Fact-checked by Grok 2 weeks ago

Activity coefficient

In , the activity coefficient is a dimensionless factor that quantifies the departure of a real mixture's thermodynamic properties from those of an ideal mixture, relating the of a component to its composition through the equation \mu_i = \mu_i^\circ + RT \ln(\gamma_i x_i), where \mu_i is the , \mu_i^\circ is the standard chemical potential, R is the , T is temperature, \gamma_i is the activity coefficient, and x_i is the mole fraction. For a component i in a solution, the activity a_i is defined as a_i = \gamma_i x_i (or \gamma_i c_i for concentration-based conventions), where \gamma_i = 1 indicates ideal behavior and deviations arise from intermolecular interactions. Activity coefficients are crucial for accurately predicting and modeling non-ideal behaviors in both liquid and gas mixtures, particularly in solutions where ionic interactions significantly affect properties like and reaction equilibria. In systems, they correct for the difference between actual concentrations and their effective thermodynamic activities, as expressed by a_X = \gamma_X [X], enabling the use of concentrations in expressions K = \prod (a_X)^{v_X} when \gamma values are known or approximated. For nonelectrolyte mixtures, activity coefficients derive from the excess via \Delta G^E = RT \sum n_i \ln \gamma_i, ensuring thermodynamic consistency through relations like the Gibbs-Duhem equation. Common models for estimating activity coefficients include the Debye-Hückel equation for dilute electrolytes, \log \gamma_i = -0.51 z_i^2 \sqrt{\mu} / (1 + 0.33 \alpha \sqrt{\mu}), where z_i is ion charge, \mu is , and \alpha is an , as well as activity coefficient models like Margules, van Laar, and for multicomponent systems. These models are fitted to experimental data from phase equilibria, such as vapor-liquid measurements, to predict behaviors in processes like and . Overall, activity coefficients bridge ideal approximations and real-world applications in and , with values typically ranging from less than 1 (attractive interactions) to greater than 1 (repulsive interactions).

Definition and Thermodynamic Foundations

General Thermodynamic Definition

In thermodynamics, ideal solutions are characterized by the absence of intermolecular interactions beyond those in the pure components, leading to additive enthalpies of mixing and entropies that follow the ideal entropy of mixing. For such systems, the partial vapor pressure of a solvent follows Raoult's law, expressed as P_i = x_i P_i^\circ, where P_i is the partial pressure of component i, x_i is its mole fraction, and P_i^\circ is the vapor pressure of the pure component. For solutes in ideal dilute solutions, Henry's law applies, P_i = k_H x_i, where k_H is the Henry's law constant specific to the solute-solvent pair. Real solutions, however, exhibit non-ideal behavior due to interactions such as solute-solvent attractions or repulsions, resulting in deviations from these laws and necessitating correction factors to maintain thermodynamic consistency. The concept of the activity coefficient was introduced by in 1907 as a means to extend the theory to real systems by accounting for these deviations in a thermodynamically rigorous manner. Lewis proposed the activity a_i of a component i as an effective concentration that preserves the form of ideal equations while incorporating non-ideality, defined such that the \mu_i follows the relation \mu_i = \mu_i^0 + RT \ln a_i, where \mu_i^0 is the standard chemical potential, R is the , and T is the . To derive the activity coefficient \gamma_i, start from the general expression for the in a , which must reduce to the ideal form in the limit of . For the in a binary mixture, the ideal chemical potential is \mu_1^\text{ideal} = \mu_1^\circ + [RT](/page/RT) \ln x_1, corresponding to where the equals x_1 f_1^\circ (with f_1^\circ as the fugacity of pure ). In non-ideal cases, the fugacity is f_1 = \gamma_1 x_1 f_1^\circ, so \mu_1 = \mu_1^\circ + [RT](/page/RT) \ln (\gamma_1 x_1), introducing \gamma_1 as the activity coefficient that captures the deviation (\gamma_1 = 1 for ). For a solute, the is often taken at infinite dilution, where defines the reference fugacity f_2^\circ = k_H, leading to \mu_2 = \mu_2^\circ + [RT](/page/RT) \ln (\gamma_2 x_2), with \lim_{x_2 \to 0} \gamma_2 = 1 by convention, ensuring the activity a_2 = \gamma_2 x_2 approaches the ideal dilute behavior. This framework applies generally to multicomponent systems, where each \gamma_i adjusts the to reflect excess interactions. The activity coefficients are thermodynamically linked to the excess Gibbs free energy of mixing G^E, which quantifies the non-ideal contribution to the total Gibbs energy. For a multicomponent solution, this relation is given by G^E = RT \sum_i x_i \ln \gamma_i, where the sum is over all components, and G^E = 0 for ideal solutions since \gamma_i = 1 for all i. This equation derives from the partial molar excess Gibbs energy \bar{G}_i^E = RT \ln \gamma_i, integrated over the composition via the Gibbs-Duhem relation. In multicomponent systems, particularly those involving dissociated species, a mean activity coefficient \gamma_\pm is often defined to represent the collective non-ideal behavior, such as for a dissociating into \nu_+ cations and \nu_- anions, where \gamma_\pm = (\gamma_+^{\nu_+} \gamma_-^{\nu_-})^{1/\nu} with \nu = \nu_+ + \nu_-, ensuring the mean activity a_\pm = \gamma_\pm m^\nu (on a scale) aligns with measurable properties like .

Activity in Non-Electrolyte Solutions

In non-electrolyte solutions, which consist of neutral molecular species without significant ionic contributions, the activity coefficient \gamma_i corrects for deviations from ideal behavior due to short-range intermolecular interactions such as van der Waals forces, hydrogen bonding, and differences. For volatile components in these mixtures, the partial vapor pressure P_i is described by the modified : P_i = \gamma_i x_i P_i^\ast, where x_i is the of component i in the liquid phase, and P_i^\ast is the of the pure component i at the same ./16%3A_The_Chemical_Activity_of_the_Components_of_a_Solution/16.03%3A_Expressing_the_Activity_Coefficient_as_a_Deviation_from_Raoult%27s_Law) This formulation applies to both solvents and solutes, with \gamma_i = 1 indicating ideal behavior and \gamma_i \neq 1 reflecting non-idealities; for non-volatile solutes, activity is often referenced to , but the focus here remains on molecular mixtures like binary liquids. In systems such as alcohol-water mixtures, significant positive deviations occur due to differing polarities and hydrogen-bonding capabilities, leading to \gamma > 1. For ethanol-water at 25°C, the activity coefficient of ethanol varies from approximately 4–5 at low ethanol mole fractions (indicating reduced solubility and higher fugacity) to near 1 at equimolar compositions, highlighting a maximum azeotrope formation. Similarly, in hydrocarbon blends like benzene-cyclohexane, interactions are more similar, resulting in \gamma values close to 1 across compositions, with minor positive deviations (\gamma \approx 1.05–1.1) arising from subtle differences in chain length or branching. These examples illustrate how \gamma_i quantifies the enhanced volatility or immiscibility in non-ideal non-electrolyte systems. The activity coefficients in non-electrolyte solutions are thermodynamically linked to excess properties, particularly the molar excess G^E, through the relation \frac{G^E}{RT} = \sum_i x_i \ln \gamma_i, where R is the and T is . This derives from integrating the Gibbs-Duhem relation over , expressing non-ideality as the difference between the actual of mixing and the entropic contribution./24%3A_Solutions_I_-_Volatile_Solutes/24.09%3A_Gibbs_Energy_of_Mixing_of_Binary_Solutions_in_Terms_of_the_Activity_Coefficient) Positive G^E corresponds to \gamma_i > 1 for limited , while negative values indicate \gamma_i < 1 for favorable interactions, as seen in some polar-organic solvent pairs. A common framework for estimating \gamma_i in non-electrolyte mixtures assumes regular solution behavior, where entropic effects are ideal but enthalpic contributions dominate via differences in solubility parameters \delta_i. In regular solution theory, the activity coefficient relates to the interaction parameter as \ln \gamma_i \propto V_i (\delta_i - \delta_j)^2 / RT, with V_i the molar volume of component i, capturing dispersion forces in non-polar systems like hydrocarbons. This approach simplifies predictions but assumes random mixing and neglects specific association effects prevalent in polar non-electrolytes. Despite its utility, the treatment of activity in non-electrolyte solutions has limitations, as it overlooks variations in long-range interactions like dipole-dipole forces in highly polar mixtures or conformational entropy changes in associating liquids, leading to inaccuracies beyond simple binaries. Additionally, while ionic effects are absent, the model fails when trace charges or amphiphilic behavior introduces complexities not captured by pairwise solubility parameter differences.

Activity in Electrolyte Solutions

In electrolyte solutions, the concept of activity extends beyond neutral molecules to account for electrostatic interactions between ions, which cause significant deviations from ideal behavior even at moderate concentrations. While single-ion activities a_+ and a_- can be theoretically defined via the chemical potential \mu_i = \mu_i^\circ + RT \ln a_i for cations and anions, they are not directly measurable due to the principle of electroneutrality, which requires that solutions remain electrically neutral and prevents isolation of individual ion contributions without a counterion. Instead, practical thermodynamics employs the mean ionic activity a_\pm, defined for a salt like NaCl (\ce{NaCl -> Na+ + Cl-}) as a_\pm = (a_+ a_-)^{1/2}, which represents the geometric mean and ensures thermodynamic consistency. The corresponding mean ionic activity coefficient \gamma_\pm relates to mole fractions or molalities, such as a_\pm = \gamma_\pm (x_+ x_-)^{1/2} \nu, where \nu is the number of ions per formula unit (here \nu = 2) and x_+, x_- are the mole fractions of the ions; this formulation allows electrolyte solutions to be treated as fully dissociated for strong electrolytes./16%3A_The_Chemical_Activity_of_the_Components_of_a_Solution/16.17%3A_Activities_of_Electrolytes_-_The_Mean_Activity_Coefficient) To quantify the effects of ionic interactions, the I is introduced as a key parameter, defined as I = \frac{1}{2} \sum_i m_i z_i^2, where m_i is the of i and z_i its charge. This measure weights contributions by the square of the charge, capturing the stronger influence of multivalent ions on non-idealities. The mean ionic activity coefficient \gamma_\pm is then expressed as a function of ionic strength, typically \ln \gamma_\pm = f(I), enabling predictions of deviations that scale with ion density and charge. For instance, in dilute solutions, the Debye-Hückel limiting provides a foundational approximation: \ln \gamma_\pm = -A |z_+ z_-| \sqrt{I}, where A is a constant depending on , constant, and solvent properties (approximately 0.51 for at 25°C). This arises from treating the ionic atmosphere around each as a linearized Poisson-Boltzmann distribution, highlighting long-range Coulombic effects at low concentrations (I < 0.001 mol/kg)./25%3A_Solutions_II_-_Nonvolatile_Solutes/25.06%3A_The_Debye-Huckel_Theory) The framework for activity in electrolytes was pioneered by Gilbert N. Lewis and Merle Randall in their 1923 treatise, which systematically applied thermodynamic principles to ionic solutions, introducing ionic strength and mean activities to reconcile experimental data on colligative properties and solubilities with theoretical expectations. Their work emphasized the need for activity corrections in electrolyte thermodynamics, laying the groundwork for later developments like the Debye-Hückel theory published the same year. Despite these advances, the inherent challenge of single-ion activities persists, as any measurement involving ion-selective electrodes or potentials inherently includes liquid junction effects that violate local electroneutrality assumptions, rendering absolute single-ion values conventional rather than absolute. This limitation underscores the reliance on mean activities for rigorous thermodynamic calculations in electrolyte systems.

Determination of Activity Coefficients

Common Experimental Methods

Common experimental methods for determining activity coefficients in solutions include electromotive force (EMF) measurements, vapor pressure-based techniques such as osmometry and isopiestic equilibration, solubility assessments, and colligative property analyses like freezing point depression or boiling point elevation. These approaches are widely used for both non-electrolyte and electrolyte systems, providing data across finite concentration ranges with typical precisions of ±0.01 in ln γ_i and often conducted at 25°C. Electromotive force (EMF) measurements using ion-selective electrodes offer a direct way to evaluate ionic activity coefficients in electrolyte solutions. In this setup, a galvanic cell is constructed with the ion-selective electrode responsive to the target ion and a reference electrode, such as Ag/AgCl. The measured cell potential follows the Nernst equation: E = E^0 + \frac{RT}{nF} \ln(a_i) where E^0 is the standard cell potential, R is the gas constant, T is temperature, n is the number of electrons transferred, F is Faraday's constant, and a_i = \gamma_i m_i is the activity of the ion (with \gamma_i as the activity coefficient and m_i as molality). By varying the solution concentration and measuring E, \gamma_i is calculated after accounting for the standard potential, typically yielding mean ionic activity coefficients with random errors below ±0.5 mV, corresponding to precisions of about ±0.01 in ln \gamma_+. This method is particularly effective for uni-univalent electrolytes like NaCl over concentrations up to several molal. Vapor pressure osmometry and isopiestic methods determine activity coefficients indirectly through osmotic coefficients in both non-electrolyte and electrolyte solutions. Vapor pressure osmometry involves equilibrating a solution droplet with solvent vapor in a controlled chamber, where the temperature difference induced by vapor pressure lowering is proportional to the osmotic pressure; this yields the osmotic coefficient \phi from the relation between solution vapor pressure and pure solvent vapor pressure. The isopiestic method extends this by equilibrating multiple solutions in a vacuum chamber until equal vapor pressures are achieved, often using a reference electrolyte like NaCl for calibration, allowing \phi to be obtained from molality ratios. Activity coefficients are then derived from the : \ln \gamma_i = \int_0^{m} (\phi - 1) \, d \ln m for mean ionic activity coefficients in electrolytes, providing accurate data up to high concentrations (e.g., saturation) with uncertainties typically below ±0.005 in \phi. These techniques are standard for aqueous systems at 25°C and require precise temperature control for equilibration. Solubility measurements assess activity coefficients for sparingly soluble salts in mixed solvents by leveraging saturation equilibria. For a salt like AgCl in a solvent mixture, the solubility product K_{sp} = a_{Ag^+} a_{Cl^-} = \gamma_+^2 m_+^2 remains constant, so variations in saturation molality m_{sat} with solvent composition allow \gamma_+ to be calculated as \gamma_+ = \sqrt{K_{sp}} / m_{sat}, assuming ideal behavior in the solid phase. This indirect method is useful for non-aqueous or mixed-solvent systems where direct measurements are challenging, often applied to salts with solubilities below 0.01 M, and provides \gamma values with precisions around ±0.02 over temperature ranges including 25°C. Representative examples include determining \gamma for CaSO4 in ethanol-water mixtures from measured saturation points. Freezing point depression and boiling point elevation measurements exploit colligative properties to derive activity coefficients from deviations in non-electrolyte and electrolyte solutions. For freezing point depression, the solvent activity a_w is related to the depression \Delta T_f by \ln a_w = -\frac{\Delta H_f \Delta T_f}{R T_0^2}, where \Delta H_f is the solvent's heat of fusion and T_0 is the normal freezing point; solute activity coefficients follow from integration via Gibbs-Duhem, yielding \gamma_i for dilute to moderate concentrations. Boiling point elevation uses a similar approach with vapor pressure data. These methods are precise (±0.001 K in \Delta T) for aqueous solutions at temperatures near 0°C or 100°C, but corrections for heat capacity are needed; they typically achieve ±0.01 accuracy in ln \gamma_i for systems like sucrose or NaCl solutions.

Measurements at Infinite Dilution

The activity coefficient at infinite dilution, denoted \gamma_i^\infty, is defined as \gamma_i^\infty = \lim_{x_i \to 0} \gamma_i, where x_i is the mole fraction of solute i and \gamma_i is its activity coefficient. This limiting value isolates solute-solvent interactions, excluding any solute-solute contributions that become negligible at vanishing concentrations. It serves as a fundamental measure of non-ideality in dilute solutions, reflecting the extent to which the solute's chemical potential deviates from ideality due to solvent effects alone. A primary experimental method for determining \gamma_i^\infty is gas-liquid chromatography (GLC), which leverages the solute's retention behavior on a column coated with the solvent as the stationary phase. In GLC, the solute is injected at trace amounts to approximate infinite dilution conditions, and its net retention time t_R is measured. The activity coefficient is then calculated using the relation \ln \gamma_i^\infty = \ln \left[ \frac{RT}{V_s} \cdot \frac{t_R}{j \phi} \right] - \frac{B}{V_s} \cdot \frac{2 P_0}{RT}, where R is the gas constant, T is the absolute temperature, V_s is the molar volume of the solvent, j is the James-Martin correction factor for pressure drop across the column, \phi is the fugacity coefficient of the solute, B is the second virial coefficient of the solute, and P_0 is the outlet pressure. This equation, derived from chromatographic theory, accounts for gas-phase non-idealities and column hydrodynamics. Measurements are typically conducted over a temperature range to derive thermodynamic properties like excess enthalpies via the . For systems where direct infinite dilution data are challenging, \gamma_i^\infty can be obtained by extrapolating activity coefficients measured at finite concentrations using virial expansions of the excess Gibbs energy. A common form is \ln \gamma_i = \ln \gamma_i^\infty + B x_i + C x_i^2 + \cdots, where B and C are composition-dependent virial coefficients capturing pairwise and higher-order interactions, respectively. This approach is particularly useful for validating models or when GLC is impractical, such as for highly volatile or reactive solutes. Values of \gamma_i^\infty are essential for parameterizing thermodynamic models like UNIQUAC or NRTL, enabling predictions of phase equilibria and azeotrope formation. For instance, \gamma_i^\infty > 1 indicates positive deviations from ideality, often corresponding to mutual solubility in immiscible pairs, such as hydrocarbons in where \gamma_i^\infty can exceed 1000. Tabulated \gamma_i^\infty data for numerous systems are compiled in the DECHEMA Activity Coefficients at Infinite Dilution () database, facilitating applications in selection and design.

Advanced and Specialized Techniques

Radiochemical methods emerged in the as a means to precisely measure activity coefficients of trace s in mixed solutions, particularly where conventional techniques were limited by low concentrations. These approaches utilized radioactive tracers to track distribution with high sensitivity, enabling determinations in systems with levels as low as 10^{-6} M. For instance, early work at employed isotopes such as ^{24} (half-life 15 hours) and ^{22} (half-life 2.6 years) in synthetic ion-exchange resins to quantify distribution ratios between the resin phase and aqueous solutions (0.005–1.5 m), yielding activity coefficients for salts like NaCl and KCl with precisions of ±0.004–0.005. In solvent variants of these methods, radioactive tracers facilitate measurement of distribution coefficients () for ions partitioning between aqueous and phases, from which activity coefficients are derived. The relation is given by D = K \frac{\gamma_{\org}}{\gamma_{\aq}} where K is the extraction equilibrium constant (often determined independently), \gamma_{\org} is the activity coefficient in the phase (typically near for dilute conditions), and \gamma_{\aq} is the aqueous activity coefficient. This approach is especially valuable for trace actinides or metals in complex matrices. A representative example involves (IV) extraction from using thenoyltrifluoroacetone in , where a radioactive tracer (concentration ~10^{-6} M) was distributed and quantified via alpha counting; the resulting \gamma_{\Pu(\ClO_4)_4} values increased with acidity, from approximately 0.86 at 0.1 M HClO₄ to 13.8 at 4 M. Neutron scattering techniques provide structural insights into concentrated solutions, serving as indirect probes for local activities by revealing ion pairing, shells, and spatial correlations that underlie non-ideal behavior. Total neutron scattering, combining elastic and inelastic components, captures the pair distribution functions for ions and , allowing computation of Kirkwood-Buff integrals that link microscopic to thermodynamic properties like activity coefficients. In a 7.3 m CaCl_2 , for example, neutron data highlighted enhanced ion-water correlations and reduced self-diffusion coefficients (e.g., D_{Ca^{2+}} ≈ 0.2 × 10^{-9} m²/s versus 0.79 × 10^{-9} m²/s for dilute), indicating local activity enhancements due to restrictions that contribute to overall \gamma deviations exceeding 20% from ideality. Spectroscopic methods, including NMR and Raman, offer local probes of activity in concentrated solutions by examining solvation dynamics and speciation. Raman spectroscopy detects vibrational shifts in ion-solvent bonds, quantifying coordination numbers and ion pairing that influence effective local concentrations and thus activity. For saturated LiTFSI electrolytes (up to 5 m), Raman analysis of the TFSI anion symmetric stretch (~740 cm^{-1}) revealed a transition from solvent-separated ion pairs at low concentration to contact ion pairs at high concentration, correlating with water activity reductions through increased ion pairing, which impacts ionic conductivity and stability in battery applications. Similarly, ^{7}Li and ^{19}F NMR in these systems measures chemical shifts and relaxation times to assess local ion environments, providing hydration numbers (e.g., ~4 for Li^+ in 3 m solutions) that inform activity corrections via solvation models. Computational simulations, particularly molecular dynamics (MD), enable estimation of activity coefficients in challenging systems via free energy perturbation (FEP), where the excess chemical potential \mu^{\ex} of an ion is computed by gradually charging it in solution, yielding \gamma = \exp(\mu^{\ex}/RT). This thermodynamic integration approach captures ion-solvent and ion-ion interactions explicitly. In NaCl aqueous solutions (up to 5 m), implicit-water MD with FEP predicted mean \gamma_{\pm} values agreeing with experiments within 5% (e.g., 0.66 at 1 m versus 0.657 measured), highlighting the role of polarization in concentrated regimes; detailed theory of these connections to molecular properties is addressed elsewhere. Under extreme industrial conditions such as (>100 ) and (>200°C), activity coefficients are determined using specialized volumetric and acoustic techniques in pressure-resistant cells. Ultrasonic measurements with piezoelectric transducers assess (u) and attenuation, which, combined with data, yield and hydration numbers for model-based \gamma calculations. For strong electrolytes like NaCl, the concentration dependence of (n_h) is incorporated into \ln \gamma_{\pm} = -n_h \ln a_w + f(I) where a_w is water activity and f(I) accounts for ionic strength; at 300 bar and 150°C, such methods predict \gamma_{\NaCl} increases of ~10% over ambient due to reduced ion association. These probes are critical for geothermal brines or supercritical processes.

Theoretical Models and Calculations

Models for Dilute Solutions

In dilute electrolyte solutions, the Debye-Hückel theory provides a foundational theoretical framework for calculating activity coefficients by accounting for long-range electrostatic interactions between ions treated as point charges in a continuum solvent. The theory assumes complete dissociation of the electrolyte, spherical non-polarizable ions, and a linear response of the ionic atmosphere to the central ion's charge, leading to a Poisson-Boltzmann description of the electrostatic potential. This model is valid primarily for ionic strengths I < 0.1 M, where short-range interactions and ion pairing are negligible. The extended Debye-Hückel equation for the common logarithm of the mean ionic activity coefficient \gamma_\pm incorporates a finite ion size parameter a to improve accuracy beyond the limiting law, expressed as: \log_{10} \gamma_\pm = -\frac{A |z_+ z_-| \sqrt{I}}{1 + B a \sqrt{I}} + C I Here, A and B are parameters dependent on the solvent's dielectric constant \epsilon and temperature T, with A \propto (\epsilon T)^{-3/2} and B \propto (\epsilon T)^{-1/2}, while C is an empirical term for higher-order corrections; for aqueous solutions at 25°C, A \approx 0.509 (mol/kg)^{-1/2}. The term C I accounts for linear deviations at slightly higher dilutions. This form extends the applicability to moderate dilutions while retaining the electrostatic focus. An empirical refinement of the Debye-Hückel approach, the Davies equation, further extends predictions to ionic strengths up to approximately 0.3 M by incorporating a negative linear term to better fit experimental data for mean activity coefficients: \log_{10} \gamma_\pm = -A |z_+ z_-| \left( \frac{\sqrt{I}}{1 + \sqrt{I}} - 0.3 I \right) This equation omits explicit ion size dependence but uses the same A parameter, making it simpler for practical calculations in dilute regimes without needing species-specific fits. It performs well for 1:1 electrolytes like , where deviations from ideality are dominated by ionic atmosphere effects. For non-electrolyte solutions in the dilute limit, activity coefficients can be modeled using the low-concentration approximations of regular solution theories such as the one-parameter or equations, which capture asymmetric deviations from ideality through excess Gibbs energy terms. In a binary mixture where solute 1 is dilute (x_1 \ll 1), the model simplifies to \ln \gamma_1 = A x_2^2, with A reflecting solute-solvent interactions; similarly, the model simplifies to \ln \gamma_1 \approx A_{12}^\infty at infinite dilution, a constant reflecting solute-solvent interactions. These models assume random mixing with local composition effects negligible in dilution. Parameters in these models, including a, B, C, and A in the Margules limit, are typically estimated by least-squares fitting to experimental data such as osmotic coefficients, vapor pressures, or solubility measurements at varying dilutions, ensuring the model reproduces observed non-idealities while adhering to thermodynamic consistency. For Debye-Hückel variants, fits often prioritize low-I data to isolate electrostatic contributions, with ion-size a values around 3–5 Å derived iteratively./25%3A_Solutions_II_-_Nonvolatile_Solutes/25.07%3A_Extending_Debye-Huckel_Theory_to_Higher_Concentrations)

Models for Concentrated Solutions

For concentrated electrolyte solutions, where ionic strengths exceed 1 M and approach saturation, models must account for short-range ion interactions, ion pairing, and higher-order effects that dominate over long-range electrostatics described by limiting theories like . These models incorporate empirical parameters fitted to experimental data to predict accurately across a wide range of compositions, including mixed salts and brines. The Pitzer model, developed for strong electrolytes, expresses the natural logarithm of the mean activity coefficient as \ln \gamma_\pm = f^\gamma + \sum_j \beta_{MX}(I) m_j + \sum_{j,k} C_{MX,jk} m_j m_k, where f^\gamma is a Debye-Hückel-like term modified for higher concentrations, \beta_{MX}(I) are ionic strength-dependent second virial coefficients capturing binary ion interactions, and C_{MX,jk} are third virial coefficients for ternary interactions, all serving as adjustable parameters derived from osmotic and activity coefficient measurements. This formulation enables predictions for single and mixed electrolyte systems up to high molalities, with parameters tabulated for common ions like Na^+, Cl^-, and SO_4^{2-}. For example, in NaCl solutions at 6 m, the model yields \gamma_\pm \approx 0.65, closely matching experimental vapor pressure data. The model has been widely adopted for geochemical applications due to its robustness in multisalt environments. The Stokes-Robinson model addresses hydration effects in concentrated solutions by relating the activity coefficient to that in pure water and the solvent's activity, given by \ln \gamma_\pm = \ln \gamma_\pm^{\text{water}} + \frac{\nu_h}{\nu} \ln a_w, where \nu_h is the effective hydration number (e.g., 4 for NaCl), \nu is the stoichiometric coefficient, and a_w is the water activity computed from osmotic coefficients. This approach assumes ions bind water molecules, reducing the "free" solvent available, and performs well for salts like LiCl and CaCl_2 at concentrations above 5 m, where it predicts salting-out behavior in mixed systems. Hydration numbers are estimated from solubility or partial molar volume data, making the model semi-empirical and suitable for systems where water structuring is prominent. For non-electrolyte solutions at high concentrations, the UNIQUAC model separates contributions into combinatorial (entropic, size- and shape-based) and residual (enthalpic, interaction-based) parts: \ln \gamma_i = \ln \gamma_i^C + \ln \gamma_i^R, with the combinatorial term \ln \gamma_i^C = \ln \frac{\phi_i}{x_i} + 1 - \frac{\phi_i}{x_i} - l_i \ln \frac{\phi_i}{\theta_i} - 5 q_i \ln \frac{1 - \phi_i + \phi_i / r_i}{1 - \phi_i} depending on relative volume parameters r_i and surface area parameters q_i, while the residual term \ln \gamma_i^R = q_i \left(1 - \ln \frac{\sum_j \tau_{ji} \phi_j}{\sum_j \phi_j} - \sum_j \frac{\phi_j \tau_{ij}}{\sum_k \phi_k \tau_{kj}}\right) incorporates binary interaction parameters \tau_{ij} = \exp\left(-\frac{u_{ij} - u_{ji}}{RT}\right). These parameters are derived from group contributions or regression to vapor-liquid equilibrium data, enabling predictions for polymer-solvent or organic mixtures like ethanol-water, where UNIQUAC predicts activity coefficients accurately across compositions, with \gamma_\text{ethanol} reaching ~2.5 at intermediate mole fractions near the azeotrope (x ≈ 0.4). UNIQUAC excels in handling molecular asymmetry absent in electrolyte-focused models. Ion trio models extend virial expansions to explicitly include three-body interactions for specific ion associations in concentrated brines, such as Na^+-Ca^{2+}-Cl^- triplets, through higher-order terms that correct for clustering beyond pairwise approximations. These are particularly vital in multisalt systems like seawater evaporites, where they improve solubility predictions by 10-20% over binary-only models at ionic strengths above 4 M. Seminal implementations, like those in the , fit ternary coefficients to mineral saturation data for accurate speciation in Na-K-Mg-Ca-Cl-SO_4-H_2O systems up to 25°C. Such models are validated for ionic strengths I > 1 M, extending reliably to in many aqueous systems, though parameter availability limits applicability to well-studied ions; deviations occur in asymmetric or high-temperature cases requiring extensions.

Connections to Molecular and Ionic Properties

In the Debye-Hückel theory, the parameter a represents the effective ionic , appearing in the denominator of the limiting expression for activity coefficients to account for the finite size of ions and prevent unphysical close approaches. This parameter encapsulates the hydrated ion size, typically ranging from 3 to 5 for common monovalent ions such as Li⁺, Na⁺, Cl⁻, and Br⁻, based on empirical fits to experimental data. The Stokes-Robinson hydration model connects activity coefficients to properties by incorporating the number of water molecules bound in primary shells around , which reduces the effective free water available for osmotic behavior. numbers, such as approximately 4-6 for Na⁺ and 1-2 for Cl⁻, influence the activity coefficient through corrections that reflect stronger ion-water binding, thereby altering the non-ideal behavior in solutions. In the mean spherical approximation (), a statistical mechanical approach, activity coefficients for are derived from pair correlation functions g(r), which describe the of relative to a central and capture local composition effects beyond mean-field descriptions. The excess , related to \ln \gamma, integrates contributions from these g(r) functions, highlighting how short-range ion-ion and ion-solvent interactions govern deviations from ideality in primitive models. For non-electrolyte solutions, the \delta links infinite-dilution activity coefficients \gamma^\infty to molecular cohesion energies, as expressed in regular solution where \ln \gamma^\infty \approx \frac{V_2}{[RT](/page/RT)} (\delta_1 - \delta_2)^2, with V_2 the solute . This underscores how differences in cohesive densities between solute and drive non-ideal mixing, with \delta values (e.g., 23.4 ^{1/2} for ) providing a measure of intermolecular forces like and . Modern parameterizations of activity coefficient models increasingly incorporate calculations, particularly (DFT), to compute accurate ion-water interaction energies and structures. These DFT-derived potentials, such as binding energies for first solvation shells (e.g., -100 to -400 kJ/mol for halides), inform parameters in models like extended Debye-Hückel or , improving predictions for concentrated solutions by bridging microscopic interactions with macroscopic .

Factors Affecting Activity Coefficients

Dependence on Temperature and Pressure

The temperature dependence of activity coefficients arises from their connection to the excess Gibbs energy and can be expressed through the partial derivative relation derived from thermodynamic state functions: \left( \frac{\partial \ln \gamma_i}{\partial T} \right)_P = -\frac{\bar{H}_i^E}{RT^2}, where \bar{H}_i^E is the partial molar excess enthalpy of component i, R is the gas constant, and T is the absolute temperature. This relation indicates that the rate of change of \ln \gamma_i with temperature at constant pressure reflects the enthalpic contributions to non-ideality; for many systems, positive \bar{H}_i^E values lead to decreasing \gamma_i as temperature increases. For pressure dependence, the analogous thermodynamic relation is \left( \frac{\partial \ln \gamma_i}{\partial P} \right)_T = \frac{\bar{V}_i^E}{RT}, where \bar{V}_i^E is the partial molar excess volume of component i. Excess volumes \bar{V}_i^E are typically small in liquid mixtures at moderate pressures due to near-incompressibility, resulting in weak pressure effects on \gamma_i; however, they become significant at high pressures (e.g., >100 MPa), where volume changes can alter non-ideal interactions. In some systems, particularly at infinite dilution or for specific solutions, empirical correlations approximate the dependence using Arrhenius-like forms, such as \ln \gamma(T) = A + B/T, where A and B are fitted parameters capturing effects. These forms are useful for modeling over limited ranges and align with the exponential sensitivity implied by the excess relation. For aqueous solutions, activity coefficients often decrease with increasing at fixed composition, attributed to reduced structuring around ions. A representative example is NaCl(aq), where \gamma for the mean ionic activity coefficient drops from approximately 0.657 at 298 K to 0.626 at 373 K for 1 mol·kg⁻¹ solutions, reflecting weakened ion-water interactions. Typical values for the are on the order of d \ln \gamma / dT \approx -0.001 K⁻¹ for NaCl over 273–373 K, derived from parametric fits to experimental osmotic and activity data.

Dependence on Composition and Solvent Effects

The activity coefficient of a solute in a varies significantly with the overall , reflecting deviations from ideal mixing due to intermolecular interactions. In binary mixtures, this dependence is often modeled using expansions of the excess , such as the Margules equation, which expresses the natural logarithm of the activity coefficient for component i as \ln \gamma_i = \sum_j A_{ij} x_j, where A_{ij} are composition-independent parameters and x_j is the of component j. This one-parameter form for symmetric systems captures symmetric deviations in activity coefficients, while higher-order terms (e.g., two-parameter Margules) account for asymmetry, providing a simple yet effective representation for regular solution behavior in non-electrolyte binaries. For multicomponent systems, mixing rules extend binary models to predict activity coefficients across compositions, influencing phase behavior topologies. The Van Konynenburg and Scott delineates six types of binary fluid phase diagrams based on critical lines and coexistence regions, where the choice of activity coefficient model (e.g., via parameters in excess Gibbs energy functions) determines the emergence of phenomena like liquid-liquid immiscibility or azeotropy. These rules ensure thermodynamic consistency, such as satisfying the Gibbs-Duhem equation, and are crucial for systems exhibiting type III or V behavior, where composition-dependent activity coefficients drive discontinuous critical curves. Solvent choice profoundly impacts activity coefficients by altering solute-solvent affinities, with poorer solvents yielding higher values due to reduced . For instance, the infinite-dilution activity coefficient of (\gamma^\infty) is approximately 4.5 in —a polar where hydrogen bonding stabilizes the solute—but exceeds 10 in , a nonpolar solvent where ethanol's polarity leads to and limited . In systems, such as - mixtures, adding a like to can lower \gamma^\infty for hydrophobic solutes by improving , though excessive may reverse this effect through preferential interactions. In electrolyte solutions, salts can induce salting-out (increased activity coefficients) or salting-in (decreased values) for nonelectrolytes via ion-solvent competition. The Setschenow equation quantifies this as \log(\gamma / \gamma^0) = k_s C_s, where \gamma^0 is the activity coefficient without salt, k_s is the salting constant (positive for salting-out, e.g., 0.18 L/mol for NaCl with benzene), and C_s is salt concentration; this empirical relation arises from reduced water availability around the solute. For electrolytes themselves, composition effects are pronounced: the mean ionic activity coefficient \gamma_\pm for HCl in aqueous solution starts near 0.81 at 0.1 mol/kg, dips to a minimum around 0.75 at 0.5 mol/kg, and rises sharply to over 3 at 5 mol/kg due to ion pairing and decreased dielectric screening at higher concentrations.

Applications in Thermodynamics

Role in Chemical Equilibrium Constants

In chemical equilibrium, the thermodynamic equilibrium constant K is defined in terms of activities a_i = \gamma_i m_i / m^\circ (where \gamma_i is the activity coefficient, m_i the molality, and m^\circ = 1 mol kg^{-1} the standard molality), ensuring K is dimensionless and independent of concentration units or non-ideality effects. In contrast, the concentration-based equilibrium constant K_c (or K_m in molality terms) relates to K via K = K_c \prod_i \gamma_i^{\nu_i}, where \nu_i are the stoichiometric coefficients (positive for products, negative for reactants); this correction accounts for deviations from ideality due to intermolecular interactions. For electrolyte solutions, mean ionic activity coefficients \gamma_\pm are often used to simplify the product for charged species. A key application arises in acid-base equilibria, such as the dissociation of a weak HA \rightleftharpoons H^+ + A^-, where the thermodynamic K_a relates to the concentration-based K_a^c by K_a = K_a^c \gamma_\pm^2 / \gamma_\text{HA}, with \gamma_\text{HA} \approx 1 for species. This leads to the observed pK_a shifting from the infinite-dilution value pK_a^0 according to pK_a = pK_a^0 - log \gamma_\pm, where \gamma_\pm < 1 at finite ionic strengths reduces the apparent acidity. Such corrections are vital for accurate calculations in ionic media, as uncorrected K_c overestimates extents. The temperature dependence of equilibrium constants follows the van't Hoff equation \frac{d \ln K}{dT} = \frac{\Delta H^\circ}{RT^2}, applied to the thermodynamic K and thus inherently incorporating activity coefficients evaluated at each temperature. However, for concentration-based constants, the full temperature variation includes an additional term \frac{d \ln K_c}{dT} = \frac{\Delta H^\circ}{[RT^2](/page/RT-2)} + \sum_i \nu_i \frac{d \ln \gamma_i}{dT}, reflecting how \gamma_i typically decrease with rising temperature due to enhanced thermal motion weakening ion interactions. This coupling ensures predictions of equilibrium shifts with temperature account for both enthalpic driving forces and non-ideal solution behavior. In like ammonia synthesis (N_2 + 3H_2 \rightleftharpoons 2NH_3), high pressures (100-300 bar) induce significant non-ideality, requiring coefficients (analogous to activity coefficients for gases) to compute the true K from partial pressures; deviations can shift predicted yields by up to 20% at operating conditions. Similarly, for solubility equilibria, the solubility product K_{sp} for sparingly soluble salts like AgCl(s) \rightleftharpoons Ag^+ + Cl^- is corrected as K_{sp} = K_{sp}^c \gamma_\pm^2, where elevated suppresses \gamma_\pm and reduces apparent solubility via the amplified by non-ideality. Activity coefficients are essential for precise chemical in natural waters, such as where ionic strengths reach 0.7 mol kg^{-1}, enabling accurate determination and system modeling; neglecting them leads to errors exceeding 0.2 units in CO_2 equilibration calculations critical for studies.

Use in Equilibria and

Activity coefficients play a crucial role in describing vapor-liquid (VLE) by accounting for non-ideal behavior in mixtures, enabling accurate predictions of phase compositions and separations in processes. The modified expresses the of component i in the vapor phase as y_i P = \gamma_i x_i P_i^\circ, where y_i is the vapor , P is the total , x_i is the liquid , \gamma_i is the activity coefficient, and P_i^\circ is the saturation of pure i. This relation allows for the calculation of activity coefficients from experimental VLE data or vice versa, facilitating the design of separation columns. In azeotrope formation, where liquid and vapor compositions are identical (x_i = y_i), the activity coefficients at the azeotropic point satisfy \gamma_i = P / P_i^\circ, providing a direct link between non-ideality and the composition at which minima or maxima occur. For the -water system, which forms a minimum-boiling at approximately 95.6 wt% at , activity coefficients deviate significantly from unity—ethanol's \gamma exceeds 1 in water-rich mixtures due to positive deviations from ideality—enabling prediction of the azeotropic composition through models like or NRTL fitted to VLE data. For liquid-liquid equilibrium (LLE), activity coefficients ensure equality of chemical potentials across immiscible phases, expressed as \gamma_i^\alpha x_i^\alpha = \gamma_i^\beta x_i^\beta for each component i in phases \alpha and \beta. This condition underpins the calculation of distribution coefficients for solutes partitioning between phases, such as in processes, where the distribution ratio m = \frac{x_i^\beta}{x_i^\alpha} \approx \frac{\gamma_i^\alpha}{\gamma_i^\beta} \cdot \frac{S_i^\beta}{S_i^\alpha}, with S_i denoting in the pure phase. Activity coefficients thus quantify phase selectivity, with values greater than unity in the aqueous phase often enhancing organic phase partitioning for hydrophobic solutes. Activity coefficients also govern solubility enhancement in mixed solvents, particularly through cosolvency, where addition of a cosolvent like to increases the of poorly -soluble drugs beyond linear additivity. For hydrophobic pharmaceuticals, such as ibuprofen, cosolvency arises from reduced activity coefficients in the mixed solvent due to favorable solute-cosolvent interactions, leading to increases of up to several orders of magnitude; for instance, rises from 0.13 mg/mL in to approximately 10 mg/mL in 50 vol% - mixtures at 25°C. This effect is modeled using log-linear relationships incorporating composition-dependent \gamma, aiding formulation design for oral delivery. The (UNIversal Functional Activity Coefficient) group contribution method provides a predictive framework for estimating activity coefficients in VLE and LLE without experimental data, by decomposing molecules into functional groups and summing combinatorial and residual contributions to excess Gibbs energy. Developed from regression on binary VLE data, accurately predicts azeotropic compositions and LLE binodals for systems like alcohol-water or hydrocarbon-alcohol mixtures, with average deviations in VLE of 5-10% for group interaction parameters derived from over 1000 binaries. In LLE applications, it estimates distribution coefficients for solvents, supporting in . An illustrative example is the solubility of NaCl in ethanol-water mixtures, where activity coefficients decrease with increasing ethanol content due to salting-out effects, reducing NaCl solubility from 6.15 mol/kg in pure to below 1 mol/kg in 80 wt% at 25°C. Models incorporating for ionic \gamma predict this behavior, with mean activity coefficients around 0.66 in and approximately 0.72 in 20 wt% mixtures at 25°C, decreasing further to about 0.41 at higher contents.

Applications in Electrochemical Systems

In electrochemical systems, activity coefficients play a crucial role in accurately predicting cell potentials through the Nernst equation, which relates the electrode potential to the activities of species involved in the half-reaction. The standard form of the Nernst equation is given by E = E^\circ - \frac{RT}{nF} \ln \left( \prod_i a_i^{\nu_i} \right), where E is the cell potential, E^\circ is the standard potential, R is the gas constant, T is the temperature, n is the number of electrons transferred, F is Faraday's constant, a_i are the activities of species i with stoichiometric coefficients \nu_i, and the product is over reactants and products. Since activities are defined as a_i = \gamma_i c_i (with \gamma_i as the activity coefficient and c_i as the concentration), the equation can be rewritten as E = E^\circ - \frac{RT}{nF} \ln \left( K_c \prod_i \gamma_i^{\nu_i} \right), where K_c is the equilibrium constant in terms of concentrations; deviations from ideality (\gamma_i \neq 1) thus introduce corrections to the potential that are essential for non-dilute solutions. In lithium-ion batteries, mean activity coefficients (\gamma_\pm) are particularly important for optimizing electrolyte performance, as they influence the thermodynamic stability and ion solvation in concentrated salt solutions like LiPF_6 in carbonate solvents. Accurate determination of \gamma_\pm via electromotive force measurements or vapor pressure osmometry allows for better prediction of salt solubility limits and minimization of concentration gradients that degrade battery efficiency during charge-discharge cycles. For instance, in high-energy-density cells, incorporating \gamma_\pm values ranging from 0.5 to 0.8 helps refine models for electrolyte formulation, reducing overpotential and extending cycle life. Activity coefficients also adjust models of ion diffusion and conductivity in electrochemical transport processes, where the Nernst-Einstein relation links ionic conductivity \sigma to self-diffusion coefficients D_i via \sigma = \frac{z_i^2 F^2 c_i D_i}{RT} under ideal conditions (z_i is the charge). In non-ideal solutions, this relation is modified by a thermodynamic factor incorporating \gamma_i, specifically \frac{\partial \ln a_i}{\partial \ln c_i} = 1 + \frac{\partial \ln \gamma_i}{\partial \ln c_i}, to account for the chemical potential gradient driving flux; neglecting this leads to overestimation of transport rates by up to 20-30% in concentrated electrolytes. This correction is vital for simulating migration in porous electrodes and membranes. Practical examples highlight the impact of activity coefficient errors in electrochemical devices. In pH electrodes, where the potential follows E = E^\circ - \frac{[RT](/page/RT)}{F} \ln a_{\ce{H+}} with a_{\ce{H+}} = \gamma_{\ce{H+}} c_{\ce{H+}}, ignoring \gamma_\pm in high-ionic-strength media like can introduce errors of 10-20 mV, equivalent to 0.2-0.3 units and compromising measurement accuracy in . Similarly, in electrochemical desalination processes such as , non-ideal activity coefficients in multi-ion electrolytes (e.g., NaCl-dominated with \gamma_\pm \approx 0.65-0.75) affect drops and , with uncorrected models overpredicting removal rates by 15-25%. In modern fuel cells, activity coefficients exhibit strong dependence on and , influencing proton and under varying operating conditions. At low relative (below 50%), reduced lowers \gamma for ions, increasing ohmic losses and degrading performance by up to 100 mV; elevated (60-80°C) further modulate \gamma through changes in ion pairing and , necessitating control for optimal . These effects are modeled to enhance in automotive applications.

References

  1. [1]
    Activity coefficient
    ### Definition and Summary of Activity Coefficient from Thermopedia
  2. [2]
    [PDF] Chapter 11 - Activity Models - MSU College of Engineering
    We refer to this "correction factor" as the activity coefficient. To derive the thermodynamic meaning of the activity coefficient, note: ∆G. nRT. G. nRT. G. nRT.
  3. [3]
    [PDF] Activity Activity Coefficients
    where K is the equilibrium constant, and aX is the activity of X and described by the activity coefficient γX and [X]:. aX = γX[X].
  4. [4]
    [PDF] Lecture 11: 10.15.05 Models of the chemical potential
    Oct 17, 2005 · When the activity coefficient = 1, the activity is equal to the mole fraction and the chemical potential reduces to: N; = N; + RTI. IDEAL ...
  5. [5]
    [PDF] 663 Chapter 19: Real Solutions Relate the observations that “oil and ...
    ... equation and n = nA + nB shows the relationship between the activity coefficients and the excess Gibbs energy: G. E. = nA (RT ln γA) + nB (RT ln γB). 19.2.5. G.
  6. [6]
    Hints & Examples - AIOMFAC home - McGill University
    On molality-basis, the so-called mean molal activity coefficient is then defined as the geometric mean, γ±(m) = [γ+ν+ × γ-ν-]1/(ν+ + ν-); here the superscript ( ...
  7. [7]
    Activity Coefficient - an overview | ScienceDirect Topics
    ... Raoult's law applies (Sposito and Mattigod, 1979). Thus, for exchanger phases, an activity coefficient is the ratio of an actual activity to a mole fraction.
  8. [8]
    [PDF] vapour pressure and activity in water–ethanol mixtures - HAL
    Apr 6, 2018 · The activities in the liquid phase are calculated from the mean vapour phase densities at a given composition, providing a good qualitative ...
  9. [9]
    SOLUBILITY. XII. REGULAR SOLUTIONS 1 - ACS Publications
    Large decrease in the melting point of benzoquinones via high-n eutectic mixing predicted by a regular solution model.
  10. [10]
    A Flory–Huggins model based on the Hansen solubility parameters
    In this work, we present a methodology, based on which the Hansen solubility parameters can be incorporated in the Flory–Huggins model.
  11. [11]
    [PDF] REGULAR SOLUTION THEORY AND ITS APPLICATIONS
    Jul 15, 2016 · Provided the ratio of activity coefficients remains constants irrespective of the total concentration and total composition of the two phases, ...
  12. [12]
    Thermodynamics - Gilbert Newton Lewis, Merle Randall
    Electrolyte Solutions. 298. Theories of Electrolyte Solutions. 332. Galvanic Cells and Electrode Potentials. 349. Thermal Properties of Solutions. 373 ...
  13. [13]
    Thermodynamics : Lewis, Gilbert Newton, 1875-1946 - Internet Archive
    Oct 15, 2010 · First published in 1923 under title: Thermodynamics and the free energy of chemical substances. ... Randall, Merle, 1888-1950, joint author.
  14. [14]
    The Impossibility of Measuring Individual Ion Activity Coefficients ...
    The present analysisconsiders the results obtained when applying the procedure to typical friendlycases liable to precise thermodynamic–mathematical treatments.Missing: single- electroneutrality
  15. [15]
    Isopiestic Determination of the Osmotic and Activity Coefficients of K ...
    The solid and dashed lines show calculated differences of Goldberg's 23 recommended values for the osmotic coefficient and the solute activity coefficient from ...
  16. [16]
    THE FREEZING POINT AND ACTIVITY COEFFICIENT OF ...
    Calculations of Freezing Point Depression, Boiling Point Elevation, Vapor Pressure and Enthalpies of Vaporization of Electrolyte Solutions by a Modified ...
  17. [17]
    Vapor Pressure Osmometry Determination of the Osmotic and ...
    From the experimental osmotic coefficient data, the water activity, water activity coefficient and vapor pressure values were evaluated.
  18. [18]
    The activity coefficients of non-electrolytes in seawater - ScienceDirect
    The activity coefficient (γi) is equal to the ratio of the solubility in water divided by the value in the solution (S0/S) for sparingly soluble non- ...
  19. [19]
    New Model to Predict Infinite Dilution Activity Coefficients Based on ...
    Mar 21, 2023 · Infinite dilution activity coefficient (γ∞) is a physical quantity reflecting solute–solvent interaction. It plays a role in the ...
  20. [20]
    Measurement of Activity Coefficients at Infinite Dilution Using Gas ...
    The GLC technique was chosen as the most preferable measuring technique for the planned investigations because GLC shows several advantages in comparison to ...
  21. [21]
    Gas–liquid chromatography measurements of activity coefficients at ...
    The average γ 13 ∞ values for the different amounts of solvent on the column packing within the temperature range from T = (298.15 K to 368.15) K are tabulated ...
  22. [22]
    Comparison of the Osmotic Virial Equation with the Margules Activity ...
    Jan 23, 2019 · The main goal of this paper is to compare the general polynomial forms of the osmotic virial equation and the Margules model for the liquid activity ...
  23. [23]
    trends in solvent impact on infinite dilution activity coefficients of ...
    Oct 1, 2021 · The infinite dilution activity coefficient, γ i ∞ , is a frequently used molecular descriptor to pre-select a solvent for various kinds of fluid separations.<|control11|><|separator|>
  24. [24]
    ACT – Activity Coefficients at Infinite Dilution - DDBST
    Activity coefficients at infinite dilution (ACT) are available for a very large number of binary mixtures and are of great importance if high purities need to ...
  25. [25]
    radiochemical measurements of activity coefficients in mixed ...
    Two principal methods have been used in the past to investigate activity coefficients of individual electrolytes in mixed electrolytic solutions. The first ...
  26. [26]
    [PDF] Lawrence Berkeley National Laboratory - eScholarship
    The activity coefficient of Pu(Cl04)4 was measured in perchloric acid solutions. The procedure used was to measure the extraction of the neutral plutonium(IV)- ...
  27. [27]
    The “good,” the “bad,” and the “hidden” in neutron scattering and ...
    May 18, 2022 · The neutron total scattering technique is a unique scientific tool that allows one to unravel the structure of ionic solutions at an ...
  28. [28]
    Understanding Solvation Behavior of the Saturated Electrolytes with ...
    Nov 11, 2021 · In this work, the solvation structures of LiTFSI-based saturated electrolytes were investigated by SAXS/WAXS and Raman spectroscopy. Seven ...
  29. [29]
    Activity coefficients of aqueous electrolytes from implicit-water ...
    Nov 9, 2021 · We obtain activity coefficients in NaCl and KCl solutions from implicit-water molecular dynamics simulations, at 298.15 K and 1 bar, ...
  30. [30]
    Theory for the determination of activity coefficients of strong ...
    Theory for the determination of activity coefficients of strong electrolytes with regard to concentration dependence of hydration numbers ... speed of sound ...Missing: ultrasonic | Show results with:ultrasonic
  31. [31]
    [PDF] The Debye-Huckel Theory of Ion-Ion Interactions - Dalal Institute
    Hence, the Debye-Huckel's model gives the theoretical value of 𝛥𝜇i−I arising from ion-ion interactions. Now before we check the validity of the Debye-Huckel ...
  32. [32]
    [PDF] Predicting activity coefficients with the Debye‐Hückel theory using ...
    Local composition model for excess Gibbs energy. 430 of electrolyte systems. Part I: Single solvent, single completely dissociated electrolyte. 431 systems ...Missing: seminal | Show results with:seminal<|control11|><|separator|>
  33. [33]
    An analysis of the parameters in the Debye-Hückel theory
    This work investigates the effects of the two most important parameters of the Debye-Hückel theory – the size parameters of ions and the relative static ...
  34. [34]
    [PDF] GUIDELINES FOR THE EXTRAPOLATION TO ZERO IONIC ...
    This guideline describes the scientific basis for extrapolating equilibrium data to zero ionic strength using the SIT approach, and other methods.
  35. [35]
    [PDF] Extending the Van Laar Model to Multicomponent Systems
    Apr 15, 2010 · Of all these models, the van Laar model is the only one that was originated from a combina- tion of an empirical equation of state with the ...<|separator|>
  36. [36]
    Thermodynamics of electrolytes. II. Activity and osmotic coefficients ...
    Article September 1, 1973. Thermodynamics of electrolytes. II. Activity and osmotic coefficients for strong electrolytes with one or both ions univalent.
  37. [37]
    Application of the Pitzer model for the estimation of activity ...
    Abstract. The Pitzer model can be applied to predict the mean activity coefficients of salts in aqueous electrolyte solutions as a function of the molality.Missing: seminal | Show results with:seminal
  38. [38]
    Ionic Hydration and Activity in Electrolyte Solutions - ACS Publications
    Properties of aqueous solutions as demonstrated by calcium chloride. The effect of temperature on the extent of dissociation and on the number of bound waters.Missing: original | Show results with:original
  39. [39]
    Introduction to Activity Coefficient Models - Thermo - Read the Docs
    The activity coefficients that are used for phase equilibria are derived from the partial mole number derivative of excess Gibbs energy.
  40. [40]
    Traceable Values for Activity and Osmotic Coefficients in Aqueous ...
    Oct 29, 2020 · We present traceable three-parameter extended Hückel equations for the activity coefficient of sodium chloride and for the osmotic coefficient of water in ...Introduction · Theory · Results and Discussion · References
  41. [41]
    Pitzer Equation - an overview | ScienceDirect Topics
    The equations that give the pressure coefficient for In γ± and ϕ are. (18.49) ( ∂ ln γ ± ∂ p ) T = V ¯ 2 − V 2 ∘ ¯ υ R T. (18.50) ( ∂ ϕ ∂ p ) T = − ( V ...
  42. [42]
    Quickly Calculating the Activity Coefficient of a NaCl Solution Based ...
    Nov 5, 2024 · The model can be used to calculate the activity coefficient of solutions under high-temperature and high-pressure conditions.
  43. [43]
    [PDF] TEMPERATURE EFFECTS ON THE ACTIVITY COEFFICIENT OF ...
    Therefore, the Ion. Interaction Model is a simple and accurate technique to calculate the equilibrium properties of mixed electrolyte solutions. The above ...
  44. [44]
    Characteristics of activity coefficient models for liquid solutions
    Apr 14, 2004 · For the Margules model, Figure 1 (a) shows a somewhat atypical graph of ln γ1 and of ln γ2 vs. x1 corresponding to r = 3. In Figure 1(a), ln γ2 ...Missing: expansion original
  45. [45]
    Critical lines and phase equilibria in binary van der Waals mixtures
    This paper presents theoretical studies of phase equilibria in binary mixtures obeying the van der Waals equation, especially liquid-liquid equilibria that can ...
  46. [46]
    General Principles and Strategies for Salting-Out Informed by the ...
    Aug 15, 2017 · The degree of salting-out, reflecting the activity coefficient of ... A Setschenow equation for two-phase systems can also be used to express ...Introduction · Results and Discussion · Conclusion · Experimental Section
  47. [47]
    [PDF] Osmotic Coefficients and Mean Activity Coefficients of Uni-univalent
    Once this standardization is achieved the isopiestic method is used effectively to measure osmotic and activity coefficients to high concentrations, in some.
  48. [48]
    Development of Methods for the Determination of pKa Values - PMC
    Aug 8, 2013 · Activity coefficients (γ) of different ionic species in solution strongly depend on ionic strength (Debye-Hückel theory).
  49. [49]
    [PDF] How pKa values depend on temperature, pressure and ionic strength
    γ[H+]γ[A−] γ[HA]. (Eq. 14). Page 9. 8 where a is the activity, and γ is the activity coefficient. By using the activity term, a more precise equilibrium ...<|separator|>
  50. [50]
    Temperature Correction of log K - aqion
    Nov 15, 2023 · This formula is known as the van't Hoff equation, where the T-dependence of K is determined by the enthalpy change ΔH 0 alone.
  51. [51]
    Temperature Dependences of Limiting Activity Coefficients, Henry's ...
    Gas/liquid partition coefficients calculated at experimental temperatures using the van't Hoff equation reported in the cited source. e. Limiting activity ...
  52. [52]
    Thermodynamic Properties in Ammonia Synthesis
    The thermodynamic equilibrium constant K for the ammonia synthesis reaction is. K = aNH3 a?/22 ai./22 where aj is the activity of component i, which is given as.
  53. [53]
    [PDF] Determining Ksp of Lead(II) Iodide - Fountainhead Press
    The activity coefficient acts as a correction factor to account for decreases in ionic activity at high ionic strengths. While a detailed study of activity ...
  54. [54]
    Chemical speciation models based upon the Pitzer activity ...
    Jul 20, 2022 · Accurate chemical speciation models of solutions containing the ions of seawater have applications in the calculation of carbonate system equilibria.
  55. [55]
    [PDF] Vapor Liquid Equilibrium (VLE): 10.213 04/29/02 A Guide Spring 2002
    We use activity coefficient to relate the fugacity of the pure species ... yi P = γixiPisat (modified Raoult's Law). 3) non-ideal gas and liq. solution ...
  56. [56]
    [PDF] Non-Ideality Through Fugacity and Activity - University of Delaware
    Equation 42 is sometimes referred to as a modified Raoult's Law. The non-ideality of the liquid phase is totally contained in the activity coefficient ...
  57. [57]
    [PDF] VAPOR LIQUID EQUILIBRIA: A REVIEW | TSI Journals
    Jun 3, 2012 · The activity coefficient γi is often called as. “Raoult's law correction factor” and is highly concentration dependent. Eq. (4) and (5) apply ...
  58. [58]
    [PDF] Determination of Vapor-Liquid Equilibrium Parameters Using ...
    Sep 1, 2011 · At the azeotropic point, since the vapor and liquid compositions are the same, the activity coefficients are determined by the total pressure ...
  59. [59]
    LLE – Liquid-Liquid Equilibria - DDBST
    In case of low mutual solubility, activity coefficients at infinite dilution can directly be calculated from the solubility data. Several different ...
  60. [60]
    [PDF] GROUP-CONTRIBUTION METHODS IN ESTIMATING LIQUID ...
    The distribution coefficient, K, is the ratio of the composition of the solute in each liquid phase in equilibrium: Y.
  61. [61]
    Solubility enhancement of hydrophobic compounds by cosolvents
    The purpose of this study is to investigate the relationship between the hydrophobicity of the drug and its solubilization behavior in water–cosolvent mixtures, ...
  62. [62]
    [PDF] Review of the cosolvency models for predicting solubility of drugs in ...
    Water is the main solvent and the aqueous solubility is one of the most important properties of a drug molecule. The drugs are often low soluble in water and ...Missing: enhancement | Show results with:enhancement
  63. [63]
    role of solute hydrophobicity on the solubilization effect - PubMed
    Jun 30, 2010 · The hydrophobicity of the drug has a direct impact on the solubilization obtained in the water-cosolvent mixtures. However, the role of ...
  64. [64]
    Vapor-Liquid Equilibria by UNIFAC Group Contribution. Revision ...
    Aug 7, 2025 · The UNIFAC group contribution method for estimating liquid phase activity coefficients is widely used for phase equilibrium calculations.Missing: LLE | Show results with:LLE
  65. [65]
    Advancing thermodynamic group-contribution methods by machine ...
    Jan 15, 2025 · Among GC methods, UNIFAC stands out as the most sophisticated and widely adopted approach for predicting activity coefficients in liquid ...
  66. [66]
    Analysis of the UNIFAC-Type Group-Contribution Models at the ...
    The Modified UNIFAC model for predicting activity coeffs. presented in this work is based on the well-known UNIFAC model. Two changes are introduced in Modified ...
  67. [67]
    Activity Coefficients of Sodium Chloride in Water–Ethanol Mixtures
    Activity coefficients of NaCl were determined in water–ethanol solvents, in the range 5–20% (w/w) ethanol, from emf data. The molalities varied.
  68. [68]
    Modeling of the solid-liquid equilibrium of NaCl, KCl and NH 4 Cl in ...
    Jan 15, 2021 · The present work studies of NaCl, KCl and NH 4 Cl solubility in binary liquid solutions containing water and ethanol in the temperature range from 288.15 K to ...
  69. [69]
    [PDF] Electrochemistry-based Battery Modeling for Prognostics
    In an ideal condition the activity coefficients are unity and so the last term disappears, resulting in the classical form of the Nernst equation. In the case ...<|separator|>
  70. [70]
    [PDF] Lecture #4 of 17 - UCI Department of Chemistry
    Oct 31, 2023 · Half reactions, at non-unity activity, obey the Nernst equation… Take ∆G = ∆G0 + RT ln Q and use the relation ∆G = −nFE,. But first ...
  71. [71]
    Measuring the Salt Activity Coefficient in Lithium-Battery Electrolytes
    Aug 6, 2025 · The mean activity coefficient γ ± is associated with the deviation of an electrolyte solution from its ideal behavior and may be obtained by ...
  72. [72]
    Optimization for maximum specific energy density of a lithium-ion ...
    Sep 24, 2020 · Average molar activity coefficient. F. Faraday's constant, 96,487 (C ... Multiscale design optimization of lithium ion batteries using adjoint ...
  73. [73]
    Correlations from Ion Pairing and the Nernst-Einstein Equation
    We present a new approximation to ionic conductivity well suited to dynamical atomic-scale simulations, based on the Nernst-Einstein equation.
  74. [74]
    [PDF] Ion Transport in Organic Electrolyte Solutions for Lithium-ion ... - lidsen
    Sep 27, 2021 · Accurate transport properties must be evaluated for designing and/or optimization of lithium-ion and other metal-ion batteries. In this review, ...<|separator|>
  75. [75]
    pH Electrode Theory - Cole-Parmer
    Oct 15, 2018 · pH = -log a · (1) ; aH+ = γH+ • M · (2) ; γH+ = activity coefficient of H · MH+ = molarity of H+ in mols/liter ; H2O = H+ + OH · (3) ; K = (aH+ • aH-) / ...
  76. [76]
    [PDF] Effect of composition and nonideal solution behavior on desalination ...
    Apr 1, 2013 · As discussed, the ideal solution approximation is commonly used in order to avoid the problem of evaluating activity coefficients entirely [8–12] ...<|control11|><|separator|>
  77. [77]
    Implications for Fuel-Cell Operation with Low Relative Humidities
    Jun 14, 2021 · We investigate the effect of proton activity on the open-circuit voltage (OCV) of a H 2 /H 2 cell by individually controlling RH A and RH C at ≤30%.
  78. [78]
    Temperature Dependence of Oxygen Reduction Activity at Pt/Nb ...
    Apr 16, 2021 · We have examined the temperature dependence of the activities for the oxygen reduction reaction (ORR) at Pt catalysts supported on Nb-doped SnO 2 (Pt/Nb-SnO 2 ...