enzymology, 259, 43-94.

Parsegian, V.A., Rand, R.P., & Rau, D.C. (2000). Osmotic stress, crowding, preferential

hydration, and binding: A comparison of perspectives. Proceedings of the National Academy of

Sciences of U.S.A., 97, 3987–3992.

Pierce, P., Kang, M., Aburi, M., Weerasinghe, S., & Smith, P. E. (2008). Recent applications

of Kirkwood–Buff theory to biological systems. Cell Biochemistry and Biophysics, 50, 1-22.

Ploetz, E. A., & Smith, P. E. (2013). Local fluctuations in solution: Theory and applications.

Advances in Chemical Physics, 153, 311-372.

Sakakibara, M., & Noguchi, H. (1977). Interaction of l1S fraction of soybean protein with

calcium ion. Agricultural and Biological Chernistry, 41, 1575-1580.

42

861

862

863

864

865

866

867

868

869

870

871

872

873

874

875

876

877

878

879

Sapir, L., & Harries, D. (2016). Macromolecular compaction by mixed solutions: Bridging

versus depletion attraction. Current Opinion in Colloid & Interface Science, 22, 80–87.

Schellman, J.A. (1987). Selective binding and solvent denaturation. Biopolymers, 26, 549-559.

Scilingo, A. A., & Anon, M. C. (1996). Calorimetric study of soybean protein isolates: Effect

of calcium and thermal treatments. Journal of Agricultural and Food Chemistry, 44, 3751−3756.

Shimizu, S. (2004). Estimating hydration changes upon biomolecular reactions from osmotic

stress, high pressure, and preferential hydration experiments. Proceedings of the National

Academy of Sciences of U.S.A., 101, 1195-1199.

Shimizu, S. (2011a). Molecular origin of the cosolvent-induced changes in the thermal stability

of proteins. Chemical Physics Letters, 514, 156-158.

Shimizu, S. (2011b). The effect of urea on hydrophobic hydration: preferential interaction and

the enthalpy of transfer. Chemical Physics Letters, 517, 76-79.

Shimizu, S. (2015). Caffeine dimerization: Effects of sugar, salts, and water structure. Food &

Function, 6, 3228-3235.

Shimizu, S., & Abbott, S. (2016). How entrainers enhance solubility in supercritical carbon

dioxide. Journal of Physical Chemistry, 120, 3713–3723.

Shimizu, S., & Boon, C. L. (2004). The Kirkwood-Buff theory and the effect of cosolvents on

biochemical reactions. Journal of Chemical Physics, 121, 9147-9155.

Shimizu, S., Booth, J.J., & Abbott, S. (2013). Hydrotropy: Binding models vs. statistical

thermodynamics. Physical Chemistry Chemical Physics, 15, 20625-20632.

43

880

881

882

883

884

885

886

887

888

889

890

891

892

893

894

895

896

897

898

899

Shimizu, S., & Matubayasi, N. (2014a) Preferential hydration: Dividing surface vs excess

numbers. Journal of Physical Chemistry B, 118, 3922-3930.

Shimizu, S., & Matubayasi, N. (2014b). Hydrotropy: Monomer-micelle equilibrium and

minimum hydrotrope concentration. Journal of Physical Chemistry B 118, 10515–10524.

Shimizu, S., & Matubayasi, N. (2014c) Gelation: The role of sugars and polyols on gelatin and

agarose. Journal of Physical Chemistry B, 118, 13210-13216.

Shimizu, S., McLaren, W.M., & Matubayasi, N. (2006). The Hofmeister series and protein-salt

interactions. Journal of Chemical Physics, 124, 234905/1-234905/5.

Shulgin, I. L., & Ruckenstein, E. (2006). A Protein Molecule in a Mixed Solvent: The

Preferential Binding Parameter via the Kirkwood-Buff Theory. Biophysical Journal, 90, 704–

707.

Smith, P. E. (2004). Cosolvent interactions with biomolecules:  Relating computer simulation

data to experimental thermodynamic data. Journal of Physical Chemistry B, 108, 18716–18724.

Smith, P. E. (2008). On the Kirkwood–Buff inversion procedure. Journal of Chemical Physics,

129, 124509/1 - 124509/5.

Stenner, R., Matubayasi, N., & Shimizu, S. (2016). Gelation of carrageenan: Effects of sugars

and polyols. Food Hydrocolloids, 54, 284-292.

Tanford, C. (1968). Protein denaturation. Advances in Protein Chemistry, 23, 121-282.

Tanford, C. (1969). Extension of theory of linked functions to incorporate effects of protein

hydration. Journal of Molecular Biology, 39, 539-544.

44

900

901

902

903

904

905

906

907

908

909

910

911

912

913

914

915

916

917

918

919

Tanford, C. (1970). Protein denaturation. Part C: Theoretical models for the mechanism of

denaturation. Advances in Protein Chemistry, 24, 2-95.

Timasheff, S. N. (1998). In disperse solution, “osmotic stress” is a restricted case of

preferential interactions. Proceedings of the National Academy of Sciences, USA, 95, 7363-7367.

Timasheff, S. N. (2002a). Protein-solvent preferential interactions, protein hydration, and the

modulation of biochemical reactions by solvent components. Proceedings of the National

Academy of Sciences, 99, 9721-9726.

Timasheff, S. N. (2002b) Protein hydration, thermodynamic binding and preferential

hydration. Biochemistry, 41, 13473-13482.

van der Sman, R. G. M. (2016). Sugar and polyol solutions as effective solvent for biopolymer

Food Hydrocolloids, 56, 144–149.

van der Sman, R. G. M., Paudel, E., Voda A., & Khalloufi, S. (2013). Hydration properties of

vegetable foods explained by Flory–Rehner theory. Food Research International, 54, 804–811.

Walstra, P. (2003). Physical Chemistry of Foods. Marcel Dekker, New York.

Watase, M., Kohyama, K., & Nishinari, K. (1992). Effects of sugars and polyols on the gel-sol

transition of agarose by differential scanning calorimetry. Thermochimica Acta, 206 163-173.

Wyman, J. (1948). Heme proteins. Advances in Protein Chemistry, 4, 407-531.

Wyman, J. (1964). Linked functions and reciprocal effects in hemoglobin – A second look.

Advances in Protein Chemistry, 19, 223-286.

45

920

921

922

923

924

925

926

927

928

929

930

931

932

933

934

935

936

937

938

Yuan, Y. J., Velev, O. D., Chen, K., Campbell, B. E., Kaler, E. W., & Lenhoff, A. M. (2002).

Effect of pH and Ca2+-induced associations of soybean proteins. Journal of Agricultural and

Food Chemistry, 50, 4953-4958.

Zhang, L., Zeng, M. (2008). Proteins as source of Materials. In Monomers, Polymers and

Composites from Renewable Resources, edited by Mohamed Naceur Belgacem, Alessandro

Gandini. Elsevier, Oxford.

Zhao, H., Li, W., Qin F., & Chen, J. (2016). Calcium sulphate-induced soya bean protein tofu-

type gels: Influence of denaturation and particle size. International Journal of Food Science and

Technology, 51, 731–741.

Zhou, H.X., Rivas, G., & Minton, A.P. (2008). Macromolecular crowding and confinement:

Biochemical, biophysical, and potential physiological consequences. Annual Reviews of

Biophysics, 37, 375-397.

46

939

940

941

942

943

944

945

946

947

948

949

950

951

952

Figure and Table Captions

Table 1: The effect of salts on the thermal denaturation of soy proteins analyzed with the Kirkwood-Buff theory. aExperimental data taken from Liu et al. (2010); bEstimated using the empirical formula by Chalikian and Filfil (2003), in which 360 and 190 kDa (Zhang & Zeng, 2008) have been used as the upper-limit molecular weights of glycinin and β-conglycinin, respectively; cCalculated at the low concentration limit using the data at 0 and 5mM salt concentrations.

Table 2: The effect of polyols (ethylene glycol (EG), glycerol and sorbitol) on the thermal denaturation of soy proteins analyzed with the Kirkwood-Buff theory. aExperimental data taken from Gekko et al. (1999); bEstimated using the empirical formula by Chalikian and Filfil (2003), in which 360 and 190 kDa (Zhang & Zeng, 2008) have been used as the upper-limit molecular weights of glycinin and β-conglycinin, respectively.

Figure 1: Schematic representations of preferential interactions: (a) preferential interaction of cosolvents (dark red) with the biomolecule (black) compared to that of water (blue) with the biomolecule; (b) preferential water-biomolecule interaction over water-cosolvent interaction.

Figure 2: Wyman-Tanford analysis of the cosolvent effect. See Eq. (1).

Figure 3: Intuitive derivation of the Kirkwood-Buff theory. The first step is to divide the solution into the “local” (vicinity of the solute) and “bulk” (far away from the solute). The goal is to evaluate the net excess/deficiency of water (blue) and cosolvent (dark red) molecules in the “local” region relative to the bulk.

Figure 4: The microscopic interpretation of the excess cosolvent number, Nu2 (Eq. (12)). “Excluded volume”: Due to steric repulsion (excluded volume effect) the distance between the solute (black) and the cosolvent (dark red) cannot be shorter than the contact distance; this, in the language of the radial distribution function, corresponds to g (r )=0. Such distance ranges contribute negatively to the excess number due to the depletion of cosolvent density. “Contact”: At the distance of solute-cosolvent contact, an attractive interaction would increase the occurrence of such a configuration, making g(r ) peak higher, thereby contributing positively to the excess number. “Bulk”: As r becomes larger, the cosolvent density approaches to that of the bulk.

Figure 5. Wyman-Tanford plot for the gel → sol transition of the soy protein isolate in the presence of glycerol (black) and ethylene glycol (red). The experimental data (circle and square) are taken from Gekko et al. (1999). The source of experimental data used for analysis are the following: density of ethylene glycol-water mixture (Azfal et al., 2009); vapor pressure of ethylene glycol-water mixture (Horstmann et al., 2009); density and vapor pressure of glycerol-water mixture (Glycerin producer’s association, 1963). Lines representing fitting are: glycerol (T g → s=−0.0019 ( RT ln a1 )2−0.0185 ( RT ln a1 )+350.72) and ethylene glycol (

47

953954955956

957958959960

961962

963964965

966967968

969

970971972973

974975976977978979980981982

983984985986987988989

T g → s=95.17

(1+exp[ RT lna1−118.987.7 ])). These fitting lines are used for differentiation (Eq. (21)).

Multiplying Δ Sg → s=¿5.5 J K-1 (mole crosslinks)-1 (Gekko, 1999) can convert this figure to a Wyman-Tanford plot.

Figure 6. Kirkwood-Buff integrals for the gel → sol transition of soy bean isolate protein in the presence of glycerol and ethylene glycol, calculated using Figure 5 and the KB theory. Note that the KB integrals have been calculated approximately at 353 K for glycerol and ethylene glycol. Contribution from ΔGu 1 has been assumed to be constant (which is indeed negligible for larger n2 for glycerol and smaller n2 for ethylene glycol) for which the pure water value from Gekko (1993) have been used in the absence of data in the presence of cosolvents. The KB integral changes are expressed in the units of cm3 per mole of cross links.

48

990

991992

993994995996997998999

1000