pmc logo imageJournal ListSearchpmc logo image
Logo of biophysjBiophysical Journal - SubscribeBiophysical Journal - SubmissionBiophysical Journal - Current IssueBiophysical SocietyBiophysical J
Biophys J. 2006 January 15; 90(2): 704–707.
Published online 2005 November 4. doi: 10.1529/biophysj.105.074112.
PMCID: PMC1367075
A Protein Molecule in a Mixed Solvent: The Preferential Binding Parameter via the Kirkwood-Buff Theory
Ivan L. Shulgin and Eli Ruckenstein
Department of Chemical & Biological Engineering State University of New York at Buffalo Amherst, New York
Address reprint requests to Eli Ruckenstein, Tel.: 716-645-2911, ext. 2214; Fax: 716-645-3822; E-mail: feaeliru/at/acsu.buffalo.edu.
Received September 7, 2005; Accepted October 12, 2005.
 
In a recent article (Schurr, J. M., D. P. Rangel, and S. R. Aragon. 2005. A contribution to the theory of preferential interaction coefficients. Biophys. J. 89:2258–2276), a detailed derivation of an expression for the preferential binding coefficient via the Kirkwood-Buff theory of solutions was presented. The authors of this Comment (Shulgin, I. L., and E. Ruckenstein. 2005. A protein molecule in an aqueous mixed solvent: fluctuation theory outlook. J. Chem. Phys. 123:054909) also recently established on the basis of the Kirkwood-Buff theory of solutions an equation for the preferential binding of a cosolvent to a protein. There are other publications that relate the preferential binding parameter to the Kirkwood-Buff theory of solutions for protein + binary mixed solvents. The expressions derived in the two articles mentioned above are different because the definitions of the preferential binding parameter are different. However, there are articles in which the definitions of the preferential binding parameter are the same, but the derived equations that relate the preferential binding parameter to the Kirkwood-Buff integrals are different. The goal of this Comment is to examine the various expressions that relate the preferential binding parameter to the Kirkwood-Buff theory.
INTRODUCTION

An important characteristic of a solution of a protein (component 2) in a mixture water (1) + cosolvent (3) is the preferential binding parameter equation M1 (16)

equation M2
(1)
where equation M3 is the molality of component i, P is the pressure, T is the absolute temperature, and equation M4 is the chemical potential of component i. The preferential binding parameter can be also defined at a molarity scale by
equation M5
(2)
where equation M6 is the molar concentration of component i. It should be emphasized that equation M7 and equation M8 are defined at infinite protein dilution.

The preferential binding parameter equation M9 was determined experimentally (57) and provides information regarding the interactions between a protein and the components of the mixed solvent. As a rule (15), equation M10 the protein is preferentially hydrated, for cosolvents such as glycerol, sucrose, etc., which can stabilize at high concentrations the protein structure and preserve its enzymatic activity (35), and equation M11 the protein is preferentially solvated by cosolvents (such as urea), which can cause protein denaturation.

In literature (8) a number of different definitions of the preferential binding parameter (coefficient) have been employed. They can be connected by thermodynamic relations for ternary mixtures (8). In this Comment the preferential binding parameter will be mostly defined by Eqs. 1 and 2.

Because the preferential binding parameter is a meaningful physical quantity, attempts have been made to relate it to a general theory of solutions, such as the Kirkwood-Buff theory of solutions (9). Several authors reported results in this direction (1017). The authors of this Comment derived the following equation for equation M12 (16):

equation M13
(3)
where equation M14 and equation M15 are the Kirkwood-Buff integrals defined as (9)
equation M16
(4)
where equation M17 is the radial distribution function between species α and β, and r is the distance between the centers of molecules α and β.

Equation 3 differs from the expression of equation M18 employed in Shimizu (10,11):

equation M19
(5)

In a recent article in this journal (17), the Kirkwood-Buff theory of solutions was used to express the preferential binding coefficient equation M20 defined as

equation M21
(6)
in terms of the Kirkwood-Buff integrals. It was found (17) that
equation M22
(7)

As noted in Schurr et al. (17) the preferential binding coefficient equation M23 defined by Eq. 6 differs from the preferential binding parameter equation M24 defined by Eq. 2.

However, Eqs. 3 and 5 are different equations even though they are based on the same definition of the preferential binding parameter and have the same theoretical basis: the Kirkwood-Buff theory of solutions. To make a selection between Eqs. 3 and 5 a simple limiting case, the ideal ternary mixture, will be examined using the traditional thermodynamics, and the results will be compared to those provided by Eqs. 3 and 5.

IDEAL TERNARY MIXTURE

Let us consider an ideal ternary mixture. According to the definition of an ideal mixture (18), the activities of the components (ai) are equal to their mol fractions (equation M25) and their partial molar volumes are equal to those of the pure components (equation M26).

Because

equation M27
(8)
one can write for an ideal mixture
equation M28
(9)

For isothermal-isobaric conditions

equation M29
(10)
and
equation M30
(11)
where V is the molar volume of the ternary mixture.

When equation M31 is a constant, Eqs. 10 and 11 lead to

equation M32
(12)
and when equation M33 is a constant, Eqs. 10 and 11 lead to
equation M34
(13)

By inserting Eqs. 12 and 13 into Eq. 9 at infinite dilution of component 2, one obtains the following expression for equation M35 of an ideal ternary mixture:

equation M36
(14)

On the other hand, expressions for equation M37 for an ideal ternary solution can be also derived by combining Eq. 3 or Eq. 5 with the following Kirkwood-Buff integrals for ideal ternary mixtures (16):

equation M38
(15)
equation M39
(16)
equation M40
(17)
where k is the Boltzmann constant and equation M41 is the isothermal compressibility.

Equation 3 leads to

equation M42
(18)
whereas Eq. 5 to
equation M43
(19)

DISCUSSION

One can see that the result obtained on the basis of Eq. 3 (Eq. 18) coincides with Eq. 14 derived from general thermodynamic considerations, whereas that based on Eq. 5 does not. The numerical difference between the two expressions is very large because the molar volume of a protein is, usually, much larger than the molar volume of the cosolvent.

Whereas the above discussion involves equation M44 the quantity equation M45 which is usually determined experimentally (27), is related to equation M46 through the equation (1,16)

equation M47
(20)
where equation M48 is the partial molar volume of the protein at infinite dilution. equation M49 and equation M50 can be expressed at infinite dilution of component 2 in terms of the Kirkwood-Buff integrals as follows (19):
equation M51
(21)
and (9)
equation M52
(22)

By combining Eqs. 3, 20, 21, and 22, one obtains after some algebra the following simple expression:

equation M53
(23)

Whereas equation M54 and equation M55 depend on the protein characteristics, equation M56 and equation M57 depend only on the characteristics of the protein-free mixed solvent.

For usual cosolvents (organic solvents, salts, etc.), one can use the following approximation of Eq. 23 in the dilute cosolvent range:

equation M58
(24)

Indeed, equation M59 and equation M60 are much smaller than the Kirkwood-Buff integrals for the pairs involving the protein (equation M61 and equation M62). Table 1 provides their values for the system water (1) + lysozyme (2) + urea (3) (pH 7.0, 20°C).

TABLE 1TABLE 1
Numerical values of the Kirkwood-Buff integrals for the water (1) + lysozyme (2) + urea (3) (pH 7.0, 20°C) system

However, when equation M63 and equation M64 are large, and this occurs when the cosolvent is, for example, a polymer (equation M65 (cm3/mol) for the system water/polyethylene glycol 2000 at a weight fraction of polyethylene glycol of 0.02 (21)), the complete Eq. 23 should be used. This conclusion is valid for all large cosolvent molecules (polymers, biomolecules, etc.).

Let us consider the biochemical equilibrium between infinitely dilute native (N) and denaturated (D) states of a protein in a mixed solvent. The changes of the preferential binding parameters equation M66 and equation M67 in this process are given by

equation M68
(25)
equation M69
(26)
and
equation M70
(27)

Equations 25 and 26 follow from Eqs. 3 and 23 by taking into account that equation M71 and equation M72 are characteristics of the protein-free mixed solvent at infinite protein dilution.

The equilibrium constant K of biochemical equilibrium between infinitely dilute native (N) and denaturated (D) states of a protein in a mixed solvent can be expressed in terms of equation M73 (22)

equation M74
(28)
where equation M75 can be provided by experiment (23).

Using for equation M76 and equation M77 expressions from Shulgin and Ruckenstein (16), Eq. 28 can be also rewritten in the form

equation M78
(29)
where equation M79 and equation M80 is the activity coefficient of component i at a mol fraction scale. Let us note that equation M81 is characteristic of the protein-free mixed solvent at infinite protein dilution.

Acknowledgments

We are indebted to Prof. J. Michael Schurr (Dept. of Chemistry, University of Washington, Seattle, WA) for helpful comments regarding this manuscript and for drawing our attention to the fact that the coefficients equation M82 and equation M83 are different.

References
1.
Casassa, E. F., and H. Eisenberg. 1964. Thermodynamic analysis of multi-component solutions. Adv. Protein Chem. 19:287–395. [PubMed].
2.
Noelken, M. E., and S. N. Timasheff. 1967. Preferential solvation of bovine serum albumin in aqueous guanidine hydrochloride. J. Biol. Chem. 1967:5080–5085.
3.
Kuntz, I. D., and W. Kauzmann. 1974. Hydration of proteins and polypeptides. Adv. Protein Chem. 28:239–345. [PubMed].
4.
Timasheff, S. N. 1993. The control of protein stability and association by weak interactions with water: how do solvents affect these processes? Annu. Rev. Biophys. Biomol. Struct. 22:67–97. [PubMed].
5.
Timasheff, S. N. 1998. Control of protein stability and reactions by weakly interacting cosolvents: the simplicity of the complicated. Adv. Protein Chem. 51:355–432. [PubMed].
6.
Record, M. T., W. T. Zhang, and C. F. Anderson. 1998. Analysis of effects of salts and uncharged solutes on protein and nucleic acid equilibria and processes: a practical guide to recognizing and interpreting polyelectrolyte effects, Hofmeister effects and osmotic effects of salts. Adv. Protein Chem. 51:281–353. [PubMed].
7.
Zhang, W. T., M. W. Capp, J. P. Bond, C. F. Anderson, and M. T. Record, Jr. 1996. Thermodynamic characterization of interactions of native bovine serum albumin with highly excluded (glycine betaine) and moderately accumulated (urea) solutes by a novel application of vapor pressure osmometry. Biochemistry. 35:10506–10516. [PubMed].
8.
Anderson, C. F., E. S. Courtenay, and M. T. Record, Jr. 2002. Thermodynamic expressions relating different types of preferential interaction coefficients in solutions containing two solute components. J. Phys. Chem. B. 106:418–433.
9.
Kirkwood, J. G., and F. P. Buff. 1951. The statistical mechanical theory of solutions. J. Chem. Phys. 19:774–777.
10.
Shimizu, S. 2004. Estimating hydration changes upon biomolecular reactions from osmotic stress, high pressure, and preferential interaction coefficients. Proc. Natl. Acad. Sci. USA. 101:1195–1199. [PubMed].
11.
Shimizu, S. 2004. Estimation of excess solvation numbers of water and cosolvents from preferential interaction and volumetric experiments. J. Chem. Phys. 120:4989–4990. [PubMed].
12.
Shimizu, S., and D. J. Smith. 2004. Preferential hydration and the exclusion of cosolvents from protein surfaces. J. Chem. Phys. 121:1148–1154. [PubMed].
13.
Shimizu, S., and C. L. Boon. 2004. The Kirkwood-Buff theory and the effect of cosolvents on biochemical reactions. J. Chem. Phys. 121:9147–9155. [PubMed].
14.
Smith, P. E. 2004. Local chemical potential equalization model for cosolvent effects on biomolecular equilibria. J. Phys. Chem. B. 108:16271–16278.
15.
Smith, P. E. 2004. Cosolvent interactions with biomolecules: relating computer simulation data to experimental thermodynamic data. J. Phys. Chem. B. 108:18716–18724.
16.
Shulgin, I. L., and E. Ruckenstein. 2005. A protein molecule in an aqueous mixed solvent: fluctuation theory outlook. J. Chem. Phys. 123:054909. [PubMed].
17.
Schurr, J. M., D. P. Rangel, and S. R. Aragon. 2005. A contribution to the theory of preferential interaction coefficients. Biophys. J. 89:2258–2276. [PubMed].
18.
Prausnitz, J. M., R. N. Lichtenthaler, and E. Gomes de Azevedo. 1986. Molecular Thermodynamics of Fluid-Phase Equilibria, 2nd Ed. Prentice-Hall, Englewood Cliffs, NJ.
19.
Ruckenstein, E., and I. Shulgin. 2001. Entrainer effect in supercritical mixtures. Fluid Phase Equilib. 180:345–359.
20.
Chitra, R., and P. E. Smith. 2002. Molecular association in solution: a Kirkwood-Buff analysis of sodium chloride, ammonium sulfate, guanidinium chloride, urea, and 2,2,2-trifluoroethanol in water. J. Phys. Chem. B. 106:1491–1500.
21.
Vergara, A., L. Paduano, and R. Sartorio. 2002. Kirkwood-Buff integrals for polymer solvent mixtures. Preferential solvation and volumetric analysis in aqueous PEG solutions. Phys. Chem. Chem. Phys. 4:4716–4723.
22.
Timasheff, S. N. 1992. Water as ligand: preferential binding and exclusion of denaturants in protein unfolding. Biochemistry. 31:9857–9864. [PubMed].
23.
Timasheff, S. N., and G. Xie. 2003. Preferential interactions of urea with lysozyme and their linkage to protein denaturation. Biophys. Chem. 105:421–448. [PubMed].