Next Article in Journal
Closed-Form Solutions for Current Distribution in Ladder-Type Textile Heaters
Previous Article in Journal
A New Numerically Improved Transient Technique for Measuring Thermal Properties of Anisotropic Materials
Previous Article in Special Issue
Molecular, Crystalline, and Microstructures of Lipids from Astrocaryum Species in Guyana and Their Thermal and Flow Behavior
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Linking Solution Microstructure and Solvation Thermodynamics of Mixed-Solvent Systems: Formal Results, Critical Observations, and Modeling Pitfalls

Independent Researcher, Knoxville, TN 37922-3108, USA
Thermo 2024, 4(3), 407-432; https://doi.org/10.3390/thermo4030022
Submission received: 29 August 2024 / Revised: 16 September 2024 / Accepted: 20 September 2024 / Published: 22 September 2024

Abstract

:
This review provides a critical assessment of the current state of affairs regarding the solvation thermodynamics involving mixed-solvent systems. It focuses specifically on (i) its rigorous molecular-based foundations, (ii) the underlying connections between the microstructural behavior of the mixed-solvent environment and its thermodynamic responses, (iii) the microstructural characterization of the behavior of the mixed-solvent environment around the dilute solute via unique fundamental structure-making/-breaking functions and the universal preferential solvation function, (iv) the discussion of potential drawbacks associated with the molecular simulation-based determination of thermodynamic preferential interaction parameters, and (v) the forensic examination of frequent modeling pitfalls behind the interpretation of preferential solvation from experimental data of Gibbs free energy of solute transfer.

1. Introduction

An effective approach to gaining an understanding of solution nonidealities involves the formal interfacing between the microstructural evolution of a system as described by statistical mechanics, and its macroscopic manifestation rationalized by chemical thermodynamics. The sources of success behind this strategy can usually be attributed to the following elements: (a) the inhomogeneous distribution of species in a one-component homogeneous (i.e., one phase diffusively stable) system, i.e., its microstructure, results from the interspecies interactions, so that the non-zero correlation functions reflect the strength of the interactions and translate into their deviations from the uniform (i.e., ideal gas) distribution counterpart [1]; (b) the microstructure of a homogeneous mixture stems from the differences (e.g., interspecies interaction asymmetries) between the unlike pairs giving rise to contrasting pair correlation functions [2,3]; (c) the Kirkwood–Buff fluctuation formalism that connects rigorously the interspecies correlation function integrals, characterizing the microstructure, with the isothermal–isobaric composition dependence of the volumetric and chemical partial molar properties of the system [4]; and (d) the fundamental solution thermodynamics that provides the formal description of the excess properties, i.e., the deviations from a precisely defined ideal solution reference, in terms of the residual properties, i.e., their departure from the ideal gas model [5,6,7].
Naturally, these elements have been the foundational support for the drawing of rigorous microscopic-to-macroscopic interpretations of solvation processes based on the determination of the isobaric–isothermal solute-induced perturbation of the solvent environment, resulting from the solute–solvent intermolecular interaction asymmetry, and the ensuing perturbation propagation across the system [3,8,9]. In fact, this local microstructural distortion, formally described by the related direct correlation function, propagates across the system up to a distance given by the correlation length of the environment as described by the corresponding indirect correlation function [8,10], and allows its macroscopic interpretation in terms of the thermodynamics of solution non-idealities [2,9,11,12,13,14,15].
Unsurprisingly, a significantly large effort in solution chemistry has focused on the development of experimental techniques for the microstructural characterization of fluid mixtures to interpret, and gain an understanding of, the thermophysical properties of real solutions in terms of the structure of the first few solvation shells centered at the species of interest. For that purpose, these techniques have comprised scattering approaches, such as X-ray [16,17,18,19,20,21,22,23] and neutron diffraction [21,24,25,26,27], assisted by empirical microstructural refinement [28,29,30,31,32,33,34], molecular simulation [35,36,37,38,39,40,41], and complemented with infrared [42,43], Raman [44,45,46,47,48], NMR [49,50], and femtosecond pump-probe spectroscopies [51,52]. The effort has frequently led to the unfortunate depiction of the solvation behavior of a species as either structure-making or structure-breaking [53,54,55,56,57], later accompanied by the fancier kosmotrope (order-maker) and chaotrope (disorder-maker) monikers [58], which have contributed further to the ambiguity and confusion when they were used as interchangeable alternative notation [58,59,60,61,62]. In fact, the recurrent factor leading to the structure-making/-breaking confusion has been, until recently, the lack of an explicit and unambiguous formal relation between the microstructural impact of the solute on the surrounding solvent environment, also known as solute-induced local density perturbation, and the macroscopic (thermodynamic) response that can be experimentally measured [63,64,65].
The need for a fundamental understanding of solvation phenomena becomes even more pressing when dealing with species in binary mixed-solvent environments, where the presence of a second solvent, also known as cosolvent or entrainer, modifies the solubility of a solute and the diffusive stability of the solvating environment, through the combination of the contrasting solvating power of the individual solvents, and resulting in significant deviations from solution ideality [66,67,68,69,70,71]. This need has been typically addressed via a variety of binding models [72,73,74,75], supplemented with molecular simulations and integral equation approximations [76,77,78,79,80,81,82], as well as fluctuation-based modeling [83,84,85,86].
On the one hand, the identification of so-called thermodynamic preferential interaction parameters to describe how the chemical potential of the i -solute species in solution, μ i T , P , χ i , χ α , could be modified by changing the concentration χ α of the α -species in an α = j , k mixed-solvent system pioneered a formal thermodynamic description for the interpretation of the experimental evidence [87,88,89]. These parameters have been explicitly defined as isothermal χ -composition derivatives under specific thermodynamic constraints Q α , Q β , including μ α , P as Γ μ α ( χ i 0 ) x j χ k / χ i T P μ α and μ j , μ k as Γ μ j μ k ( χ i 0 ) x j χ k / χ i T μ j μ k , where χ = ρ , m , x denotes the species concentration as either molar, molal, or mole fraction scales, while μ α identifies the chemical potential of the α -species at the prevailing state conditions and composition [90,91,92].
On the other hand, under this thermodynamic approach, we typically encounter a variety of thermodynamic preferential interaction parameters leading to a lack of uniqueness in the resulting microscopic markers for a preferential solvation ( P S ) event. In fact, depending on the chosen thermodynamic constraints Q α , Q β , the composition derivatives within an open (grand canonical) Γ μ i μ k ( χ i ) x j , semi-open (semi-grand), Γ μ α ( χ i ) x j , and a closed (isothermal–isobaric) Γ i α ( χ i ) x j systems translate formally into different functional microstructural descriptions of the system [93]. In other words, how could we assert anything about the P S behavior of a mixed-solvent system when we have neither a precise microstructural manifestation of the differential solute affinity of the solvent environment relative to that of the cosolvent around the solute, nor a unique macroscopic preferential interaction parameter counterpart?
This mystifying scenario was originally underscored by Ben-Naim when first asking “… how do we measure the P S of a given solute in a given mixture?” and then,“… what are the molecular reasons that cause a solute to prefer one component over the other and, hence, alter the composition in its local environment?” [83]. Adding to the confusion was the need for an identification of the formal microstructure-to-macroscopic (i.e., statistical mechanics to thermodynamics) property relationships required to connect a precisely defined molecular-based signature of P S with the underlying solute–solvent, solute–cosolvent, and solvent–cosolvent intermolecular interaction asymmetries leading to the thermodynamic solution non-idealities [93].
The preferential solvation of solute species, momentarily portrayed loosely as the differential affinity of the solute in the mixed-solvent environment (see Section 2.3 for a formal definition), embodies the differential solute-induced perturbations between the solvent and the cosolvent environments. Therefore, it implies a microstructural signature linked explicitly to the experimentally available macroscopic counterparts [91,93,94,95,96], that leads to the rigorous interpretation of the preferential solvation phenomena in terms of the fundamental structure-making/-breaking functions  S α β T , P , x α for any solute in mixed solvents [65,93]. Consequently, it provides the molecular-based framework for the proper interpretation, potential tuning of the species solvation behavior, and eventual thermodynamic modeling of mixed-solvent systems.
In this context, the main goals of this review include addressing the current state of affairs regarding the solvation thermodynamics of mixed-solvent systems, including (i) its rigorous molecular-based foundations, (ii) the underlying connections between so-called preferential interaction parameters and the system microstructure, (iii) the microstructural characterization of the behavior of the mixed-solvent environments around the dilute solute via fundamental structure-making/-breaking functions and the universal preferential solvation function, (iv) the discussion of potential drawbacks behind the molecular simulation-based determination of preferential interaction parameters, and (v) the examination of frequent modeling pitfalls behind the interpretation of preferential solvation from the experimental data of Gibbs free energy of solute transfer.

2. Fundamentals from Statistical Mechanics and Chemical Thermodynamics

The rational manipulation of these solvation phenomena requires a fundamental understanding of the molecular-based mechanism associated with the impact of the state conditions and composition of the mixed-solvent, the solute–solvent, solute–cosolvent, and solvent–cosolvent intermolecular asymmetries on the microstructure of the system as well as the resulting thermodynamic responses, including the diffusive stability of the homogeneous phase.

2.1. Molecular-Based Description of the Structure-Making/-Breaking Functions and Their Attributes

The key elements of interest in the solvation of a solute in a mixed-solvent are the solute-induced microstructural evolution of the surrounding environment and its associated thermodynamic responses [3,9]. In particular, we must recognize the key role played by the relative affinity of the solute with its mixed-solvent environment in the manipulation of the solvation behavior of any solute. In fact, the contrasting affinity represented by the differential perturbation of the solvent relative to that of the cosolvent around the solute becomes the typical microstructural signatures that can be rigorously linked to experimentally available macroscopic quantities [91,94,95,96,97]. For that purpose, we have recently developed an explicit formalism to achieve such microscopic-to-macroscopic connection according to the statistical mechanics-based fundamental structure-making/-breaking functions S α β T , P , x α associated with any species in solution [65,93], whose macroscopic interpretation would lead naturally to the thermodynamic modeling of solvation in mixed-solvents.
The magnitude of the local density perturbation of the β -species environment around any α -species can be rigorously determined in terms of the total correlation function integrals (TCFIs), also known as Kirkwood–Buff integrals [4], as follows [65,98]:
S α β T , P , x α = ρ x β G α β G β β
with S α β T , P , x α S β α T , P , x α where N α β = ρ x β G α β describes the absolute average number of β -species molecules around any central α -species (solute or solvent) at the environmental state conditions and composition, while ρ identifies the number density of the system. Therefore, the function S α β T , P , x α characterizes the average number of β -species around the central α -species in excess/deficit to that encountered around any β -species, and consequently, embodies the required microscopic information to measure the magnitude of the α -species-induced microstructural perturbation of the β -species environment,
G β β T , P , x α β - s p e c i e s   p e r t u r b a t i o n α - s p e c i e s   s o l v a t i o n G α β T , P , x α
where the arrow describes the perturbation of the solvent environment, as a TCFI, from its original magnitude to the final value. In fact, this evolution, G α β G β β T P x , characterizes three possible scenarios as follows:
G α β G β β T P x G α β > G β β S α β > 0   s t r u c t u r e   m a k e r G α β = G β β S α β = 0   n e i t h e r   m a k e r   n o r   b r e a k e r G α β < G β β S α β < 0   s t r u c t u r e   b r e a k e r
Therefore, according to Equations (1)–(3), the solute-induced perturbation of the solvent environment translates unambiguously as a strengthening of the β -species environment around the α -species when G α β > G β β , i.e., S α β > 0 , and the α -species behaves as a structure-maker. Otherwise, when G α β < G β β , i.e., S α β < 0 , the solvent perturbation weakens the β -species environment around the α -species and the α -species becomes a structure-breaker. Obviously, there is a natural transition between the two cases, i.e., when the α -species behaves as the β -species for which G α β = G β β 0 , i.e., S α β = 0 , and consequently, the solvent environment remains unperturbed.
While the three conditions in Equation (3) describe microscopically three potential structure-making/-breaking scenarios for any electrolyte or non-electrolyte binary solution, we need to be able to assess those conditions based on its thermodynamic properties. For that purpose, in Refs. [3,65] we have derived the following formal expressions for the fundamental structure-making/-breaking function  S α β T , P , x α for any α -species at any state condition and composition,
S α β T , P , x α = 1 ρ υ ^ α / ν D ρ υ ^ α < ν D S α β > 0   s t r u c t u r e   m a k e r ρ υ ^ α = ν D S α β = 0   n e i t h e r   m a k e r   n o r   b r e a k e r ρ υ ^ α > ν D S α β < 0   s t r u c t u r e   b r e a k e r
where ν is the stoichiometric coefficient of the electrolyte α -species (i.e., ν = 1 for a non-electrolyte), υ ^ α denotes its partial molar volume, and D denotes the diffusive stability coefficient [7]. In other words, Equation (4) describes the thermodynamic response to the α -species-induced effects on the microstructure of the β -species, and reveals rigorously and unambiguously which of the possible behaviors of the α -species without requiring any explicit microstructural information of the system.
Note that, as hinted above by the inequality S α β T , P , x α S β α T , P , x α , these structure-making/-breaking functions for finite concentrations are defined as pairs, and are formally connected to other solvation-related thermodynamic properties, including the isochoric–isothermal pressure effect upon solvation, P / x α T ρ , the isothermal–isobaric diffusive stability coefficient, D = 1 + ln γ j / ln x j T P , and the universal preferential solvation of a species at infinite dilution in mixed-solvents as lengthily discussed elsewhere [3,93]. However, as one of the species approaches infinite dilution, then
lim x α 0 S α β T , P , x α S α β = 1 υ ^ α / ν υ β o lim x α 0 S β α T , P , x α S β α = 1
so that, the system becomes characterized by just one fundamental structure-making/-breaking function S α β T , P .

2.2. Thermodynamic Preferential Interaction Parameters in Open, Semi-Open, and Closed Systems

Thermodynamic preferential interaction (or binding) parameters in ternary systems are defined according to the number of diffusible species, and consequently, associated with a particular experimental approach including vapor pressure osmometry [99,100], isopiestic distillation [101,102], and equilibrium dialysis [103,104]. Thus, depending on which species in the ternary system is diffusible, either solvent, cosolvent, or both, these preferential interaction parameters are defined as Γ Q j Q k ( χ i ) x j = χ k / χ i T Q j Q k , i.e., the isothermal change in the diffusible solvent concentration in response to the dissolution of a solute when the chemical potential of the diffusible species is kept constant [105,106]. Unfortunately, according to the derivations by Smith [91] and others [90,92,93], these thermodynamic responses lead to a variety of composition-dependent microstructural manifestations depending on the (open/closed) nature of the system under study as illustrated in Table 1. See Appendix B of Ref. [93] for a detailed derivation of these formal expressions.
We can additionally define the thermodynamic preferential solvation parameter in the isothermal–isobaric closed system, Γ i α ( χ i 0 ) x α μ i / μ α T P , sometimes identified in the literature by the symbol ν i α for α = j , k , [86,107] as follows:
Γ i α ( x i 0 ) x α ρ α G i β G i α
where μ i T , P , x j = lim x i 0 μ i k T ln ρ i T P describes the standard chemical potential of the i -solute in the mixed-solvent environment. Therefore, for the α + β -binary solvent, the two preferential solvation parameters are related as x α / Γ i α ( x i 0 ) x α + x β / Γ i β ( x i 0 ) x α = 0 by the Gibbs–Duhem relation.
In summary, given the diversity of the microstructural outcomes as attested in Table 1, it becomes difficult to choose one of the thermodynamic preferential solvation parameters, and its microscopic analog, as the signature for the preferential solvation phenomenon ( P S ). In other words, at this juncture, we could not assert anything about the P S behavior of a solute in a mixed-solvent system given the lack of a precise microstructural manifestation of the differential solute affinity of the solvent environment relative to that of the cosolvent around the solute.

2.3. Universal Preferential Solvation Function and Its Link to the Fundamental Structure-Making/-Breaking Functions

One of the first preferential solvation measures was introduced by Ben-Naim [83] and described as the first-order coefficient of the expansion of the local mole fraction deviation of the solvent around the infinitely dilute i -solute in a mixed j + k -solvent system, δ j i = x j i l o c a l x j , leading to the following formal definition:
δ j i 0 = x j x k G i j G i k
Equation (7) comprises at least two essential ingredients for the formal definition and assessment of P S ; namely, it measures the difference in the i -solute affinities between the solvent and the cosolvent; G i j G i k , and the alluded difference is explicitly and rigorously connected to the Gibbs free energy of transfer between a pure solvent and the desired mixed-solvent environment. Alternative definitions of P S have been given in the literature; including the use of ν i α , Equation (6) [108,109];
P S x j ρ k G i j G i k = ρ o δ j i 0 / x j
where ρ o = lim x i 0 ρ and [110],
P S x j G i j G i k + ρ υ ^ k υ ^ j x i
Unfortunately, as discussed elsewhere [93], Equations (7) and (8) share some common drawbacks regarding the misrepresentation of the actual preferential solvation at the infinite dilution of either the k - or the j -species. In fact, according to Equation (7), δ j i 0 ρ k 0 = 0 regardless of the relative affinity of the i -solute with either the j -solvent or the k -cosolvent, and represents a rather misleading characterization of the relative affinity of the i -solute between the pair of solvents. By considering the definition of relative affinity, R A x α , we find that
R A x k 0 lim x k 0 G i k G i j 0
Equation (10) implies that as x k 0 , the system behaves as an infinitely dilute i + j -binary system so that R A x j 1 = μ i x j / μ j x j T P 0 except for the unlikely case when G i k = G i j at the actual state conditions of the mixed-solvents. Moreover, we should also highlight a yet unreported unfortunate feature of Equation (9) which predicts the unphysical outcome P S , I G _ i x i 0 0 for an ideal gas solute I G _ i , i.e.,
P S , I G _ i x i 0 = ρ i υ ^ k υ ^ j 0
unless υ ^ k = υ ^ j T P x .
The most promising alternative P S definition was recently introduced by the author as a universal state function [93],
P S T , P , x j = ρ o G i k G i j
considering that any formal unambiguous definition must obey a pair of natural constraints in the description of an infinitely dilute i -solute in a j + k -mixed solvent at fixed state conditions. The alluded scenarios are as follows: (a) a non-interacting (ideal gas) i -species must depict a null preference for either solvent species, P S I G _ i x j = 0 , and (b) when a j -solvent and the k -cosolvent behave identically, the i -species prefers equally either species in the j + k -mixed environment, a condition that renders identical solute–solvent and solute–cosolvent intermolecular interaction asymmetries, i.e., P S x j = 0 with G j j = G j k = G k k = k T κ j o υ j o 0 . Moreover, P S x j 0 represents the microstructural response of the j + k mixed solvent to the solute–solvent and solute–cosolvent interaction asymmetries, i.e., differential affinities of the solute between the solvent and the cosolvent whose macroscopic manifestations are given by the thermodynamic excess properties describing the solution non-idealities [7].

2.4. Links between the Thermodynamic Preferential Interaction Parameters and the Universal Preferential Solvation Function

The existing concerns propelled a careful scrutiny of the thermodynamic definitions of preferential interaction parameters and the alternative microstructural signatures of preferential solvation. These concerns were spelled out as a pair of questions: “does preferential interaction mean preferential solvation?, or alternatively, has preferential interaction been unintentionally taken as an equivalent to preferential solvation?” [93]. The microstructural interpretation of the thermodynamic preferential interaction parameters in Table 1 implies that most of them cannot describe preferential solvation, Γ Q j Q k ( χ i 0 ) x j P S x j , because they cannot fulfill the null requirement discussed in the previous section. More problematic is the fact that, as addressed elsewhere, [93] the thermodynamic preferential interaction Γ Q j Q k ( χ i 0 ) x j parameters and the preferential solvation P S x j markers might plausibly be describing either the same microscopic marker, yet resulting from two different thermodynamic responses, or representing two different microstructural scenarios.
In fact, from a microstructural viewpoint, we established the Γ μ j μ k ( m i 0 ) x j = Γ i α ( x i 0 ) x j identity, where the left- and right-hand sides of this expression are representations of two significantly different thermodynamic responses given the nature of the systems involved. To be more precise, the left-hand side describes an open system where Γ μ j μ k ( m i 0 ) x j = m k / m i T μ j μ k m i 0 measures the rate of change in the molal concentration of the k -cosolvent when perturbing the molality of the i -solute around its infinite dilution while the chemical potential of the mixed-solute is kept constant. In contrast, the right-hand side describes a closed system where Γ i α ( x i 0 ) x j = μ i / μ α T P defines the rate of change in the standard chemical potential of the i -solute caused by a perturbation of the composition of the α -solvent. Likewise, Γ μ j μ k ( m i 0 ) x j = Γ i α ( x i 0 ) x j = 0 might be describing two significantly different intermolecular interaction asymmetries, such as a system comprising a real i -solute in an ideal mixed-solvent whose species behave identically, in contrast to another system involving an ideal gas i -solute in a real mixed-solvent. Obviously, these systems represent two completely different microstructural environments, i.e., G i j = G i k 0 for the first scenario, and G i j = G i k = 0 for the second.
In summary, while the description of the thermodynamic preferential interaction parameters and the preferential solvation counterparts are all rigorous for the infinitely dilute solute in a binary mixed-solvent, we must be cautious when discussing preferential interaction and preferential solvation so that they are not taken as equivalent concepts, and avoid the misinterpretations of the experimental evidence, leading to the mischaracterization of the solvation behavior of species in mixed-solvent environments.

3. Exploring the Preferential Solvation Behavior of Solutes in Mixed-Solvents

The differential solvation behavior of an i -solute at infinite dilution in a j + k -binary solvent environment has been described in terms of precisely defined fundamental structure-making/-breaking functions  S α β [93], and its macroscopic (thermodynamic) responses through the Kirkwood–Buff fluctuation formalism [111]. This depiction ensures a highly desirable uniqueness in the outcome while imposing no constraints on the type of solute, the nature of the mixed-solvents, or the solute–solvent and solute–cosolvent intermolecular interaction asymmetries.
To explore the differential solvation behavior of any i -solute at infinite dilution in a j + k -binary solvent, let us consider an isobaric–isothermal ternary system comprising N j solvent and N k cosolvent particles with N i solute species such that N j + N k N i 0 . Under these conditions, we denote the number density of an α -species as ρ α = ρ x α , where ρ N i + N j + N k / V with the mole fraction x α = N α / N . To connect the microscopic behavior of the solvents around the i -solute and the macroscopic (thermodynamic) responses, we recall that at isothermal conditions, the species densities in the phase are functions of the chemical potentials in the grand canonical (open) ensemble, so that for our system we have,
k T d ln ρ i T = 1 + ρ i G i i d μ i + ρ j G i j d μ j + ρ k G i k d μ k
Consequently, after invoking the definition of the standard chemical potential μ i T , P , x j and the ρ i 0 limit, we have the following:
d μ i x j T P = ρ j G i j d μ j ρ k G i k d μ k
where the system is now a binary j + k -mixed solvent whose Gibbs–Duhem relation ρ j d μ j + ρ k d μ k T P = 0 leads to the desired expression linking thermodynamics, the left-hand side of Equation (15), and microstructure, the right-hand side of Equation (15), as follows:
μ i T , P , x j / μ j T P = ρ j G i k G i j = x j P S
Equation (15) embodies the mechanism behind the tuning of the solvation behavior of an i -solute in a (diffusively stable, i.e., finite D x j > 0 ) mixed j + k -solvent environment through the manipulation of the chemical potential of the j -solvent.
To draw a more explicit link, we need to formally determine the five T C F I ’s (also known as Kirkwood–Buff integrals), G α β , through the inversion of the Kirkwood–Buff formalism [112]. Typically, we first extract the first three functional relations G α β υ ^ α , υ ^ β , κ o , D of the mixed-solvent phase given explicitly as follows (see Appendix C of Ref. [93] for details):
G j k = k T κ o υ ^ j υ ^ k ρ o D 1
G j j = G j k + ρ o υ ^ k / D 1 ρ j 1
G k k = G j k + ρ o υ ^ j / D 1 ρ k 1
in terms of the species partial molar volumes υ ^ i , υ ^ j , υ ^ k , the diffusive stability coefficient D , and the isothermal compressibility κ o of the mixed-solvent environment. For the remaining pair G i α υ ^ i , υ ^ β , κ o , D , μ i x j / x j T P integrals, we need the composition dependence of the partial molar volume of the i -solute, υ ^ i x j , and the rate of change in the standard chemical potential of the i -solute in response to variations in the mixed-solvent composition, μ i x j / x j T P . In particular, this thermodynamic derivative is usually available in the following equivalent forms, whose choice might depend on either the type of solute or the nature of the experimental solubility data (see Appendix D of Ref. [113] for details)
μ i x j / x j T P = Δ t r g i x j / x j T P k T ln ϕ ^ i z / x j T P μ i x j / x j T P
where the first line on the right of Equation (19) involves the Gibbs free energy of the transfer of the i -solute between the pure k -cosolvent phase to the binary j + k -mixed solvent, and the second and third lines invoke two representations of the isothermal–isochoric residual chemical potential of the i -solute at infinite dilution, μ i r T , ρ o , x j , with μ i * T , ρ o , x j identifying the pseudo chemical potential [114]. Then, from Equations (15) and (19), we complete the inversion, i.e.,
G i j = k T κ o υ ^ i x k υ ^ k D 1 μ i x j / x j T P
and,
G i k = k T κ o υ ^ i + x j υ ^ j D 1 μ i x j / x j T P
Finally, from Equations (15)–(19), we find that either the microscopic quantity P S T , P , x j ,
P S T , P , x j = S i k S j k / x k S i j S k j / x j
or the thermodynamic property μ i T , P , x j / μ j T P ,
μ i T , P , x j / μ j T P = x j / x k S i k S j k S i j S k j
is described entirely in terms of a linear combination of the same four fundamental structure-making/-breaking functions. These equations provide the formal and rigorous interpretation of the source of the differential solvation, i.e., it results from a linear combination of solute-induced and solvent-induced effects on the microstructure of the mixed-solvent environment, while highlighting the link to the non-ideality of the mixed-solvent ternary system.

3.1. Preferential Solvation of Non-Polar Gasses in Polar Mixed-Solvent Environments

The understanding of the solvation behavior of volatile solutes in mixed-solvent environments is of great theoretical and practical relevance [65,93,115,116]. In fact, while traditionally challenging to model properly given the large solute–solvent intermolecular interaction asymmetries [65,93,115,116], the availability of a complete experimental dataset provided the opportunity to apply the discussed formalism to study a series of non-polar gases in 2,2,2-trifluoroethanol–water mixed-solvent systems based on the solubility of He , Ne , Ar , Kr , Xe , H 2 , N 2 , O 2 , CH 4 , C 2 H 4 , C 2 H 6 , CO 2 , CF 4 , and SF 6 at ambient conditions in light water [117], 2,2,2-trifluoroethanol (TFE) [118], and the 2,2,2-trifluoroethanol–water mixed-solvent over the entire range of composition [119,120].
For the purpose of the Kirkwood–Buff inversion, the solubility data were represented by μ i x j / x j T P = ln ϕ ^ i z / x j T P in Equations (20) and (21), and complemented with the composition dependence of the partial molar volumes [121], the isothermal compressibility [122], and the activity coefficients [123] of the gases in 2,2,2-trifluoroethanol–water mixed-solvent. Moreover, for the partial molar volume of the i -solutes at infinite dilution in the mixed-solvents, we invoked the composition dependence resulting from the quasi-ideal solution (also known as Krichevskii’s relation [124,125]) behavior of Henry’s law constant of the solutes in the mixed-solvent environments, ln H i , j + k E x j 0 [119,120], i.e.,
ln H i , j + k K r i c h e v s k i i I S T , P , x j x j ln H i , j T , P + 1 x j ln H i , k T , P
where H i , α T , P identifies Henry’s law constant of the i -solute in a pure α -solvent at the prevailing state conditions (e.g., see Figure 1 in Ref. [113]). This relation, written in the following alternative form,
ln ϕ ^ i , j + k T , P , x j x j ln ϕ ^ i , j + 1 x j ln ϕ ^ i , k
whose isothermal-pressure derivative, after assuming the validity of Equation (24) over the isothermal P ± Δ P range, leads to the sought expression for υ ^ i , j + k T , P , x j ,
υ ^ i , j + k T , P , x j x j υ ^ i , j T , P + 1 x j υ ^ i , k T , P
where υ ^ i , α T , P denotes the partial molar volume counterparts to H i , α T , P .
The isobaric–isothermal composition-dependent microstructure of the TFE + H 2 O -system becomes represented by G T F E   T F E   x T F E < 0.32 > 0 with a maximum at x T F E 0.16 and G T F E   T F E   x T F E 0.32 < 0 , as well as G H 2 O   H 2 O   x T F E < 0.05 < 0 and G H 2 O   H 2 O   x T F E 0.05 > 0 , resulting in a significantly large positive deviation from Lewis–Randall ideality, Δ T F E   H 2 O L R I S T , P , x T F E = 0 , [112,126] as depicted in Figure 1.
On the one hand, this picture shows a significant magnification of G H 2 O   H 2 O   x T F E from its pure component counterpart G H 2 O   H 2 O o   =   k T κ H 2 O o υ H 2 O o , with its maximum value at x T F E 0.25 , while G T F E   H 2 O   0 x T F E 1 < 0 and exhibits a minimum at x T F E 0.2 . This microstructural behavior translates also into a significant reduction in the diffusive stability coefficient, D x T F E 0.2 0.1 , from its Lewis-Randall ideality D L R I S x T F E = 1 counterpart as depicted in Figure 2a, which indicates a close approach to the incipient immiscibility condition [127]. On the other hand, the observed maxima in the total correlation function integrals G H 2 O   H 2 O   x T F E 0.25 and G T F E   T F E   x T F E 0.16 are the microstructural manifestations of like-species aggregation in Figure 1, which can be identified as a H 2 O -induced structure-breaking of the TFE -environment, S H 2 O   T F E   x T F E < 0 , and simultaneously, a TFE -induced structure-breaking of the H 2 O -environment, S T F E   H 2 O x T F E < 0 , as evidenced in Figure 2b.
For the interpretation of the differential solvation behavior of the non-polar gases in the mixed TFE + H 2 O -solvent environment, we evaluate simultaneously the isobaric-isothermal behavior of the two relevant structure-making/-breaking functions S i   T F E x T F E and S i   H 2 O x T F E , as well as the resulting preferential solvation function  P S x T F E , where the i -species identifies one of the fourteen non-polar gas solutes at infinite dilution and ambient conditions. In fact, from Figure 3a,b we can highlight some common features underlying the behavior of S i   T F E x T F E and S i   H 2 O x T F E , which translates into the final differential solvation of the infinitely dilute solutes between the solvent and the cosolvent.
On the one hand, the function S i     T F E x T F E in Figure 3a starts at S i     T F E x T F E = 0 = 0 , follows through a negligibly small negative value at x T F E 0.03 , becomes quickly large and positive at increasing TFE -concentration, goes through a maximum within 0.15 < x T F E < 0.25 , to finally decrease to S i   T F E x T F E = 1 = ρ T F E o G   i   T F E G T F E   T F E o 0 . This behavior indicates that the solvation of the gas solutes enhances the microstructure of the TFE -solvent, i.e., they behave as structure-making i -solutes, except for the negligibly small structure-breaking effect for x T F E 0.03 , and for i = CF 4 , SF 6 at x T F E > 0.6 . On the other hand, in Figure 3b S i   H 2 O x T F E starts at S i   H 2 O x T F E = 0 = ρ H 2 O o G   i   H 2 O G H 2 O   H 2 O o 0 , decreases significantly up to a minimum at x T F E 0.2 (similar to that for the diffusive stability coefficient D ), and converges at S i   H 2 O x T F E = 1 = 0 . Thus, the solvation of these gas solutes strongly weakens the microstructure of the H 2 O -solvent, i.e., they behave as structure-breaking i -solutes over the entire range of mixed-solvent composition.
The observed opposite structure-making/-breaking trends in Figure 3a,b give rise to a prominent change in the sign of the preferential solvation functions of the gas solutes, i.e., from P S x T F E 0 0 to P S x T F E 1 0 , through a significantly large minimum P S 0.15 x T F E 0.19 0 , as depicted in Figure 4a. This behavior appears as a common feature of all these gas solutes in the mixed TFE + H 2 O -solvent environment, and becomes the signature of the fact that these gas solutes are better solvated by water than by the mixed TFE + H 2 O -environment when the solvent is highly dilute in TFE , i.e., except for the narrow concentration range. Indeed, for 0.03 x T F E 0.05 , these gas solutes are equally solvated by the two solvents, while at x T F E 0.03 these solutes become better solvated by TFE than water. Thus, if we wish to tune the solubility of these gas solutes, Figure 4b illustrates how the manipulation of the chemical potential of the TFE -solvent provides the way to accomplish the task, i.e., μ i x T F E / μ T F E T P or its equivalent form,
μ i x T F E / μ T F E T P = k T ln H i , T F E + H 2 O x T F E / ρ k T / μ T F E T P
In other words, Equation (27) and Figure 4b indicate that by perturbing the chemical potential of the TFE -solvent (through changes in the composition of the mixed-solvent), we can modify the standard chemical potential of any of the gas solutes, and consequently, their solubility, i.e., the magnitude of the i -solute’s Henry law constant.

3.2. Preferential Solvation of Pharmaceutical Species in Polar Mixed-Solvent Environments

The incorporation of a second solvent might have opposite effects on the resulting solute solubility; cosolvents might increase the solubility of non-polar drugs over that in aqueous environments, a desired goal for the formulations of concentrated solutions of non-polar drugs for systemic therapeutic treatments [128,129]; meanwhile, a co-non-solvent will drastically reduce the solubility of a solute in mixed-solvent environments whose individual components exhibit good solvency, to become a versatile tool for the manipulation of solute aggregation and/or crystallization toward specific pharmaceutical applications [130,131,132]. For example, we have recently explored the preferential solvation behavior of a few pharmaceutical solutes [133], for which the Kirkwood–Buff inversion involved the solubility data as μ i x j / x j T P = Δ t r g i x j / x j T P   in Equations (20) and (21), and comprising a common mixed ethyl acetate (j) + ethyl alcohol (k) environment, including indomethacin [134], naproxen [134], meloxicam [135], and piroxicam [136], used as nonsteroidal anti-inflammatory drugs, while mevastatin [137] prescribed as a cholesterol-lowering agent, and posaconazole [138], involved in the treatment of fungal infections [139].
In Figure 5a, we display the universal preferential solvation P S x E T A C behavior of these six pharmaceutical solutes in ethyl acetate–ethanol mixed-solvent systems, and contrast them with the corresponding first-order preferential solvation parameter δ E T A C   i 0 x E T A C in Figure 5b. The remarkably similar outcome of the universal preferential solvation behavior of this pharmaceutical species is their increasing trend with the increasing concentration of ethyl acetate, except for posaconazole which shows a significant increase up to x E T A C 0.6 , and then a decrease to slightly negative, when it reaches the condition of pure ethyl acetate. The fact that P S x E T A C < x 0 < 0 in Figure 5a, where x 0 the composition of the mixed-solvent at which P S x E T A C = x 0 = 0 , confirms that these pharmaceutical solutes prefer ethyl acetate over ethanol as the solvation environment within the indicated range of ethyl acetate mole fraction; in other words, it manifests microscopically as G i   E T A C x E T A C < x 0 > G i   E T O H x E T A C < x 0 and macroscopically as Δ t r g i 0 < x E T A C < x 0 < 0 , Δ t r g i x E T A C < x 0 / x E T A C T P < 0 as well as μ i x E T A C / μ E T A C x E T A C T P < 0 . Note, however, that not all the observed trends are monotonous in mixed-solvent composition; in fact, they exhibit inflection points (e.g., piroxicam and meloxican) as well as relative extrema (e.g., mevastatin and posaconazole).
It is also useful to highlight the contrasting behavior between the universal P S x E T A C function and the always non-monotonous Ben-Naim’s first-order δ E T A C   i 0 x E T A C parameter comprising three roots, i.e., δ E T A C   i 0 x E T A C = 0 for x E T A C = 0 ; x E T A C = 1 ; 0 < x E T A C < 1 . The contrasting behavior underscores the main drawback of δ E T A C   i 0 x E T A C as a marker for preferential solvation as extensively discussed elsewhere [93].

4. Discussion on Alternative Exploration Routes and Relevant Observations

Whenever we have the ability to either generate (e.g., by molecular simulations) or extract (e.g., from scattering experiments) the proper microstructure of a mixture, there is an opportunity for us to study a variety of solvation phenomena and attempt linking microstructural details with thermodynamic behavior. This approach is usually guided by statistical mechanics in conjunction with chemical thermodynamics as previously stated in the introduction. However, to obtain meaningful outcomes, we need to be aware of the proper foundations, the underlying hypothesis, and the potential limitations of the invoked formalisms as we discuss and illustrate below.

4.1. Molecular Simulation Approaches to Thermodynamic Preferential Interaction Parameters

Molecular-based simulation of model systems has been targeted as a source of information on the preferential interaction/solvation of infinitely dilute solutes in mixed-solvent has been molecular simulation of model ternary systems [80,140,141]. These studies typically invoke the molality-based thermodynamic preferential interaction parameter Γ μ j μ k ( m i 0 ) (e.g., Table 1), i.e.,
Γ μ j μ k ( m i 0 ) = x k P S T , P , x j = N i k r e s , N k / N j N i j r e s ,
where the above microstructural interpretation relies on the Kirkwood–Buff fluctuation formalism, and therefore, on the ability to extract a meaningful representation of N α β r e s , ρ β G α β from simulation, i.e.,
N α β r e s , = 4 π ρ β 0 g α β r α β , ω α , ω β ω α ω β 1 r α β 2 d r α β
with the orientational average ω α ω β , where ω α describes the orientation of the α -species and r α β defines the distance between the centers of mass of the two species [1].
The second line of Equation (28), however, is actually approximated according to the truncated version of the corresponding volume integrals as follows,
Γ μ j μ k ( m i 0 ) = N i k r e s , N k / N j N i j r e s , n i k r < r c μ V T n k μ V T / n j μ V T n i j r < r c μ V T
where the μ V T denotes a grand-canonical ensemble average and r c describes the truncation/cutoff distance which defines the radius of the preferential interaction shell, also known as the correlation radius, while N i α r e s , = n i α r c μ V T and N α = n α r c μ V T . Therefore, the approximation in the second line of Equation (30) assumes implicitly that,
n i k r r c μ V T n k μ V T / n j μ V T n i j r r c μ V T 0
In turn, the right-hand-side of Equation (28) is estimated in terms of the following molecular simulation time averages, i.e.,
Γ μ j μ k ( m i 0 ) τ = 1 τ r u n / τ r e c Γ μ j μ k ( m i 0 ) τ / τ r u n / τ r e c
with (e.g., S11 of Ref. [142])
Γ μ j μ k ( m i 0 ) τ = n i k r < r c , τ n k n i k r < r c , τ n j n i j r < r c , τ n i j r < r c , τ
where n α β r < r c , τ denotes the number of α -species separated by a center of mass-to-center of mass r -distance from the β -species, while n α is the total number of α -species in the simulation system. Given that r c < 10 Å is used in most cases [140,141,143], and considering the simulation evidence on the radial distribution functions of representative aqueous systems [79,142,144,145,146,147,148], it becomes difficult to expect an accurate representation of Equation (28) based on the simulation determination of Equation (30), i.e., the condition described by Equation (31) would not be fulfilled, especially for large i -solutes such as proteins.
As long as r c refers to the center of mass-to-center of mass, the shortcomings from this issue might be alleviated by increasing the cutoff distance; however, the simulation approach suffers from another significantly more serious issue that has been usually overlooked. This issue concerns the actual statistical mechanic meaning of N α β r e s , = ρ β G α β . In fact, in the formulation of the preferential interaction parameter, Equation (28), N α β r e s , r < r c must be evaluated as a volume integral over the α -center of mass to β -center of mass radial distances. However, the reported simulation results from the application of Equation (29) frequently involve a radial distance other than the required center of mass-to-center of mass, including the “closest distance from the solvent or cosolvent to any atom of the protein solute ” [144], the “the distance from the solvent/cosolvent to the surface of the solute protein” [82,149,150], and “the radial distance from the solvent or cosolvent to the protein van der Waals surface” [141,151]. Obviously, these simulation prescriptions do not, and cannot, describe the truncated form N α β r e s , r c = ρ β G α β r c in that, the simulated N α β r e s , r c becomes defined as a truncated integral over a radial correlation function between sites other than the species centers of mass, i.e., it neither represents Kirkwood-Buff integrals, nor can it fulfill any Kirkwood-Buff fluctuation theory relationship [4]. Any choice of pair distance other than the center-to-center of mass in the determination of the simulation averages introduces an implicit angular dependence into the pair correlation functions invalidating all further relationships, including the one used in the simulation approach, e.g., Equations (28) and (29).

4.2. Local Composition-Based Preferential Solvation Approach

As discussed in §3, the starting point of the microstructural-based analysis of preferential solvation is the inversion of the Kirkwood-Buff fluctuation theory to obtain the formal expressions for the calculation of the five relevant Kirkwood-Buff G α β integrals. In addition to performing such formal inversion, Ben-Naim [83] proposed a means to measure the preferential solvation of a solute at infinite dilution in a binary α + β -mixed solvent according to the idea of local composition and the deviation from its bulk counterpart, i.e.,
δ α i R c = x α i R c x α = x α x β G α i G β i / x α G α i + x β G β i + V c
where the local mole fraction of the α -solvent around the i -solute is defined as follows:
x α i R c = N i α R c / N i α R c + N i β R c
with N i α R c describing the number of α -solvent molecules in the correlation shell of radius R c centered at the i -solute. This number becomes formally given by the following statistical mechanic expression:
N i α R c 4 π ρ α 0 R c g i α r i α r i α 2 d r i α = ρ α G α i + ρ α V c R c
which highlights its explicit dependence on the size of R c , and consequently, that of the local composition x α i R c .
The R c -dependence x α i R c becomes an intrinsic drawback of the local composition definition, and consequently, the outcome of δ α i R c and any idea of preferential solvation becomes dependent on the arbitrary choice of the coordination volume V c R c . To avoid such an issue, Ben-Naim suggested an R c -independent preferential solvation parameter δ α i 0 , resulting from the first-order expansion of Equation (34) with respect to V c 1 , i.e.,
δ α i 0 = x α x β G α i G β i
where G α β T , P represents the Kirkwood-Buff integral for the α β -pair of species. We should note that Equation (37) describes null preferential solvation at the two edges of the composition range, regardless of the solvent-to-cosolvent relative solute affinities, G α i G β i T P x = ρ o 1 P S , where typically G α i T , P , x j G β i T , P , x j as clearly illustrated in Figure 5a,b.
A variation around Ben-Naim’s δ α i 0 was introduced by Marcus, [94] who took Equation (34) and invoked a conjectured expression for the estimation of the composition-dependent correlation volume V c υ ^ i , x α i , x α such as Equation 8a,b in Ref. [152]. Thus, the pair of Equations, δ α i V c , x α i and V c υ ^ i , x α i , x α , must be solved simultaneously by iteration to obtain the desired δ α i V c , x α i . It becomes immediately obvious that regardless of the procedure used in the estimation of V c υ ^ i , x α i , x α , Equation (34) involves a potential source of trouble given the lack of explicit enforcement of the splitting requirement
G i α T , P , x α = 4 π 0 R c > ζ g i α r i α r i α 2 d r i α V c R c > ζ
where ζ identifies the correlation length of the mixed-solvent environment, i.e., the distance traveled by the solute-induced perturbation of the microstructure of the solvent environment around it [3]. While this requirement is automatically fulfilled by the first-order preferential solvation parameter δ α i 0 x α x β G i α G i β , we must also be mindful that the resulting R c = 3 V c υ ^ i , x α i , x α / 4 π 3 from the iterative solution of the pair, δ α i V c , x α i and V c υ ^ i , x α i , x α , might translate into an R c ζ as suggested by the microstructural information from the molecular simulation of aqueous systems [147,148].
Here, we should also bring to the reader’s attention a pair of frequently encountered modeling pitfalls when applying this approach. Note that, when discussing the Kirkwood-Buff formal inversion, we highlighted the condition of diffusive stability, as described by a finite positive stability coefficient D 0 x j 1 [7]. In other words, the binary mixed-solvent system must exhibit full miscibility over the entire compositional range at the prevailing state conditions. Otherwise, the Kirkwood-Buff inversion cannot be achieved within the miscibility gap, and the failure to recognize this condition has led to a variety of bizarre interpretations of the faulty preferential solvation calculations as thoroughly discussed elsewhere [93,113]. Moreover, it should be obvious that no matter how we approach the iterative solution of δ α i V c , x α i and V c υ ^ i , x α i , x α , the calculated preferential solvation δ α i V c , x α = x α i R c , x α x α must always result in a physically meaningful local mole fraction, i.e., 0 < x α i R c , x α < 1 . Unfortunately, we have encountered a considerably large number of published research concerning the solvent effects on the solubility of pharmaceutical species (as discussed and sampled in SI.1 and SI.2 of the Supplementary Information of Ref. [133]) where the data comprised δ α i V c , x α < 0 that led to physically meaningless x α i V c , x α = δ α i V c , x α + x α < 0 .
Given the number of publications exhibiting the same problem, we performed a forensic analysis of the available tabulated data to find that most of the affected publications invoking Marcus’ approach—the iterative solution of δ α i V c , x α i given by Equation (34) and the correlation volume V c υ ^ i , x α i , x α as given in Ref. [152]—followed literally either Equations 5.33a,b of Ref. [67], Equation (6) of Ref. [153], Equations 2.51–2.52 of Ref. [154], or Equations 6.25a,b of Ref. [155]. A compilation of publications involving this Kirkwood-Buff inversion issue (about 165 publications to date) is given in the Supplementary Information of Ref. [133].
It turned out that the unfortunate misprint in the original sources, Refs. [67,153,154,155], written in our notation as follows:
G i j , f a u l t y = k T κ o υ ^ i + x k υ ^ k k T D 1 μ i x j / x j T P = G i j + 2 x k υ ^ k k T D 1 μ i x j / x j T P
has gone unnoticed, where G i j , f a u l t y identifies the misprinted form of the correct G i j quantity given by Equation (20), after the equivalences represented by Equation (19). Equation (39) highlights the misprinted sign in the third term of its first line, an error that might have become immediately obvious should the users have checked the difference between Equations (20) and (21) to recover the original solubility expression, Equation (15), or its equivalent form,
μ i x j / x j T P = k T D ρ o G i k G i j
instead of the faulty outcome,
μ i x j / x j T P f a u l t y = k T D G i k G i j , f a u l t y x j υ ^ j x k υ ^ k 1
which satisfies the correct Equation (15) only when x j = 1 . Moreover, the introduction of G i j , f a u l t y G i k into the numerator of the preferential solvation parameter, Equation (34), will generate a faulty preferential solvation parameter, i.e.,
δ j i f a u l t y x j = δ j i x j x j G i j + x k G i k + V c + 2 x j x k 2 υ ^ k k T D 1 μ i x j / x j T P x j G i j , f a u l t y + x k G i k + V c f a u l t y
Moreover, as a matter of illustration in Figure 6a,b, we display side by side the comparison between ρ o δ j i 0 , f a u l t y = x j x k ρ o G i j , f a u l t y G i k against the properly inverted ρ o δ j i 0 = x j x k P S x j x k ρ o G i j G i k for the same set of systems in Figure 5a,b. Figure 6a,b shows clearly the contrasting behavior between ρ o δ E T A C   i 0 , f a u l t y x E T A C and ρ o δ E T A C   i 0 x E T A C , e.g., ρ o δ E T A C   i 0 x E T A C < 0.37 > 0 while ρ o δ E T A C   i 0 , f a u l t y x E T A C < 0.37 < 0 , where x E T A C 0.37 identifies the common middle root of δ E T A C   i 0 , f a u l t y x E T A C = 0 for the mixed ethyl acetate (j) + ethanol (k) environment. This common middle root of δ j i 0 , f a u l t y x j = 0 for a mixed-solvent system becomes independent of the identity of the solute as demonstrated in Appendix A or Ref. [133].
In summary, the use of the incorrect G i j , f a u l t y rather than G i j Kirkwood–Buff integral as input in the simultaneous determination of the preferential solvation parameter δ j i V c , x j i and corresponding correlation volume V c υ ^ i , x j i , x j leads to the iteration of Equation (42) rather than Equation (34) and steers always to the incorrect, and frequently, unphysical results as thoroughly discussed elsewhere [133].

4.3. Controversial Definitions and Claims about Preferential Solvation

We frequently find unfortunate misconceptions surrounding the differential solvation of a solute at infinite dilution in a binary j + k -solvent when the mixed-solvent behaves as a Lewis-Randall ideal solution, i.e., Δ j k L R I S T , P , x j G j j + G k k 2 G j k = 0 [67,152,156,157,158]. For instance, in Ref. [152] the author indicates that “A feature to be noted in the application of Equation (7) is that even for ideal mixtures (where G E = 0 , V E = 0 and G i i + G j j = 2 G i j the δ i j are non-zero, as if preferential solvation existed in them”, and consequently, follows a suggestion that the three Kirkwood-Buff integrals “be corrected” to force δ i j L R I S = 0 . These rather controversial ideas have been proven misleading and possibly wrong, as discussed by Ben-Naim [159,160]. In fact, under the Lewis-Randall ideality condition, G j k G j j = G j k G k k [126], which indicates that there is an exact compensation of affinities between j k - and either j j - or k k -intermolecular type interactions. In other words, even when the mixed-solvent system (at some precise state conditions and composition) exhibits Δ j k L R I S T , P , x j = 0 , the average solvent (cosolvent) environment around any cosolvent (solvent) species differs from that around itself as formally described by the differences G j k G j j = G j k G k k 0 . The only possible exception occurs when the j -solvent behaves identically as the k -cosolvent, as discussed in §2.3. Consequently, in terms of the fundamental structure-making/-breaking functions, we have that S k j / ρ j = S j k / ρ k 0 , i.e., a non-zero preferential solvation for both species (for additional details, see Ref. [93]).

4.4. Thermodynamic Consistency of the Input Properties

For the correct implementation of the formalism, we also need to handle properly the inputs, i.e., ensuring their thermodynamic consistency, including the representations of υ ^ i T , P , x j , κ o T , P , x j , and any excess property P E T , P , x j . In particular, all the regressed excess property should obey the P E T , P , x j = 1 = 0 and P E T , P , x j = 0 = 0 conditions. Unfortunately, these crucial requirements often go unfulfilled in the cited literature, when the regressed Gibbs free energy g E T , P , x j for the mixed-solvent describes g E x j = 0     and / or   x j = 1 0 [134,161,162,163,164,165,166], i.e., not only an obvious thermodynamic inconsistency, but also a source of corruption of the resulting composition profiles of the solvent–cosolvent intermolecular interaction asymmetry as described by Δ j k T , P , x j . Consequently, this thermodynamic inconsistency translates into an ill-behaved coefficient of diffusive stability D T , P , x j , spills over the determination of all Kirkwood–Buff integrals, as well as distorts the calculation of P S T , P , x j and S α β T , P , x j .

5. Concluding Remarks, Recommendations, and Outlook

In this review, we have placed most emphasis on the rigorous formal connections between microstructural behavior and thermodynamic responses involving the solvation of infinitely dilute species in binary mixed-solvents. The explicit microscopic-to-macroscopic relationships capture the fundamental notion underlying the thermodynamic solvent effects and their applications in physico-chemical separation processes, while highlighting one of the main shortcomings in the existing studies of preferential solvation. Thus, the main goal behind this emphasis, regardless of the choice of modeling scheme for the solvent effects, is the rational guide to the selection of cosolvents for any specific application, e.g., to improve separation, enhance solubility, and control transport properties through the manipulation of the mixed-solvent environment. Indeed, the alluded shortcomings in recent studies have prevented revealing how the preferential solvation parameters (or any adequate relative affinity marker) might be quantitatively applied to manipulate the solvation power of a mixed-solvent environment for any desired application while avoiding the unsupported interpretations of solvation behaviors according to the handwaving mechanisms of intermolecular interactions [167,168,169,170,171].
In addition, the identified microscopic-to-macroscopic relationships in terms of rigorous fundamental structure-making/-breaking functions have offered us versatile tools to (a) study the actual preferential solvation behavior of a contrasting variety of solutes in binary mixed-solvents, (b) provide a rigorous interpretation of the differential affinity between the solute and the components of the solvent environment in terms of an unambiguous preferential solvation function, (c) discuss the ability of a popular local composition model of solvent effects and test the correctness of its underlying Equations, and (d) conduct a forensic analysis to detect the source of some thermodynamic pitfalls in a number of recently published work on the modeling of preferential solvation.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The author expresses his gratitude to Olesya Bondarenko for her kindness in translating the original Russian version of the Ref. [124].

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

Symbols
D T , P , x j diffusive or material stability coefficient 1 + ln γ j / ln x j T P
g α β r α β , ω α , ω β spatial distribution function for the α β -centers of mass interactions
g i α r i α radial distribution function for the i α -centers of mass interactions
g E T , P , x α Isobaric-isothermal excess Gibbs free energy of the binary mixed-solvent
G α β T , P , x α Kirkwood-Buff integral for the α β -interactions
H i , j I S T , P Henry’s law constant of an i -species in a j -solvent given by H i , j I S T P = f i o T P γ i L R , T P
k Boltzmann constant
K B Kirkwood-Buff
m α molality of the α -species in solution
N α number of molecules/moles of the α -species in the system
N i α R c represents the average number of α -solvent within the correlation shell of radii R c
P S T , P , x α universal preferential solvation function
P E T , P , x α generic isobaric-isothermal excess property of the mixed-solvent
Q α , Q β thermodynamic constraints in the definition of χ k / χ i T Q α Q β parameters
R A x α relative affinity described by differences in Kirkwood-Buff integrals
r c molecular simulation cutoff radius for the definition of preferential solvation, e.g., as in Equations (30)–(33).
R c radius of the correlation volume where the local composition is defined
S α β T , P , x β structure-making/breaking function for the α β -interactions
T C F I total correlation function integral, also known as Kirkwood-Buff integral
T , P state conditions defined by the system temperature and pressure
T , ρ o state conditions defined by the system temperature and density
V c correlation volume where the local composition is defined
υ ^ α T , P , x α partial molecular/molar volume of the α -species
x α liquid phase composition defined by the mole fraction of the α -species
x α i l o c a l local mole fraction of α -species around the i -species
β k T 1
χ α generic composition scale for the α -species, e.g., χ α ρ α , m α , x α
δ α i R c , x α deviation of the local mole fraction of the α -solvent around the i -solute, also known as preferential solvation parameter
δ α i 0 Ben-Naim’s first-order preferential solvation parameter
Δ α β T , P , x α linear combination of Kirkwood-Buff integrals as marker of deviations from Lewis-Randall ideality, i.e., G α α + G β β 2 G α β T P
Δ t r g i T , P , x j transfer Gibbs free energy of the i -solute
ln ϕ ^ α T , P , x α partial molecular/molar fugacity coefficient of the α -species
ln ϕ ^ i z k T 1 -times the isochoric–isothermal residual chemical potential of the i -solute at infinite dilution in the mixed-solvent environment
γ α T , P , x α Lewis-Randall’s activity coefficient of the α -species, i.e., ϕ ^ α T , P , x α / ϕ α o T , P
Γ i j x i 0 T , P , x j Isobaric-isothermal thermodynamic preferential interaction parameter
η o T , P , x j ρ o / D
κ j o T , P isothermal compressibility of the pure j -solvent
κ o T , P , x α isothermal compressibility of the mixed-solvent environment
μ α T , P , x α chemical potential of the α -species
μ α * T , P , x α pseudo-chemical potential of the α -species
μ i T , P , x j molar-based standard chemical potential of the i -solute
ρ o T , P , x α molar/molecular density of the system
ω α orientation of the α -species in the space axes
ζ T , P , x α the correlation length of the mixed-solvent environment
Sub- and super-scripts
f a u l t y property associated with the incorrectly inverted G i j T , P , x j
o pure component or mixed-solvent property
infinite dilution in either an α -solvent or an α + β -mixed solvent
i solute species
I S ideal solution
j solvent species
k cosolvent species
I G ideal gas condition
I G _ i special case of solute as an ideal gas i -species
L R I S Lewis-Randall ideality, i.e., Δ α β T , P , x α = 0
Q stability coefficient in terms of the second composition derivative of the excess Gibbs free energy of the mixed-solvent, i.e., Q = k T D
r isochoric-isothermal residual property

References

  1. Gray, C.G.; Gubbins, K.E. Theory of Molecular Fluids; Oxford University Press: New York, NY, USA, 1985; Volume 1. [Google Scholar]
  2. Chialvo, A.A. On the Solvation Thermodynamics Involving Species with Large Intermolecular Asymmetries: A Rigorous Molecular-Based Approach to Simple Systems with Unconventionally Complex Behaviors. J. Phys. Chem. B 2020, 124, 7879–7896. [Google Scholar] [CrossRef] [PubMed]
  3. Chialvo, A.A. On the Elusive Links between Solution Microstructure, Dynamics, and Solvation Thermodynamics: Demystifying the Path through a Bridge over Troubled Conjectures and Misinterpretations. J. Phys. Chem. B 2023, 127, 10792–10813. [Google Scholar] [CrossRef] [PubMed]
  4. Kirkwood, J.G.; Buff, F.P. The Ststistical Mechanical Theory of Solutions. I. J. Chem. Phys. 1951, 19, 774–777. [Google Scholar] [CrossRef]
  5. Abbott, M.M.; Nass, K.K. Equations of State and Classical Solution Thermodynamics—Survey of the Connections. ACS Symp. Ser. 1986, 300, 2–40. [Google Scholar]
  6. Chialvo, A.A. Accurate Calculation of Excess Thermal, Infinite Dilution, and Related Properties of Liquid Mixtures Via Molecular-Based Simulation. Fluid Phase Equilibria 1993, 83, 23–32. [Google Scholar] [CrossRef]
  7. O’Connell, J.P.; Haile, J.M. Thermodynamics: Fundamentals for Applications; Cambridge University Press: New York, NY, USA, 2005. [Google Scholar]
  8. Chialvo, A.A.; Cummings, P.T. Solute-Induced Effects on the Structure and the Thermodynamics of Infinitely Dilute Mixtures. AIChE J. 1994, 40, 1558–1573. [Google Scholar] [CrossRef]
  9. Chialvo, A.A. Solvation Phenomena in Dilute Solutions: Formal, Experimental Evidence, and Modeling Implications. In Fluctuation Theory of Solutions: Applications in Chemistry, Chemical Engineering and Biophysics; Matteoli, E., O’Connell, J.P., Smith, P.E., Eds.; CRC Press: Boca Raton, FL, USA, 2013; pp. 191–224. [Google Scholar]
  10. Levelt Sengers, J.M.H. Supercritical Fluids: Their Properties and Applications. In Supercritical Fluids: Fundamentals and Applications; Kiran, E., Debenedetti, P.G., Peters, C.J., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2000; pp. 1–29. [Google Scholar]
  11. Debenedetti, P.G.; Mohamed, R.S. Attractive, Weakly Attractive and Repulsive near-Critical Systems. J. Chem. Phys. 1989, 90, 4528–4536. [Google Scholar] [CrossRef]
  12. Petsche, I.B.; Debenedetti, P.G. Influence of Solute-Solvent Asymmetry Upon the Behavior of Dilute Supercritical Mixtures. J. Phys. Chem. 1991, 95, 386–399. [Google Scholar] [CrossRef]
  13. Debenedetti, P.G.; Chialvo, A.A. Solute-Solute Correlations in Infinitely Dilute Supercritical Mixtures. J. Chem. Phys. 1992, 97, 504–507. [Google Scholar] [CrossRef]
  14. O’Connell, J.P.; Liu, H.Q. Thermodynamic Modelling of near-Critical Solutions. Fluid Phase Equilibria 1998, 144, 1–12. [Google Scholar] [CrossRef]
  15. Mazo, R.M. Salting out near the Critical Point. J. Phys. Chem. B 2007, 111, 7288–7290. [Google Scholar] [CrossRef] [PubMed]
  16. Enderby, J.E. Neutron and X-Ray Scattering from Aqueous Solutions. Proc. R. Soc. Lond. A 1975, 345, 107–117. [Google Scholar]
  17. Nishikawa, K.; Iijima, T. Small-Angle X-Rays Scattering Study of Fluctuations in Ethanol and Water Mixtures. J. Phys. Chem. 1993, 97, 10824–10828. [Google Scholar] [CrossRef]
  18. Yamanaka, K.; Yamagami, M.; Takamuku, T.; Yamaguchi, T.; Wakita, H. X-Ray Diffraction Study on Aqueous Lithium Chloride Solution in the Temperature Rnge 138—373k. J. Phys. Chem. 1993, 97, 10835–10839. [Google Scholar] [CrossRef]
  19. Fulton, J.L.; Heald, S.M.; Badyal, Y.S.; Simonson, J.M. Understanding the Effects of Concentration on the Solvation Structure of Ca+2 in Aqueous Solution. I. The Perspective on Local Structure from Exafs and Xanes. J. Phys. Chem. A 2003, 107, 4688–4696. [Google Scholar] [CrossRef]
  20. D’Angelo, P.; Roscioni, O.M.; Chillemi, G.; Della Longa, S.; Benfatto, M. Detection of Second Hydration Shells in Ionic Solutions by Xanes: Computed Spectra for Ni2+ in Water Based on Molecular Dynamics. J. Am. Chem. Soc. 2006, 128, 1853–1858. [Google Scholar] [CrossRef]
  21. Ansell, S.; Barnes, A.C.; Mason, P.E.; Neilson, G.W.; Ramos, S. X-Ray and Neutron Scattering Studies of the Hydration Structure of Alkali Ions in Concentrated Aqueous Solutions. Biophys. Chem. 2006, 124, 171–179. [Google Scholar] [CrossRef]
  22. Antalek, M.; Pace, E.; Hedman, B.; Hodgson, K.O.; Chillemi, G.; Benfatto, M.; Sarangi, R.; Frank, P. Solvation Structure of the Halides from X-Ray Absorption Spectroscopy. J. Chem. Phys. 2016, 145, 044318. [Google Scholar] [CrossRef]
  23. Zitolo, A.; Chillemi, G.; D’Angelo, P. X-Ray Absorption Study of the Solvation Structure of Cu2+ in Methanol and Dimethyl Sulfoxide. Inorg. Chem. 2012, 51, 8827–8833. [Google Scholar] [CrossRef]
  24. Enderby, J.E. Ion Solvation Via Neutron Scattering. Chem. Soc. Rev. 1995, 24, 159–168. [Google Scholar] [CrossRef]
  25. Badyal, Y.S.; Barnes, A.C.; Cuello, G.J.; Simonson, J.M. Understanding the Effects of Concentration on the Solvation Structure of Ca2+ in Aqueous Solutions. Ii: Insights into Longer Range Order from Neutron Diffraction Isotope Substitution. J. Phys. Chem. A 2004, 108, 11819–11827. [Google Scholar] [CrossRef]
  26. Mancinelli, R.; Botti, A.; Bruni, F.; Ricci, M.A.; Soper, A.K. Hydration of Sodium, Potassium, and Chloride Ions in Solution and the Concept of Structure Maker/Breaker. J. Phys. Chem. B 2007, 111, 13570–13577. [Google Scholar] [CrossRef] [PubMed]
  27. Bruni, F.; Imberti, S.; Mancinelli, R.; Ricci, M.A. Aqueous Solutions of Divalent Chlorides: Ions Hydration Shell and Water Structure. J. Chem. Phys. 2012, 136, 064520. [Google Scholar] [CrossRef] [PubMed]
  28. Pusztai, L.; McGreevy, R.L. Mcgr: An Inverse Method for Deriving the Pair Correlation Function from the Structure Factor. Physica B 1997, 234, 357–358. [Google Scholar] [CrossRef]
  29. Pusztai, L. Partial Pair Correlation Functions of Liquid Water. Phys. Rev. B 1999, 60, 11851–11854. [Google Scholar] [CrossRef]
  30. Soper, A.K. Tests of the Empirical Potential Structure Refinement Method and a New Method of Application to Neutron Diffraction Data on Water. Mol. Phys. 2001, 99, 1503–1516. [Google Scholar] [CrossRef]
  31. Soper, A.K. Joint Structure Refinement of X-Ray and Neutron Diffraction Data on Disordered Materials: Application to Liquid Water. J. Phys. Condens. Matter 2007, 19, 335206. [Google Scholar] [CrossRef]
  32. Harsanyi, I.; Pusztai, L. Hydration Structure in Concentrated Aqueous Lithium Chloride Solutions: A Reverse Monte Carlo Based Combination of Molecular Dynamics Simulations and Diffraction Data. J. Chem. Phys. 2012, 137, 204503. [Google Scholar] [CrossRef]
  33. Mile, V.; Gereben, O.; Kohara, S.; Pusztai, L. On the Structure of Aqueous Cesium Fluoride and Cesium Iodide Solutions: Diffraction Experiments, Molecular Dynamics Simulations, and Reverse Monte Carlo Modeling. J. Phys. Chem. B 2012, 116, 9758–9767. [Google Scholar] [CrossRef]
  34. Pethes, I.; Pusztai, L. Reverse Monte Carlo Modeling of Liquid Water with the Explicit Use of the Spc/E Interatomic Potential. J. Chem. Phys. 2017, 146, 064506. [Google Scholar] [CrossRef]
  35. Chialvo, A.A.; Simonson, J.M. The Structure of Concentrated NiCl2 Aqueous Solutions. What Is Molecular Simulation Revealing About the Neutron Scattering Methodologies? Mol. Phys. 2002, 100, 2307–2315. [Google Scholar] [CrossRef]
  36. Chialvo, A.A.; Simonson, J.M. The Structure of CaCl2 Aqueous Solutions over a Wide Range of Concentrations. Interpretation of Difraction Experiments Via Molecular Simulation. J. Chem. Phys. 2003, 119, 8052–8061. [Google Scholar] [CrossRef]
  37. Mason, P.E.; Neilson, G.W.; Dempsey, C.E.; Brady, J.W. Neutron Diffraction and Simulation Studies of CsNo3 and Cs2Co3 Solutions. J. Am. Chem. Soc. 2006, 128, 15136–15144. [Google Scholar] [CrossRef] [PubMed]
  38. Pluharova, E.; Fischer, H.E.; Mason, P.E.; Jungwirth, P. Hydration of the Chloride Ion in Concentrated Aqueous Solutions Using Neutron Scattering and Molecular Dynamics. Mol. Phys. 2014, 112, 1230–1240. [Google Scholar] [CrossRef]
  39. Chialvo, A.A.; Vlcek, L. NO3 Coordination in Aqueous Solutions by 15n/14n and 18o/Nato Isotopic Substitution: What Can We Learn from Molecular Simulation? J. Phys. Chem. B 2015, 119, 519–531. [Google Scholar] [CrossRef]
  40. Kohagen, M.; Pluhařová, E.; Mason, P.E.; Jungwirth, P. Exploring Ion–Ion Interactions in Aqueous Solutions by a Combination of Molecular Dynamics and Neutron Scattering. J. Phys. Chem. Lett. 2015, 6, 1563–1567. [Google Scholar] [CrossRef]
  41. Chialvo, A.A.; Vlcek, L. “Thought Experiments” as Dry-Runs for “Tough Experiments”: Novel Approaches to the Hydration Behavior of Oxyanions. Pure Appl. Chem. 2016, 88, 163–176. [Google Scholar] [CrossRef]
  42. Max, J.J.; Chapados, C. Infrared Spectroscopy of Aqueous Ionic Salt Mixtures at Low Concentrations: Ion Pairing in Water. J. Chem. Phys. 2007, 127, 114509. [Google Scholar] [CrossRef]
  43. Nickolov, Z.S.; Miller, J.D. Water Structure in Aqueous Solutions of Alkali Halide Salts: Ftir Spectroscopy of the Od Stretching Band. J. Colloid Interface Sci. 2005, 287, 572–580. [Google Scholar] [CrossRef]
  44. Dillon, S.R.; Dougherty, R.C. Raman Studies of the Solution Structure of Univalent Electrolytes in Water. J. Phys. Chem. A 2002, 106, 7647–7650. [Google Scholar] [CrossRef]
  45. Smith, J.D.; Saykally, R.J.; Geissler, P.L. The Effects of Dissolved Halide Anions on Hydrogen Bonding in Liquid Water. J. Am. Chem. Soc. 2007, 129, 13847–13856. [Google Scholar] [CrossRef] [PubMed]
  46. Rinne, K.F.; Gekle, S.; Netz, R.R. Ion-Specific Solvation Water Dynamics: Single Water Versus Collective Water Effects. J. Phys. Chem. A 2014, 118, 11667–11677. [Google Scholar] [CrossRef] [PubMed]
  47. Shalit, A.; Ahmed, S.; Savolainen, J.; Hamm, P. Terahertz Echoes Reveal the Inhomogeneity of Aqueous Salt Solutions. Nat. Chem. 2017, 9, 273–278. [Google Scholar] [CrossRef] [PubMed]
  48. Balos, V.; Imoto, S.; Netz, R.R.; Bonn, M.; Bonthuis, D.J.; Nagata, Y.; Hunger, J. Macroscopic Conductivity of Aqueous Electrolyte Solutions Scales with Ultrafast Microscopic Ion Motions. Nat. Commun. 2020, 11, 1611. [Google Scholar] [CrossRef]
  49. Kim, J.S.; Wu, Z.; Morrow, A.R.; Yethiraj, A.; Yethiraj, A. Self-Diffusion and Viscosity in Electrolyte Solutions. J. Phys. Chem. B 2012, 116, 12007–12713. [Google Scholar] [CrossRef]
  50. Ma, K.; Zhao, L. The Opposite Effect of Metal Ions on Short-/Long-Range Water Structure: A Multiple Characterization Study. Int. J. Mol. Sci. 2016, 17, 602. [Google Scholar] [CrossRef]
  51. Tielrooij, K.J.; Garcia-Araez, N.; Bonn, M.; Bakker, H.J. Cooperativity in Ion Hydration. Science 2010, 328, 1006–1009. [Google Scholar] [CrossRef]
  52. Choi, J.H.; Cho, M. Ion Aggregation in High Salt Solutions. Vi. Spectral Graph Analysis of Chaotropic Ion Aggregates. J. Chem. Phys. 2016, 145, 174501. [Google Scholar] [CrossRef]
  53. Bernal, J.D.; Fowler, R.H. A Theory of Water and Ionic Solution, with Particular Reference to Hydrogen and Hydroxyl Ions. J. Chem. Phys. 1933, 1, 515–548. [Google Scholar] [CrossRef]
  54. Frank, H.S.; Evans, M.W. Free Volume and Entropy in Condensed Systems 3. Entropy in Binary Liquid Mixtures—Partial Molal Entropy in Dilute Solutions—Structure and Thermodynamics in Aqueous Electrolytes. J. Chem. Phys. 1945, 13, 507–532. [Google Scholar] [CrossRef]
  55. Ben-Naim, A. Structure-Breaking and Structure-Promoting Processes in Aqueous-Solutions. J. Phys. Chem. 1975, 79, 1268–1274. [Google Scholar] [CrossRef]
  56. Collins, K.D. Sticky Ions in Biological-Systems. Proc. Natl. Acad. Sci. USA 1995, 92, 5553–5557. [Google Scholar] [CrossRef] [PubMed]
  57. Waluyo, I.; Nordlund, D.; Bergmann, U.; Schlesinger, D.; Pettersson, L.G.M.; Nilsson, A. A Different View of Structure-Making and Structure-Breaking in Alkali Halide Aqueous Solutions through X-Ray Absorption Spectroscopy. J. Chem. Phys. 2014, 140, 244506. [Google Scholar] [CrossRef] [PubMed]
  58. Marcus, Y. Effect of Ions on the Structure of Water: Structure Making and Breaking. Chem. Rev. 2009, 109, 1346–1370. [Google Scholar] [CrossRef] [PubMed]
  59. Russo, D. The Impact of Kosmotropes and Chaotropes on Bulk and Hydration Shell Water Dynamics in a Model Peptide Solution. Chem. Phys. 2008, 345, 200–211. [Google Scholar] [CrossRef]
  60. Vranes, M.; Tot, A.; Papovic, S.; Panic, J.; Gadzuric, S. Is Choline Kosmotrope or Chaotrope? J. Chem. Thermodyn. 2018, 124, 65–73. [Google Scholar] [CrossRef]
  61. Ben Ishai, P.; Mamontov, E.; Nickels, J.D.; Sokolov, A.P. Influence of Ions on Water Diffusion-a Neutron Scattering Study. J. Phys. Chem. B 2013, 117, 7724–7728. [Google Scholar] [CrossRef]
  62. Bonetti, M.; Nakamae, S.; Roger, M.; Guenoun, P. Huge Seebeck Coefficients in Nonaqueous Electrolytes. J. Chem. Phys. 2011, 134, 114513. [Google Scholar] [CrossRef]
  63. Ben-Naim, A. Solubility, Hydrophobic Interactions and Structural Changes in the Solvent; Adams, W.A., Ed.; Chemistry and Physics of Aqueous Gas Solutions; Electrochemical Society: Princeton, NJ, USA, 1975. [Google Scholar]
  64. Ball, P.; Hallsworth, J.E. Water Structure and Chaotropicity: Their Uses, Abuses and Biological Implications. Phys. Chem. Chem. Phys. 2015, 17, 8297–8305. [Google Scholar] [CrossRef]
  65. Chialvo, A.A. On the Solute-Induced Structure-Making/Breaking Effect: Rigorous Links among Microscopic Behavior, Solvation Properties, and Solution Non-Ideality. J. Phys. Chem. B 2019, 123, 2930–2947. [Google Scholar] [CrossRef]
  66. Blandamer, M.J.; Burgess, J. Kinetics of Reactions in Aqueous Mixtures. Chem. Soc. Rev. 1975, 4, 55–75. [Google Scholar] [CrossRef]
  67. Marcus, Y. Solvent Mixtures: Properties and Selective Solvation; Taylor & Francis: Clarkesville, GA, USA, 2002. [Google Scholar]
  68. Ivanov, E.V.; Kustov, A.V. Volumetric Properties of (Water Plus Hexamethylphosphoric Triamide) from (288.15 to 308.15) K. J. Chem. Thermodyn. 2010, 42, 1087–1093. [Google Scholar] [CrossRef]
  69. Sedov, I.A.; Magsumov, T.I.; Solomonov, B.N. Solvation of Hydrocarbons in Aqueous-Organic Mixtures. J. Chem. Thermodyn. 2016, 96, 153–160. [Google Scholar] [CrossRef]
  70. Makarov, D.M.; Egorov, G.I.; Kolker, A.M. Volumetric Properties of Binary Liquid Mixtures of Water with N-Methylpyrrolidone at (278.15–323.15) K and up to 70 Mpa. J. Chem. Eng. Data 2022, 67, 1115–1124. [Google Scholar] [CrossRef]
  71. Egorov, G.I.; Makarov, D.M.; Kolker, A.M. Densities and Molar Thermal Expansions of (Water + 1,4-Dioxane) Mixture over the Temperature Range from 274.15 to 333.15 K at Atmospheric Pressure─Comparison with Literature Data. J. Chem. Eng. Data 2022, 67, 3637–3649. [Google Scholar] [CrossRef]
  72. Inoue, H.; Timasheff, S.N. Preferential and Absolute Interactions of Solvent Components with Proteins in Mixed Solvent Systems. Biopolymers 1972, 11, 737–743. [Google Scholar] [CrossRef]
  73. Schellman, J.A. A Simple Model for Solvation in Mixed Solvents. Applications to the Stabilization and Destabilization of Macromolecular Structures. Biophys. Chem. 1990, 37, 121–140. [Google Scholar] [CrossRef]
  74. Poland, D. Ligand-Binding Distributions in Biopolymers. J. Chem. Phys. 2000, 113, 4774–4784. [Google Scholar] [CrossRef]
  75. Tang, K.E.S.; Bloomfield, V.A. Assessing Accumulated Solvent near a Macromolecular Solute by Preferential Interaction Coefficients. Biophys. J. 2002, 82, 2876–2891. [Google Scholar] [CrossRef]
  76. Paulsen, M.D.; Richey, B.; Anderson, C.F.; Record, M.T. The Salt Dependence of the Preferential Interaction Coefficient in DNA Solutions as Determined by Grand-Canonical Monte Carlo Simulations. Chem. Phys. Lett. 1987, 139, 448–452. [Google Scholar]
  77. Strauch, H.J.; Cummings, P.T. Gibbs Ensemble Simulation of Phase Equilibria in Mixed Solvent Electrolytes Systems. Fluid Phase Equilibria 1992, 86, 147–172. [Google Scholar] [CrossRef]
  78. Chialvo, A.A.; Cummings, P.T. Structure of Mixed-Solvent Electrolyte Solutions Via Gibbs Ensemble Monte Carlo Simulations. Mol. Simul. 1993, 11, 163–175. [Google Scholar] [CrossRef]
  79. Chitra, R.; Smith, P.E. Preferential Interactions of Cosolvents with Hydrophobic Solutes. J. Phys. Chem. B 2001, 105, 11513–11522. [Google Scholar] [CrossRef]
  80. Smith, P.E. Cosolvent Interactions with Biomolecules: Relating Computer Simulation Data to Experimental Thermodynamic Data. J. Phys. Chem. B 2004, 108, 18716–18724. [Google Scholar] [CrossRef]
  81. Aburi, M.; Smith, P.E. A Combined Simulation and Kirkwood−Buff Approach to Quantify Cosolvent Effects on the Conformational Preferences of Peptides in Solution. J. Phys. Chem. B 2004, 108, 7382–7388. [Google Scholar] [CrossRef]
  82. Ploetz, E.A.; Karunaweera, S.; Smith, P.E. Kirkwood-Buff-Derived Force Field for Peptides and Proteins: Applications of Kbff20. J. Chem. Theory Comput. 2021, 17, 2991–3009. [Google Scholar] [CrossRef]
  83. Ben-Naim, A. Theory of Preferential Solvation of Non-Electrolytes. Cell Biophys. 1988, 12, 255–269. [Google Scholar] [CrossRef]
  84. Shimizu, S.; Smith, D.J. Preferential Hydration and the Exclusion of Cosolvents from Protein Surfaces. J. Chem. Phys. 2004, 121, 1148–1154. [Google Scholar] [CrossRef]
  85. Shulgin, I.L.; Ruckenstein, E. Preferential Hydration and Solubility of Proteins in Aqueous Solutions of Polyethylene Glycol. Biophys. Chem. 2006, 120, 188–198. [Google Scholar] [CrossRef]
  86. Smiatek, J. Aqueous Ionic Liquids and Their Effects on Protein Structures: An Overview on Recent Theoretical and Experimental Results. J. Phys. Condens. Matter 2017, 29, 233001. [Google Scholar] [CrossRef]
  87. Eisenberg, H. Biological Macromolecules and Polyelectrolytes in Solution; Clarendon Press: Oxford, UK, 1976. [Google Scholar]
  88. Record, M.T.; Anderson, C.F. Interpretation of Preferential Interaction Coefficients of Nonelectrolytes and of Electrolyte Ions in Terms of a Two-Domain Model. Biophys. J. 1995, 68, 786–794. [Google Scholar] [CrossRef] [PubMed]
  89. Timasheff, S.N. Protein Hydration, Thermodynamic Binding, and Preferential Hydration. Biochemistry 2002, 41, 13473–13482. [Google Scholar] [CrossRef] [PubMed]
  90. Schurr, J.M.; Rangel, D.P.; Aragon, S.R. A Contribution to the Theory of Preferential Interaction Coefficients. Biophys. J. 2005, 89, 2258–2276. [Google Scholar] [CrossRef] [PubMed]
  91. Smith, P.E. Equilibrium Dialysis Data and the Relationships between Preferential Interaction Parameters for Biological Systems in Terms of Kirkwood-Buff Integrals. J. Phys. Chem. B 2006, 110, 2862–2868. [Google Scholar] [CrossRef]
  92. Shulgin, I.L.; Ruckenstein, E. A Protein Molecule in a Mixed Solvent: The Preferential Binding Parameter Via the Kirkwood-Buff Theory. Biophys. J. 2006, 90, 704–707. [Google Scholar] [CrossRef]
  93. Chialvo, A.A. Preferential Solvation Phenomena as Solute-Induced Structure-Making/Breaking Processes: Linking Thermodynamic Preferential Interaction Parameters to Fundamental Structure Making/Breaking Functions. J. Phys. Chem. B 2024, 128, 5228–5245. [Google Scholar] [CrossRef]
  94. Marcus, Y. Preferential Solvation of Ions in Mixed-Solvents. 4. Comparison of the Kirkwood-Buff and Quasi-Lattice Quasi-Chemical Approaches. J. Chem. Soc. Faraday Trans. I 1989, 85, 3019–3032. [Google Scholar] [CrossRef]
  95. Matteoli, E.; Lepori, L. Kirkwood-Buff Integrals and Preferential Salvation in Ternary Nonelectrolyte Mixtures. J. Chem. Soc. Faraday Trans. 1995, 91, 431–436. [Google Scholar] [CrossRef]
  96. Smith, P.E.; Mazo, R.A. On the Theory of Solute Solubility in Mixed Solvents. J. Phys. Chem. B 2008, 112, 7875–7884. [Google Scholar] [CrossRef]
  97. Chialvo, A.A. Solute-Solute and Solute-Solvent Correlations in Dilute near-Critical Ternary Mixtures: Mixed Solute and Entrainer Effects. J. Phys. Chem. 1993, 97, 2740–2744. [Google Scholar] [CrossRef]
  98. Chialvo, A.A.; Crisalle, O.D. Osmolyte-Induced Effects on the Hydration Behavior and the Osmotic Second Virial Coefficients of Alkyl-Substituted Urea Derivatives. Critical Assessment of Their Structure-Making/Breaking Behavior. J. Phys. Chem. B 2021, 125, 6231–6243. [Google Scholar] [CrossRef] [PubMed]
  99. Zhang, W.T.; Capp, M.W.; Bond, J.P.; Anderson, C.F.; Record, M.T. 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 1996, 35, 10506–10516. [Google Scholar] [CrossRef] [PubMed]
  100. Courtenay, E.S.; Capp, M.W.; Record, M.T. Thermodynamics of Interactions of Urea and Guanidinium Salts with Protein Surface: Relationship between Solute Effects on Protein Processes and Changes in Water-Accessible Surface Area. Protein Sci. 2001, 10, 2485–2497. [Google Scholar] [CrossRef] [PubMed]
  101. Hade, E.P.K.; Tanford, C. Isopiestic Compositions as a Measure of Preferential Interactions of Macromolecules in 2-Component Solvents.Application to Proteins in Concentrated Aqueous Cesium Chloride and Guanidine Hydrochloride. J. Am. Chem. Soc. 1967, 89, 5034–5040. [Google Scholar] [CrossRef] [PubMed]
  102. Hearst, J.E. Determination of Dominant Factors Which Influence Net Hydration of Native Sodium Deoxyribonucleate. Biopolymers 1965, 3, 57–68. [Google Scholar] [CrossRef]
  103. Braunlin, W.H.; Strick, T.J.; Record, M.T. Equilibrium Dialysis Studies of Polyamine Binding to DNA. Biopolymers 1982, 21, 1301–1314. [Google Scholar] [CrossRef] [PubMed]
  104. Record, M.T.; Zhang, W.T.; Anderson, C.F. 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. In Advances in Protein Chemistry, Vol 51: Linkage Thermodynamics of Macromolecular Interactions; DiCera, E., Ed.; Academic Press: Cambridge, MA, USA, 1998; pp. 281–353. [Google Scholar]
  105. Courtenay, E.S.; Capp, M.W.; Anderson, C.F.; Record, M.T. Vapor Pressure Osmometry Studies of Osmolyte-Protein Interactions: Implications for the Action of Osmoprotectants in Vivo and for the Interpretation of “Osmotic Stress” Experiments in Vitro. Biochemistry 2000, 39, 4455–4471. [Google Scholar] [CrossRef]
  106. Anderson, C.F.; Felitsky, D.J.; Hong, J.; Record, M.T. Generalized Derivation of an Exact Relationship Linking Different Coefficients That Characterize Thermodynamic Effects of Preferential Interactions. Biophys. Chem. 2002, 101–102, 497–511. [Google Scholar] [CrossRef]
  107. Shimizu, S.; Matubayasi, N. Preferential Hydration of Proteins: A Kirkwood-Buff Approach. Chem. Phys. Lett. 2006, 420, 518–522. [Google Scholar] [CrossRef]
  108. Weerasinghe, S.; Smith, P.E. Cavity Formation and Preferential Interactions in Urea Solutions: Dependence on Urea Aggregation. J. Chem. Phys. 2003, 118, 5901–5910. [Google Scholar] [CrossRef]
  109. Shulgin, I.L.; Ruckenstein, E. Relationship between Preferential Interaction of a Protein in an Aqueous Mixed Solvent and Its Solubility. Biophys. Chem. 2005, 118, 128–134. [Google Scholar] [CrossRef] [PubMed]
  110. Pallewela, G.N.; Smith, P.E. Preferential Solvation in Binary and Ternary Mixtures. J. Phys. Chem. B 2015, 119, 15706–15717. [Google Scholar] [CrossRef] [PubMed]
  111. Smith, P.E.; O’Connell, J.P.; Matteoli, E. Fluctuation Theory of Solutions: Applications in Chemistry, Chemical Engineering and Biophysics; CRC Press: Boca Raton, FL, USA, 2013. [Google Scholar]
  112. Ben-Naim, A. Molecular Theory of Solutions; Oxford University Press: Oxford, NC, USA, 2006. [Google Scholar]
  113. Chialvo, A.A.; Crisalle, O.D. Gas Solubility and Preferential Solvation Phenomena in Mixed-Solvents; Rigorous Relations between Microscopic Behavior, Solvation Properties, and Solution Non-Ideality. Fluid Phase Equilibria 2024, 581, 114081. [Google Scholar] [CrossRef]
  114. Ben-Naim, A. Solvation Thermodynamics; Plenum Press: New York, NY, USA, 1987. [Google Scholar]
  115. Chialvo, A.A. Gas Solubility in Dilute Solutions: A Novel Molecular Thermodynamic Perspective. J. Chem. Phys. 2018, 148, 174502. [Google Scholar] [CrossRef] [PubMed]
  116. Chialvo, A.A.; Crisalle, O.D. Solvent and H/D Isotopic Substitution Effects on the Krichevskii Parameter of Solutes: A Novel Approach to Their Accurate Determination. Liquids 2022, 2, 474–503. [Google Scholar] [CrossRef]
  117. Wilhelm, E.; Battino, R.; Wilcock, R.J. Low-Pressure Solubility of Gases in Liquid Water. Chem. Rev. 1977, 77, 219–262. [Google Scholar] [CrossRef]
  118. Mainar, A.M.; Pardo, J.; Royo, F.M.; Lopez, M.C.; Urieta, J.S. Solubility of Nonpolar Gases in 2,2,2-Trifluoroethanol at 25 Degrees C and 101.33 Kpa Partial Pressure of Gas. J. Solut. Chem. 1996, 25, 589–595. [Google Scholar] [CrossRef]
  119. Mainar, A.M.; Pardo, J.I.; Santafe, J.; Urieta, J.S. Solubility of Gases in Binary Liquid Mixtures: An Experimental and Theoretical Study of the System Noble Gas Plus Trifluoroethanol Plus Water. Ind. Eng. Chem. Res. 2003, 42, 1439–1450. [Google Scholar] [CrossRef]
  120. Mainar, A.M.; Martinez-Lopez, J.F.; Langa, E.; Pardo, J.I.; Urieta, J.S. Solubilities of Several Non-Polar Gases in Mixtures Water+2,2,2-Trifluoroethanol at 298.15 K and 101.33 Kpa. Fluid Phase Equilibria 2012, 314, 161–168. [Google Scholar] [CrossRef]
  121. Sassi, M.; Atik, Z. Excess Molar Volumes of Binary Mixtures of 2,2,2-Trifluoroethanol with Water, or Acetone, or 1,4-Difluorobenzene, or 4-Fluorotoluene, or Alpha, Alpha, Alpha, Trifluorotoluene or 1-Alcohols at a Temperature of 298.15 K and Pressure of 101 Kpa. J. Chem. Thermodyn. 2003, 35, 1161–1169. [Google Scholar] [CrossRef]
  122. Matsuo, S.; Yamamoto, R.; Kubota, H.; Tanaka, Y. Volumetric Properties of Mixtures of Fluoroalcohols and Water at High Pressures. Int. J. Thermophys. 1994, 15, 245–259. [Google Scholar] [CrossRef]
  123. Cooney, A.; Morcom, K.W. Thermodynamic Behavior of Mixtures Containing Fluoroalcohols.1. (Water+2,2,2-Trifluoroethanol). J. Chem. Thermodyn. 1988, 20, 735–741. [Google Scholar] [CrossRef]
  124. Krichevskii, I.R. Thermodynamics of an Infinitely Dilute Solution in Mixed Solvents. I. The Henry’s Coefficient in a Mixed Solvent Behaving as an Ideal Solvent. Zhournal Fiz. Khimii 1937, 9, 41–47. [Google Scholar]
  125. O’Connell, J.P. Molecular Thermodynamics of Gases in Mixed Solvents. AIChE J. 1971, 17, 658–663. [Google Scholar] [CrossRef]
  126. Mazo, R.M. Statistical Mechanical Theory of Solutions. J. Chem. Phys. 1958, 29, 1122–1128. [Google Scholar] [CrossRef]
  127. Chialvo, A.A. Alternative Approach to Modeling Excess Gibbs Free Energy in Terms of Kirkwood-Buff Integrals. In Advances in Thermodynamics; Matteoli, E., Mansoori, G.A., Eds.; Taylor & Francis: New York, NY, USA, 1990; pp. 131–173. [Google Scholar]
  128. Rubino, J.T. Cosolvents and Cosolvency. In Encyclopedia of Pharmaceutical Technology; Swarbrick, J., Ed.; CRC Press: Boca Raton, FL, USA, 2006; pp. 806–819. [Google Scholar]
  129. Myrdal, P.B.; Yalkowsky, S.H. Solubilization of Drugs in Aqueous Media. In Encyclopedia of Pharmaceutical Technology; Swarbrick, J., Ed.; CRC Press: Boca Raton, FL, USA, 2006; pp. 3311–3333. [Google Scholar]
  130. Kawashima, Y.; Imai, M.; Takeuchi, H.; Yamamoto, H.; Kamiya, K. Hi Improved Flowability and Compactibility of Spherically Agglomerated Crystals of Ascorbic Acid for Direct Tableting Designed by Spherical Crystallization Process. Powder Technol. 2003, 130, 283–289. [Google Scholar] [CrossRef]
  131. Scherzinger, C.; Schwarz, A.; Bardow, A.; Leonhard, K.; Richtering, W. Cononsolvency of Poly-N-Isopropyl Acrylamide (Pnipam): Microgels Versus Linear Chains and Macrogels. Curr. Opin. Colloid Interface Sci. 2014, 19, 84–94. [Google Scholar] [CrossRef]
  132. Ren, F.; Zhou, Y.; Liu, Y.; Fu, J.; Jing, Q.; Ren, G. A Mixed Solvent System for Preparation of Spherically Agglomerated Crystals of Ascorbic Acid. Pharm. Dev. Technol. 2016, 22, 818–826. [Google Scholar] [CrossRef]
  133. Chialvo, A.A. Preferential Solvation in Pharmaceutical Processing: Rigorous Results, Critical Observations, and the Unraveling of Some Significant Modeling Pitfalls. Fluid Phase Equilibria 2024, 587, 114212. [Google Scholar] [CrossRef]
  134. Rodríguez, G.A.; Delgado, D.R.; Martínez, F. Preferential Solvation of Indomethacin and Naproxen in Ethyl Acetate Plus Ethanol Mixtures According to the Ikbi Method. Phys. Chem. Liq. 2014, 52, 533–545. [Google Scholar] [CrossRef]
  135. Cristancho, D.M.; Martínez, F. Solubility and Preferential Solvation of Meloxicam in Ethyl Acetate Plus Ethanol Mixtures at Several Temperatures. J. Mol. Liq. 2014, 200, 122–128. [Google Scholar] [CrossRef]
  136. Cristancho, D.M.; Jouyban, A.; Martínez, F. Solubility, Solution Thermodynamics, and Preferential Solvation of Piroxicam in Ethyl Acetate Plus Ethanol Mixtures. J. Mol. Liq. 2016, 221, 72–81. [Google Scholar] [CrossRef]
  137. Guoquan, Z. Study of Solute-Solvent Intermolecular Interactions and Preferential Solvation for Mevastatin Dissolution in Pure and Mixed Binary Solvents. J. Chem. Thermodyn. 2022, 175, 106884. [Google Scholar] [CrossRef]
  138. Du, C.B.; Cong, Y.; Wang, M.; Jiang, Z.Y.; Wang, M.L. Preferential Solvation and Solute-Solvent Interactions of Posaconazole in Mixtures of (Ethyl Acetate Plus Ethanol/Isopropanol) at Several Temperatures. J. Chem. Thermodyn. 2022, 165, 106661. [Google Scholar] [CrossRef]
  139. Mayo Clinic. Drugs & Supplements. Available online: https://www.mayoclinic.org/drugs-supplements (accessed on 7 May 2024).
  140. Baynes, B.M.; Trout, B.L. Proteins in Mixed Solvents: A Molecular-Level Perspective. J. Phys. Chem. B 2003, 107, 14058–14067. [Google Scholar] [CrossRef]
  141. Vagenende, V.; Trout, B.L. Quantitative Characterization of Local Protein Solvation to Predict Solvent Effects on Protein Structure. Biophys. J. 2012, 103, 1354–1362. [Google Scholar] [CrossRef]
  142. Ganguly, P.; Bubák, D.; Polák, J.; Fagan, P.; Dračínský, M.; van der Vegt, N.F.A.; Heyda, J.; Shea, J.-E. Cosolvent Exclusion Drives Protein Stability in Trimethylamine N-Oxide and Betaine Solutions. J. Phys. Chem. Lett. 2022, 13, 7980–7986. [Google Scholar] [CrossRef]
  143. Miner, J.C.; García, A.E. Equilibrium Denaturation and Preferential Interactions of an Rna Tetraloop with Urea. J. Phys. Chem. B 2017, 121, 3734–3746. [Google Scholar] [CrossRef]
  144. Kang, M.; Smith, P.E. Preferential Interaction Parameters in Biological Systems by Kirkwood-Buff Theory and Computer Simulation. Fluid Phase Equilibria 2007, 256, 14–19. [Google Scholar] [CrossRef]
  145. Karunaweera, S.; Gee, M.B.; Weerasinghe, S.; Smith, P.E. Theory and Simulation of Multicomponent Osmotic Systems. J. Chem. Theory Comput. 2012, 8, 3493–3503. [Google Scholar] [CrossRef]
  146. Ploetz, E.A.; Karunaweera, S.; Bentenitis, N.; Chen, F.; Dai, S.; Gee, M.B.; Jiao, Y.; Kang, M.; Kariyawasam, N.L.; Naleem, N.; et al. Kirkwood-Buff-Derived Force Field for Peptides and Proteins: Philosophy and Development of Kbff20. J. Chem. Theory Comput. 2021, 17, 2964–2990. [Google Scholar] [CrossRef]
  147. Heidari, M.; Kremer, K.; Potestio, R.; Cortes-Huerto, R. Finite-Size Integral Equations in the Theory of Liquids and the Thermodynamic Limit in Computer Simulations. Mol. Phys. 2018, 116, 3301–3310. [Google Scholar] [CrossRef]
  148. Milzetti, J.; Nayar, D.; van der Vegt, N.F.A. Convergence of Kirkwood–Buff Integrals of Ideal and Nonideal Aqueous Solutions Using Molecular Dynamics Simulations. J. Phys. Chem. B 2018, 122, 5515–5526. [Google Scholar] [CrossRef] [PubMed]
  149. Cloutier, T.; Sudrik, C.; Sathish, H.A.; Trout, B.L. Kirkwood-Buff-Derived Alcohol Parameters for Aqueous Carbohydrates and Their Application to Preferential Interaction Coefficient Calculations of Proteins. J. Phys. Chem. B 2018, 122, 9350–9360. [Google Scholar] [CrossRef] [PubMed]
  150. Chéron, N.; Naepels, M.; Pluharová, E.; Laage, D. Protein Preferential Solvation in Water:Glycerol Mixtures. J. Phys. Chem. B 2020, 124, 1424–1437. [Google Scholar] [CrossRef] [PubMed]
  151. Vagenende, V.; Yap, M.G.S.; Trout, B.L. Molecular Anatomy of Preferential Interaction Coefficients by Elucidating Protein Solvation in Mixed Solvents: Methodology and Application for Lysozyme in Aqueous Glycerol. J. Phys. Chem. B 2009, 113, 11743–11753. [Google Scholar] [CrossRef]
  152. Marcus, Y. Preferential Solvation in Mixed Solvents X. Completely Miscible Aqueous Co-Solvent Binary Mixtures at 298.15 K. Monatshefte Fur Chem. 2001, 132, 1387–1411. [Google Scholar] [CrossRef]
  153. Marcus, Y. On the Preferential Solvation of Drugs and Pahs in Binary Solvent Mixtures. J. Mol. Liq. 2008, 140, 61–67. [Google Scholar] [CrossRef]
  154. Marcus, Y. Ions in Water and Biophysical Implications: From Chaos to Cosmos; Springer: Dordrecht, The Netherlands, 2012. [Google Scholar]
  155. Marcus, Y. Ions in Solution and Their Solvation; John Wiley and Sons: Hoboken, NJ, USA, 2015. [Google Scholar]
  156. Matteoli, E. A Study on Kirkwood-Buff Integrals and Preferential Solvation in Mixtures with Small Deviations from Ideality and/or with Size Mismatch of Components. Importance of a Proper Reference System. J. Phys. Chem. B 1997, 101, 9800–9810. [Google Scholar] [CrossRef]
  157. Marcus, Y. Preferential Solvation in Mixed Solvents 13. Mixtures of Tetrahydrofuran with Organic Solvents: Kirkwood-Buff Integrals and Volume-Corrected Preferential Solvation Parameters. J. Solut. Chem. 2006, 35, 251–277. [Google Scholar] [CrossRef]
  158. Marcus, Y. The Structure of and Interactions in Binary Acetonitrile Plus Water Mixtures. J. Phys. Org. Chem. 2012, 25, 1072–1085. [Google Scholar] [CrossRef]
  159. Ben-Naim, A. A Critique of Some Recent Suggestions to Correct the Kirkwood-Buff Integrals. J. Phys. Chem. B 2007, 111, 2896–2902. [Google Scholar] [CrossRef] [PubMed]
  160. Ben-Naim, A. Comment on the “Kirkwood-Buff Theory of Solutions and the Local Composition of Liquid Mixtures”. J. Phys. Chem. B 2008, 112, 5874–5875. [Google Scholar] [CrossRef] [PubMed]
  161. Li, X.; Ma, M.; Chen, J.; Chen, G.; Zhao, H. Preferential Solvation of Boscalid in Ethanol/Isopropanol + ethyl Acetate Mixtures from the Inverse Kirkwood–Buff Integrals Method. J. Solut. Chem. 2017, 46, 2050–2065. [Google Scholar] [CrossRef]
  162. Zhu, Y.Q.; Cheng, C.; Zhao, H.K. Solubility and Preferential Solvation of Carbazochrome in Solvent Mixtures of N, N-Dimethylformamide Plus Methanol/Ethanol N-Propanol and Dimethyl Sulfoxide Plus Water. J. Chem. Eng. Data 2018, 63, 822–831. [Google Scholar] [CrossRef]
  163. Zheng, M.; Chen, J.; Xu, R.J.; Chen, G.Q.; Cong, Y.; Zhao, H.K. Solubility and Preferential Solvation of 3-Nitrobenzonitrile in Binary Solvent Mixtures of Ethyl Acetate Plus (Methanol, Ethanol, n-Propanol, and Isopropyl Alcohol). J. Chem. Eng. Data 2018, 63, 2290–2298. [Google Scholar] [CrossRef]
  164. Romdhani, A.; Osorio, I.P.; Martínez, F.; Jouyban, A.; Acree, W.E. Further Calculations on the Solubility of Trans-Resveratrol in (Transcutol® Plus Water) Mixtures. J. Mol. Liq. 2021, 330, 115645. [Google Scholar] [CrossRef]
  165. Osorio, I.P.; Martínez, F.; Peña, M.A.; Jouyban, A.; Acree, W.E. Solubility of Sulphadiazine in Some {Carbitol® (1) + Water (2)} Mixtures: Determination, Correlation, and Preferential Solvation. Phys. Chem. Liq. 2021, 59, 890–906. [Google Scholar] [CrossRef]
  166. Noubigh, A.; Tahar, L.B.; Eladeb, A. Solubility Modeling and Preferential Solvation of Benzamide in Some Pure and Binary Solvent Mixtures at Different Temperatures. J. Chem. Eng. Data 2023, 68, 1018–1130. [Google Scholar] [CrossRef]
  167. Zhao, X.; Farajtabar, A.; Han, G.; Zhao, H.H. Griseofulvin Dissolved in Binary Aqueous Co-Solvent Mixtures of N, N-Dimethylformamide, Methanol, Ethanol, Acetonitrile and N-Methylpyrrolidone: Solubility Determination and Thermodynamic Studies. J. Chem. Thermodyn. 2020, 151, 106250. [Google Scholar] [CrossRef]
  168. Alshehri, S.; Shakeel, F.; Alam, P.; Jouyban, A.; Martinez, F. Solubility of 6-Phenyl-4,5-Dihydropyridazin-3(2H)-One in Aqueous Mixtures of Transcutol and Peg 400 Revisited: Correlation and Preferential Solvation. J. Mol. Liq. 2021, 344, 117728. [Google Scholar] [CrossRef]
  169. Noubigh, A.; Abderrabba, M. Preferential Solvation of 5,7-Dihydroxyflavone (Chrysin) in Aqueous Co-Solvent Mixtures of Methanol and Ethanol. Phys. Chem. Liq. 2022, 60, 931–942. [Google Scholar] [CrossRef]
  170. Guo, Q.R.; Shi, W.Z.; Zhao, H.K.; Li, W.X.; Han, G.; Farajtabar, A. Solubility, Solvent Effect, Preferential Solvation and Dft Computations of 5-Nitrosalicylic Acid in Several Aqueous Blends. J. Chem. Thermodyn. 2023, 177, 106936. [Google Scholar] [CrossRef]
  171. Gao, Q.; Farajtabar, A. Glyburide in a Series Co-Solvent Solutions: Solubility and Modeling, Solvation and Quantum Chemistry Research. J. Chem. Thermodyn. 2024, 189, 107195. [Google Scholar] [CrossRef]
Figure 1. Isobaric–isothermal composition dependence of the experimentally based total correlation function integrals G α β T , P , x j and corresponding Δ j k T , P , x j G j j + G k k 2 G j k linear combination, where we highlighted the Lewis-Randal ideality condition Δ j k T , P , x j = 0 , for j = TFE and k = H 2 O mixed-solvent environment.
Figure 1. Isobaric–isothermal composition dependence of the experimentally based total correlation function integrals G α β T , P , x j and corresponding Δ j k T , P , x j G j j + G k k 2 G j k linear combination, where we highlighted the Lewis-Randal ideality condition Δ j k T , P , x j = 0 , for j = TFE and k = H 2 O mixed-solvent environment.
Thermo 04 00022 g001
Figure 2. Isobaric-isothermal composition dependence of (a) the experimentally based diffusive stability coefficient D T , P , x j for the j + k -binary mixed-solvent environment, in comparison with the Lewis-Randal ideality condition described by Δ j k T , P , x j = 0 , and (b) the two fundamental structure-making/-breaking functions  S j k T , P , x j and S k j T , P , x j , with the highlight of the condition S α β T , P , x j = 0 identifying the border between structure-making and structure-breaking behaviors for j = TFE and k = H 2 O mixed-solvent environment.
Figure 2. Isobaric-isothermal composition dependence of (a) the experimentally based diffusive stability coefficient D T , P , x j for the j + k -binary mixed-solvent environment, in comparison with the Lewis-Randal ideality condition described by Δ j k T , P , x j = 0 , and (b) the two fundamental structure-making/-breaking functions  S j k T , P , x j and S k j T , P , x j , with the highlight of the condition S α β T , P , x j = 0 identifying the border between structure-making and structure-breaking behaviors for j = TFE and k = H 2 O mixed-solvent environment.
Thermo 04 00022 g002
Figure 3. Isobaric-isothermal composition dependence of the fundamental structure-making/-breaking functions (a) S i   T F E T , P , x T F E and (b) S i   H 2 O T , P , x T F E , for the fourteen i -solutes in the TFE + H 2 O -binary solvent environment. Note the highlighted conditions S i   T F E T , P , x T F E = 0 and S i   H 2 O T , P , x T F E = 0 which identify the borders between structure-making and structure-breaking behaviors.
Figure 3. Isobaric-isothermal composition dependence of the fundamental structure-making/-breaking functions (a) S i   T F E T , P , x T F E and (b) S i   H 2 O T , P , x T F E , for the fourteen i -solutes in the TFE + H 2 O -binary solvent environment. Note the highlighted conditions S i   T F E T , P , x T F E = 0 and S i   H 2 O T , P , x T F E = 0 which identify the borders between structure-making and structure-breaking behaviors.
Thermo 04 00022 g003
Figure 4. Isobaric-isothermal composition dependence of (a) the universal preferential solvation function  P S T , P , x T F E and (b) the rate of change in the i -solute’s standard chemical potential upon perturbation of the chemical potential of the TFE -species for the fourteen i -solutes in the TFE + H 2 O -binary solvent environment.
Figure 4. Isobaric-isothermal composition dependence of (a) the universal preferential solvation function  P S T , P , x T F E and (b) the rate of change in the i -solute’s standard chemical potential upon perturbation of the chemical potential of the TFE -species for the fourteen i -solutes in the TFE + H 2 O -binary solvent environment.
Thermo 04 00022 g004
Figure 5. Preferential solvation behavior of six pharmaceutical solutes in a common mixed ethyl acetate (j) + ethanol (k) environment described by (a) the universal preferential solvation  P S x E T A C and; (b) the corresponding first-order preferential solvation parameter δ E T A C   i 0 x E T A C along the T = 313   K isotherm under ambient pressure.
Figure 5. Preferential solvation behavior of six pharmaceutical solutes in a common mixed ethyl acetate (j) + ethanol (k) environment described by (a) the universal preferential solvation  P S x E T A C and; (b) the corresponding first-order preferential solvation parameter δ E T A C   i 0 x E T A C along the T = 313   K isotherm under ambient pressure.
Thermo 04 00022 g005
Figure 6. Preferential solvation behavior of six pharmaceutical solutes in a common mixed ethyl acetate (j) + ethanol (k) environment described by (a) the faulty first-order parameter ρ o δ E T A C   i 0 , f a u l t y x E T A C and (b) the corresponding color-coordinated proper first-order parameter ρ o δ E T A C   i 0 x E T A C along the T = 313   K isotherm under ambient pressure.
Figure 6. Preferential solvation behavior of six pharmaceutical solutes in a common mixed ethyl acetate (j) + ethanol (k) environment described by (a) the faulty first-order parameter ρ o δ E T A C   i 0 , f a u l t y x E T A C and (b) the corresponding color-coordinated proper first-order parameter ρ o δ E T A C   i 0 x E T A C along the T = 313   K isotherm under ambient pressure.
Thermo 04 00022 g006
Table 1. Relations between the thermodynamic preferential interaction parameters and the Kirkwood–Buff integrals according to the nature of the system [91].
Table 1. Relations between the thermodynamic preferential interaction parameters and the Kirkwood–Buff integrals according to the nature of the system [91].
S y s t e m ρ m x
Γ μ j μ k ( χ i 0 ) G α β  1 ρ k G i k ρ k G i k G i j x k + x j ρ k G i k G i j
Γ μ j ( χ i 0 ) G α β  2 1 + ρ k G i k G k k 1 + ρ k G i k + G k j G i j G k k 1 + x j ρ k G i k G i j + G j k G k k
Γ μ k ( χ i 0 ) G α β  2 ρ k G i k G j k ρ k / ρ j + ρ k G i k G i j + G j j G j k x j ρ k G i k G i j + G j j G j k
1 open, 2 semi-open.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chialvo, A.A. Linking Solution Microstructure and Solvation Thermodynamics of Mixed-Solvent Systems: Formal Results, Critical Observations, and Modeling Pitfalls. Thermo 2024, 4, 407-432. https://doi.org/10.3390/thermo4030022

AMA Style

Chialvo AA. Linking Solution Microstructure and Solvation Thermodynamics of Mixed-Solvent Systems: Formal Results, Critical Observations, and Modeling Pitfalls. Thermo. 2024; 4(3):407-432. https://doi.org/10.3390/thermo4030022

Chicago/Turabian Style

Chialvo, Ariel A. 2024. "Linking Solution Microstructure and Solvation Thermodynamics of Mixed-Solvent Systems: Formal Results, Critical Observations, and Modeling Pitfalls" Thermo 4, no. 3: 407-432. https://doi.org/10.3390/thermo4030022

Article Metrics

Back to TopTop