Fact-checked by Grok 2 weeks ago

Combining rules

Combining rules are mathematical formulations used in and to derive the parameters of intermolecular potentials—such as the collision diameter (σ) and well depth (ε) in the —for interactions between dissimilar atoms or molecules from the corresponding parameters for like interactions. These rules enable the efficient parameterization of force fields by minimizing the need for extensive quantum mechanical calculations on every possible atomic pair, facilitating simulations of complex systems like biomolecules, liquids, and materials. The most common combining rules include the Lorentz–Berthelot rules, which compute σij as the arithmetic mean ((σii + σjj)/2) and εij as the geometric mean (√(εii εjj)) for unlike pairs i and j. The geometric mean rule applies geometric averaging to both parameters (σij = √(σii σjj) and εij = √(εii εjj)), while the Waldman–Hagler rules use a sixth-power mean for σij and a modified expression for εij to better account for polarizability differences (σij = [(σii6 + σjj6)/2]1/6 and εij = √(εii εjj) σii3 σjj3 / σij6) . These approaches are integral to popular force fields such as OPLS, AMBER, CHARMM, and GROMOS, where they model van der Waals forces in non-bonded interactions. The selection of a combining rule influences the fidelity of simulated properties, including densities, vapor pressures, and free energies, with studies showing variations up to several percent across rules for liquids. In applications ranging from to , Lorentz–Berthelot rules are often the default due to their simplicity and historical validation against experimental data for and simple fluids, though alternatives like may perform better for polar or hydrogen-bonding systems. Ongoing research optimizes these rules through and benchmarks to enhance predictive accuracy in diverse chemical environments.

Fundamentals

Definition and Purpose

Combining rules are mathematical approximations used in molecular simulations to estimate the interaction parameters for unlike pairs of atoms or molecules from the known parameters of like pairs. In multi-component systems, these rules derive cross-interaction parameters, such as the well depth ε_{ij} and the collision diameter σ_{ij}, without requiring direct experimental determination for every possible pair. This approach is essential for pair potentials that model intermolecular forces, including the , which approximates van der Waals interactions between neutral particles through a balance of repulsive and attractive terms. The primary purpose of combining rules is to reduce the number of parameters that must be specified or fitted in force fields, thereby lowering and enabling efficient simulations of complex mixtures. For a with N distinct types, explicitly defining all unlike-pair interactions would necessitate up to N(N-1)/2 additional parameters, which becomes prohibitive for large N; combining rules mitigate this by relying solely on the N like-pair parameters, promoting in and methods. This simplification facilitates the development of transferable force fields applicable to diverse chemical environments while maintaining reasonable accuracy for predicted thermodynamic and transport properties. Combining rules are particularly valuable in modeling binary mixtures of simple fluids, such as noble gases like argon-krypton or argon-xenon, where cross-interactions significantly influence phase equilibria, diffusion coefficients, and . In these cases, the rules allow simulations to capture behavior using parameters optimized for pure components, avoiding the need for extensive experimental data on mixed systems.

Historical Development

The development of combining rules began in the late amid efforts to extend kinetic theory to gas mixtures, where simplifying assumptions were needed for unlike-pair interactions in hard-sphere models. In 1881, Hendrik Antoon Lorentz introduced the for molecular diameters, σ_ij = (σ_ii + σ_jj)/2, to approximate collision properties in mixtures without detailed quantum mechanical insights, motivated by the need to fit transport data like viscosities. This additive rule laid the groundwork for handling size differences in classical models. Seventeen years later, in 1898, Daniel Berthelot extended this framework by proposing the geometric mean for attractive energies, ε_ij = √(ε_ii ε_jj), drawing from empirical observations of solubilities and properties to capture forces in van der Waals equations. Together, these Lorentz-Berthelot rules became the foundational standard for unlike-pair parameters in potentials like Lennard-Jones, prioritizing computational simplicity over exact quantum derivations. By the mid-20th century, limitations in predicting mixture properties such as second virial coefficients prompted refinements based on scattering experiments. In 1962, B. E. F. Fender and G. D. Halsey derived harmonic-mean variants for energy parameters to better align with data, emphasizing improved accuracy for low-temperature behaviors like . This shift reflected growing experimental precision, moving motivations from basic averaging to targeted fitting of virial coefficients and viscosities. In the late 20th century, further adjustments addressed discrepancies in thermodynamic predictions. Chang Lyoul Kong's 1973 rules incorporated hard-sphere corrections to the , motivated by vapor pressure deviations in simulations of simple fluids, yielding closer matches to experimental phase equilibria without additional parameters. Later, in 1993, Marvin Waldman and Arnold T. Hagler proposed modifications accounting for in rare-gas systems, driven by the need for better nonbonded force fields in biomolecular modeling, where size and shape asymmetries affected binding energies. These advancements highlighted a transition toward rules supporting efficiency while enhancing fidelity to diverse datasets. Entering the , critiques underscored ongoing challenges, with Jérôme Delhommelle and Philippe Millié's analysis demonstrating that standard Lorentz–Berthelot rules often failed to predict properties accurately in simulations of rare gas mixtures, leading to minor refinements rather than entirely new paradigms. Motivations evolved from early empirical fitting of transport coefficients to broader demands for computational scalability in complex systems, though no major novel rules emerged post-2000, indicating relative maturity in the field.

Combining Rules for the Lennard-Jones Potential

Lorentz–Berthelot Rules

The Lorentz–Berthelot rules represent the most widely adopted combining rules for estimating the unlike-pair parameters in the , particularly for mixtures of non-polar molecules such as . These rules combine the finite distance at which the interparticle potential reaches zero, denoted as \sigma_{ij}, via an , and the depth of the , \epsilon_{ij}, via a . This approach assumes additivity in the repulsive core sizes while treating attractive dispersion forces as multiplicative, providing a straightforward method for modeling intermolecular interactions in molecular simulations without requiring experimental mixture data. The specific formulations are given by the Lorentz rule for the size parameter: \sigma_{ij} = \frac{\sigma_{ii} + \sigma_{jj}}{2} and the Berthelot rule for the energy parameter: \epsilon_{ij} = \sqrt{\epsilon_{ii} \epsilon_{jj}} where \sigma_{ii} and \epsilon_{ii} are the like-pair parameters for species i, and similarly for species j. These expressions derive from foundational assumptions in classical statistical mechanics: the arithmetic averaging for \sigma_{ij} stems from the additivity of hard-sphere diameters in the limit of infinite repulsion, as originally proposed by Lorentz for kinetic theory of gases, ensuring that the effective collision diameter in mixtures aligns with linear superposition of atomic sizes. In contrast, the geometric mean for \epsilon_{ij} arises from the pairwise additivity of London dispersion energies, which scale proportionally with the product of polarizabilities and ionization potentials of the interacting atoms, leading to a multiplicative form for the attractive well depth as justified by Berthelot's early work on corresponding states theory. Historically, these rules emerged in the early 20th century to model noble gas mixtures, with Lorentz's contribution dating to 1886 in the context of gas kinetic theory and Berthelot's to 1898 for thermodynamic properties of van der Waals fluids; their combination became standard by the 1920s for practical applications in equation-of-state predictions. The rules have since been integral to the van der Waals equation of state extensions for mixtures, facilitating estimates of critical properties and phase behavior in binary systems. Their simplicity—requiring only pure-component parameters—has led to widespread implementation in molecular dynamics software, such as GROMACS, where they serve as the default for Lennard-Jones interactions in biomolecular and materials simulations, enabling efficient computation for large-scale systems without additional parameterization.

Waldman–Hagler Rules

The Waldman–Hagler rules represent a set of modified combining rules for the , designed to improve the estimation of cross-interaction parameters between dissimilar atoms by introducing a size-dependent in the well depth calculation. These rules were introduced in 1993 by Marvin Waldman and T. Hagler in response to discrepancies between standard combining rules and experimental data for rare gas interactions, with motivation drawn from calculations aimed at developing accurate force fields for organic molecules. The rules define the interaction distance parameter at the potential minimum, r_{m,ij}, using a sixth-power : r_{m,ij} = \left[ \frac{r_{m,ii}^6 + r_{m,jj}^6}{2} \right]^{1/6} The well depth, \varepsilon_{ij}, is then computed as a that incorporates the size parameters: \varepsilon_{ij} = 2 \sqrt{\varepsilon_{ii} \varepsilon_{jj}} \frac{r_{m,ii}^3 r_{m,jj}^3}{r_{m,ii}^6 + r_{m,jj}^6} This formulation effectively averages the contributions related to the attractive dispersion term (proportional to r_m^6) while adjusting the well depth to account for differences in sizes. Compared to the Lorentz–Berthelot rules, which use a simple for \varepsilon_{ij}, the Waldman–Hagler approach reduces overestimation of the well depth for pairs with significant size disparities, yielding more realistic interaction strengths in heterogeneous systems. Validation through comparisons with calculations has demonstrated improved agreement for interaction energies in dissimilar pairs.

Fender–Halsey Rules

The combining rules represent an empirical approach to determining the unlike-pair parameters for the Lennard-Jones (12-6) potential in mixtures of rare gases, emphasizing a averaging for the energy depth to better capture low-temperature interactions. Proposed by B. E. F. Fender and G. D. in 1962, these rules were developed through analysis of experimental data on systems, particularly to address discrepancies observed in mixture properties under standard geometric-mean approximations. The core formula for the unlike energy parameter \epsilon_{ij} employs the : \epsilon_{ij} = \frac{2 \epsilon_{ii} \epsilon_{jj}}{\epsilon_{ii} + \epsilon_{jj}} This yields smaller \epsilon_{ij} values compared to the , which aligns more closely with observed behaviors in dissimilar interactions. For the size parameter, \sigma_{ij} follows the convention: \sigma_{ij} = \frac{\sigma_{ii} + \sigma_{jj}}{2} These expressions were fitted specifically to second virial coefficient measurements for pure (80–125 K), (105–140 K), and argon-krypton mixtures (105–125 K), where experimental values exceeded prior predictions by 5–15%. In applications to argon-krypton mixtures, the rules demonstrate superior performance over the Lorentz–Berthelot geometric-mean approach by providing more accurate second virial coefficients at low temperatures, as the better reproduces the experimental mixed-gas data. This empirical adjustment also enhances predictions of vapor-liquid equilibria in systems, including liquid densities, where standard rules often overestimate interaction strengths. Overall, the method prioritizes fidelity to rare-gas experimental benchmarks, making it particularly suitable for thermodynamic modeling of simple mixtures.

Kong Rules

The Kong rules were introduced by Chang Lyoul Kong in 1973 to improve the accuracy of intermolecular potential parameters for unlike pairs in the Lennard-Jones (12-6) potential, particularly in addressing discrepancies in critical constants for mixtures of simple molecules. These rules were developed based on an atomic distortion model that better correlates experimental data for interactions. By separating the treatment of the attractive and repulsive components, the approach accounts for their differing physical underpinnings, with the attractive term dominated by dispersion forces and the repulsive term by Pauli exclusion and overlap effects. The , which forms the foundation for these rules, models pairwise interactions as u(r) = 4\epsilon \left[ \left( \frac{\sigma}{r} \right)^{12} - \left( \frac{\sigma}{r} \right)^6 \right], where \epsilon is the well depth and \sigma is the finite distance at which the potential is zero. For the attractive r^{-6} term, the combining rule employs a : \epsilon_{ij} \sigma_{ij}^6 = \sqrt{ \epsilon_{ii} \sigma_{ii}^6 \, \epsilon_{jj} \sigma_{jj}^6 }, reflecting the additive nature of dispersion energies from fluctuating dipoles. For the repulsive r^{-12} term, a modified is used to incorporate effects: \epsilon_{ij} \sigma_{ij}^{12} = \left[ \frac{ (\epsilon_{ii} \sigma_{ii}^{12})^{1/13} + (\epsilon_{jj} \sigma_{jj}^{12})^{1/13} }{2} \right]^{13}, where the exponent 1/13 derives from the power-law form and ensures consistency with the potential's steep repulsive wall. Validation studies have demonstrated the superiority of rules over traditional combining methods for predicting thermodynamic properties, such as vapor-liquid equilibria in mixtures; for instance, simulations of n-/CO₂ systems yield diagrams in excellent agreement with experimental data.

Other Variants

The Böhm-Ahlrichs combining rules, proposed in 1982 by Hans-Joachim Böhm and Reinhart Ahlrichs, provide a method for estimating repulsive interactions in intermolecular potentials, derived from self-consistent field (SCF) calculations of overlap effects. These rules are particularly useful for short-range repulsions in systems where quantum overlap is significant, such as in heteronuclear pairs. They have been applied in studies of van der Waals complexes to improve modeling of repulsive walls. The Hudson-McCoubrey rules, introduced in , derive ε_ij from ionization potentials I, providing a quantum-based alternative rooted in London dispersion theory for unlike molecules. The key relation is \frac{\epsilon_{ij}}{\sqrt{\epsilon_{ii} \epsilon_{jj}}} = \frac{2 I_{ii} I_{jj}}{I_{ii} + I_{jj}}, with the right-hand side as the of the ionization potentials, which corrects the geometric mean for electronic structure differences and has been extended to Mie potentials for broader fluid mixture predictions. This approach improves accuracy in vapor-liquid equilibria for binary systems where quantum effects dominate, such as noble gas mixtures, though it requires experimental data. These variants address limitations in empirical rules by integrating quantum properties, unlike the non-quantum rules which separate repulsive and attractive exponents without such corrections. Recent studies as of 2024 have evaluated these and similar rules using benchmarks, highlighting their performance in diverse systems.
VariantKey FeaturePros vs. Lorentz–BerthelotCons vs. Lorentz–BerthelotTypical Application
Böhm-AhlrichsUses for repulsive parametersBetter for systems with significant quantum overlap (e.g., improved repulsive modeling in van der Waals complexes)Requires SCF-derived data; more complexShort-range repulsions in molecular complexes
Hudson-McCoubreyUses ionization potential for ε_ijEnhanced quantum accuracy in dispersion-dominated mixtures (e.g., better agreement in rare gas VLE)Needs potentials; sensitive to sigma combiningBinary fluid phase equilibria with electronic effects

Combining Rules for Other Potentials

Good–Hope Rule

The Good–Hope rule is a combining rule for estimating the intermolecular parameter \sigma_{ij} between unlike or molecular in generalized potential functions, such as the Mie potential, which extends the Lennard-Jones form with arbitrary repulsive and attractive exponents m and n. It employs the , given by \sigma_{ij} = \sqrt{\sigma_{ii} \sigma_{jj}}, where \sigma_{ii} and \sigma_{jj} are the self-interaction parameters for i and j, respectively. This approach contrasts with the Lorentz rule's for \sigma, providing a more consistent scaling for repulsive interactions. Proposed in 1970 by Robert J. Good and Christopher J. Hope, the rule was developed specifically to address limitations in prior distance-combining methods for intermolecular potential functions, including Mie–Lennard-Jones types. It enhances accuracy in modeling pairwise interactions where the finite size of particles plays a dominant role, outperforming additive schemes like the Lorentz rule by better preserving the geometric of molecular overlaps. In particular, it improves consistency in the hard-sphere limit, where repulsive cores approximate impenetrable spheres, leading to more reliable predictions of packing and structural properties. The Good–Hope rule finds application in simulations of soft-matter systems, such as colloidal suspensions and polymer melts, where Mie-like potentials capture the dominant repulsive contributions from particle cores or chain segments. Its simplicity and effectiveness in these contexts have made it a preferred choice for parameterizing cross-interactions in models emphasizing over long-range attractions.

Hogervorst Rules

The Hogervorst combining rules provide a method for estimating the parameters of the Exp-6 intermolecular potential between unlike atoms based on the parameters of like-atom interactions, particularly suited for modeling mixtures. The Exp-6 potential, which features an exponential repulsive term and an inverse-sixth power attractive term, is expressed as U(r) = \epsilon \frac{(1 - 1/\alpha)^\alpha}{1 - 6/\alpha} \left[ \frac{6}{\alpha} \exp\left[\alpha \left(1 - \frac{r}{\rho}\right)\right] - \left(\frac{r}{\rho}\right)^{-6} \right], where \epsilon is the well depth, \rho is the finite distance at which the potential is zero, and \alpha controls the steepness of the repulsion. These rules were developed by W. Hogervorst in 1971 through a reanalysis of and properties of simple gases, including and data for , , , and , to refine parameters for the Exp-6 model with a focus on exponential repulsion in interactions. The key combining expressions are the for the well depth, \epsilon_{ij} = \frac{2 \epsilon_{ii} \epsilon_{jj}}{\epsilon_{ii} + \epsilon_{jj}}, and the arithmetic mean for the steepness parameter, \alpha_{ij} = \frac{\alpha_{ii} + \alpha_{jj}}{2}. This approach for \epsilon_{ij} bears similarity to the harmonic mean used in the Fender–Halsey rules. The Hogervorst rules excel in applications involving high-density fluids, where the accurate representation of short-range exponential repulsion is crucial for predicting properties like densities and melting points in noble gas systems.

Kong–Chakrabarty Rules

The rules were developed in 1973 by Lyoul and Manoj R. Chakrabarty to derive cross-interaction parameters for the exp-6 potential in molecular , extending the from 's prior work on Lennard-Jones potentials to enhance predictions of . The exp-6 potential models interatomic interactions as U(r) = \frac{\epsilon_{ij}}{2\alpha_{ij} - 6} \left[ \frac{6}{\alpha_{ij}} \exp\left[\alpha_{ij} \left(1 - \frac{r}{\rho_{ij}}\right)\right] - \left(2\alpha_{ij} - 6\right) \left(\frac{\rho_{ij}}{r}\right)^6 \right], where \epsilon_{ij} is the well depth, \alpha_{ij} is the steepness parameter, and \rho_{ij} is the radial distance scale. Unlike simpler potentials with two parameters, the exp-6 form requires dedicated combining rules for all three unlike-pair parameters to account for the distinct repulsive (exponential) and attractive (dispersion) contributions. The Kong–Chakrabarty rules achieve this through coupled equations that assume an for the repulsive component and a for the attractive component, yielding more accurate cross-parameters than basic averaging methods. The key relations, often expressed in terms of the zero-potential distance \sigma_{ij} (related to \rho_{ij} via \sigma_{ij} = \rho_{ij} \left( \frac{6}{\alpha_{ij}} \right)^{1/\alpha_{ij}}), are \left[ \frac{\epsilon_{ij} \alpha_{ij} e^{\alpha_{ij}}}{(\alpha_{ij} - 6) \sigma_{ij}} \right]^{2 \sigma_{ij} / \alpha_{ij}} = \left[ \frac{\epsilon_{ii} \alpha_{ii} e^{\alpha_{ii}}}{(\alpha_{ii} - 6) \sigma_{ii}} \right]^{\sigma_{ii} / \alpha_{ii}} \left[ \frac{\epsilon_{jj} \alpha_{jj} e^{\alpha_{jj}}}{(\alpha_{jj} - 6) \sigma_{jj}} \right]^{\sigma_{jj} / \alpha_{jj}} for the well depth and steepness, and \frac{\sigma_{ij}}{\alpha_{ij}} = \frac{1}{2} \left( \frac{\sigma_{ii}}{\alpha_{ii}} + \frac{\sigma_{jj}}{\alpha_{jj}} \right) for the length scale, with \rho_{ij} derived accordingly. These equations must typically be solved iteratively for \epsilon_{ij}, \alpha_{ij}, and \rho_{ij}. These rules improve upon earlier approaches by better capturing the parameter coupling in the , leading to superior agreement with experimental data for mixture properties such as in rare gas and diatomic gas systems like oxygen-rare gas collisions. They have been applied in force fields for where exp-6 terms describe non-polar components, and they address shortcomings of geometric-mean rules by incorporating differential repulsion steepness, which indirectly accounts for quantum mechanical effects like electronic overlap absent in basic formulations.

Rules for the Buckingham Potential

The Buckingham potential models pairwise interactions as U(r) = A \exp(-r / \rho) - C / r^6, where the exponential term captures short-range repulsion and the r^{-6} term accounts for long-range dispersion attraction. For interactions between unlike atom types i and j, standard combining rules apply the geometric mean to the pre-exponential factor A and dispersion coefficient C, while using the arithmetic mean for the decay length \rho:
A_{ij} = \sqrt{A_{ii} A_{jj}},
\rho_{ij} = \frac{\rho_{ii} + \rho_{jj}}{2},
C_{ij} = \sqrt{C_{ii} C_{jj}}.
These rules ensure physical consistency in multi-component systems by approximating the cross-term parameters from pure-component values, reducing the need for extensive empirical fitting.
These combining rules trace their origins to variations developed in the , evolving from the Born-Mayer-Huggins potential form introduced in the 1930s and 1940s to describe exponential repulsion in ionic systems. Seminal parametrizations, such as those by Fumi and Tosi for alkali halide crystals, adopted the for attractive and repulsive amplitudes alongside the for \rho to match experimental properties like cohesive energies and elastic constants. The rules are particularly valuable in simulations of ionic crystals, such as NaCl or KBr, and simple metals like alkali metals, where the exponential repulsion accurately represents overlap effects at short distances critical for structural stability and defect formation. This approach provides depth for solid-state applications, an area often underexplored in general molecular dynamics literature focused on van der Waals systems.

Applications

Equations of State

Combining rules play a crucial role in extending equations of state (EOS) developed for pure fluids to multicomponent mixtures by enabling the calculation of cross-interaction parameters. In cubic EOS such as the van der Waals equation, the pressure-volume-temperature relation for a mixture is given by \left( P + \frac{a_m}{V_m^2} \right) (V_m - b_m) = RT, where the mixture parameters a_m and b_m are derived from pure-component values a_i, b_i using mixing rules that incorporate combining rules for unlike pairs. Specifically, a_m = \sum_i \sum_j x_i x_j a_{ij} and b_m = \sum_i x_i b_i, with a_{ij} often computed via the geometric mean \sqrt{a_i a_j} (related to the energy parameter \epsilon_{ij}) and b_{ij} via the arithmetic mean (b_i + b_j)/2 (related to the size parameter \sigma_{ij}). These rules ensure thermodynamic consistency and allow prediction of mixture properties like fugacity and phase behavior without additional adjustable parameters beyond pure-component data. Historically, the for the attraction parameter in der Waals-type EOS originated from empirical fitting of data by in the late 19th century, who proposed it to reconcile observed deviations in gas compressibility from behavior. This approach, later formalized as part of the Lorentz-Berthelot rules, was motivated by the need to match experimental pressure-volume relations for binary gas mixtures, providing a simple yet effective way to estimate cross-terms from pure-gas EOS parameters. In the context of corresponding-states , these rules facilitate scaling of mixture properties using reduced variables, particularly for mixtures where intermolecular forces are dominated by dispersion. For instance, applying Lorentz-Berthelot rules to helium-neon or argon-krypton mixtures yields compressibility factors Z = PV/RT that align closely with experimental values at moderate pressures, capturing the slight positive deviations due to unlike-pair attractions being weaker than like-pair averages. A key analytical application lies in the virial expansion of the EOS, where the second virial coefficient B governs low-density behavior: PV/RT = 1 + B/V + \cdots. For mixtures, B = \sum_i x_i^2 B_{ii} + 2 \sum_{i<j} x_i x_j B_{ij}, with B_{ij} derived from combining rules applied to the pair potential parameters. For a Lennard-Jones potential, B_{ij} \propto \sigma_{ij}^3 \left[1 - \left(\frac{\epsilon_{ij}}{kT}\right)^{1/2} f\left(\frac{\epsilon_{ij}}{kT}\right)\right], where \sigma_{ij} = (\sigma_i + \sigma_j)/2 and \epsilon_{ij} = \sqrt{\epsilon_i \epsilon_j} follow Lorentz-Berthelot, and f is a temperature-dependent function from integration over the potential. This form highlights how combining rules influence mixture non-ideality, with the geometric mean for energy often underpredicting B_{ij} for polar or size-asymmetric systems, leading to adjusted compressibility in noble gas binaries like argon-oxygen. Experimental data for argon-oxygen mixtures at 100-300 K show that Lorentz-Berthelot-based virial coefficients reproduce compressibility factors in alignment with measurements, underscoring their utility for cryogenic air component modeling. In modern thermodynamic modeling, combining rules are integral to advanced EOS like Statistical Associating Fluid Theory (SAFT), where they parameterize the dispersion and chain-formation contributions for mixtures. For non-associating mixtures, Lorentz-Berthelot rules compute cross-segment interactions in the perturbed-chain SAFT (PC-SAFT) variant, enabling predictions of phase envelopes and densities for complex fluids. Examples include hydrocarbon-noble gas systems, where these rules, combined with one-fluid mixing, accurately describe vapor-liquid equilibria. This integration bridges molecular-level parameters to macroscopic EOS, enhancing predictive power for industrial mixtures without relying on binary-specific fittings.

Molecular Dynamics Simulations

In (MD) simulations, combining rules are essential for parameterizing non-bonded interactions in heterogeneous systems, such as those involving multiple atom types from different molecular species. Force fields like OPLS-AA employ geometric combining rules for both the Lennard-Jones well depth (ε_ij = √(ε_ii ε_jj)) and collision diameter (σ_ij = √(σ_ii σ_jj)), ensuring compatibility with organic molecules and biomolecules. In contrast, the Lorentz–Berthelot rules, which use an for σ_ij ((σ_ii + σ_jj)/2) and a geometric mean for ε_ij, serve as the default in many general-purpose simulations. Popular MD software packages allow users to select these rules via input parameters; for instance, LAMMPS uses the pair_modify mix command to specify geometric, arithmetic, or other mixing options during pair style setup, while defines combination rule types (e.g., type 2 for Lorentz–Berthelot, type 3 for OPLS geometric) in the topology file. During MD workflows, combining rules enable on-the-fly computation of interaction parameters for unlike atom pairs as neighbor lists are built, facilitating efficient handling of mixtures without pre-specifying all cross-terms. In LAMMPS, mixing coefficients are calculated at simulation initialization for each pair type I-J where explicit coefficients are absent, integrating seamlessly into the pair potential evaluation loop. constructs the Lennard-Jones parameter matrix using the selected rules prior to force computation, applying them in the Verlet scheme to update forces for all pairwise interactions within the cut-off distance. This approach is particularly valuable in complex systems like solvated biomolecules, where diverse atom types (e.g., carbon, oxygen, nitrogen) interact dynamically. The use of combining rules significantly enhances computational efficiency by reducing the number of required parameters from O(N^2) to O(N) for N atom types, avoiding the need to parameterize every possible unlike pair explicitly. For example, in protein-ligand binding simulations, these rules allow modeling of interactions between protein residues and ligand atoms using OPLS-AA parameters, enabling trajectories that capture binding dynamics and solvation effects without excessive parameterization. Validation often involves comparing simulated properties, such as diffusion coefficients in mixed solvents, to experimental data; discrepancies can arise from rule choice, highlighting the need for system-specific testing. In the , combining rules remain foundational even in advanced MD frameworks incorporating machine-learned (MLIPs), where they are applied during data generation or to extend models to mixtures. For instance, Lorentz–Berthelot rules are used to generate consistent training data for MLIPs in zeolite-organic systems, bridging classical force fields with quantum-accurate predictions. This integration supports scalable simulations of multicomponent materials while preserving the simplicity of classical parameter mixing.

Limitations and Comparisons

Known Shortcomings

The Lorentz–Berthelot combining rules, while widely used for their simplicity, exhibit notable shortcomings in accurately modeling intermolecular interactions, particularly in mixtures involving significant size disparities or polar components. These rules assume a for the well depth parameter \epsilon_{ij} and an for the size parameter \sigma_{ij}, which often leads to overestimation of \epsilon_{ij} by approximately 15% for size-mismatched pairs, resulting in excessively attractive unlike-pair interactions that deviate from experimental data. This overestimation is especially pronounced in systems like rare gas binaries (e.g., ), where simulations using Lorentz–Berthelot rules yield substantial errors in liquid–vapor phase equilibria and thermodynamic properties compared to molecular beam experiments. In polar and associating fluids, the rules fail to account for electronic effects such as charge transfer, induction, and , which alter interaction strengths beyond simple averaging. For instance, in –methane mixtures, application of Lorentz–Berthelot rules with common force fields like TraPPE-UA and /E overestimates the of methane by about 1.02 kJ/mol (roughly 12% relative to the experimental value of 8.37 kJ/mol at standard conditions), leading to underprediction of methane in . This inadequacy stems from neglecting the polar nature of water, where hydrogen bonding and hydrophobic effects dominate, causing systematic errors in properties without adjustable corrections. Furthermore, the classical nature of Lorentz–Berthelot rules ignores quantum mechanical contributions, such as dispersion forces influenced by electronic structure variations, which become critical in low-temperature or highly asymmetric systems. Delhommelle and Millié (2001) highlighted these issues through Gibbs ensemble simulations, showing that the rules produce deviations in mixture densities and critical points for Ne–Ar and Ne–Kr systems, underscoring their limitations for transferable predictions. While alternatives like the offer partial improvements by adjusting for size mismatches in some non-polar cases, they do not fully resolve failures in polar environments.

Comparative Performance

Various studies have evaluated the performance of combining rules through benchmarks on mixtures, where experimental data for properties like second virial cross coefficients (B_{12}) and phase equilibria are available for validation. The Lorentz–Berthelot () rules often overestimate the unlike-pair well depth ε_{ij} by approximately 15%, leading to errors in B_{12} of up to 20% for mixtures such as Ne-Ar and Ar-Kr. In contrast, the rules adjust ε_{ij} using a composition-dependent factor, reducing these errors to under 5% and providing better alignment with experimental virial coefficients and coexistence curves. Studies indicate that rules lead to significant errors in B_{12} for mixtures like Ne-Ar (around 18%), Ar-Kr (15%), and Ne-Kr (up to 22%), while rules reduce these to 3-6%, improving agreement with experimental data. For predictions, metrics such as mean absolute deviation (MAD) in critical temperatures (T_c) and densities (ρ_c) are commonly used. Simulations employing rules exhibit MADs of 5-10% in T_c for binaries, while and Waldman–Hagler rules lower this to 2-4%, particularly for critical densities where size mismatches amplify errors. In evaluations of vapor-liquid equilibria, the Fender–Halsey rules demonstrate lower deviations (MAD ~3% in liquid densities) for condensed-phase properties compared to arithmetic means in . For organic mixtures, weighted rules like Waldman–Hagler perform better, achieving MADs under 5% in enthalpies of due to their handling of differences. Emerging trends indicate that harmonic-based rules (e.g., Fender–Halsey) excel in liquid-state simulations by better capturing volume effects, whereas weighted formulations (e.g., Waldman–Hagler) suit heterogeneous organics with varying electronegativities. However, a post-2020 gap persists in comprehensive coverage, with limited adoption of approaches to derive adaptive combining rules from quantum data, despite initial efforts in optimizing LJ parameters for broader applicability. However, as of 2025, frameworks are increasingly applied to refine LJ parameters and combining rules for broader applicability.

References

  1. [1]
    Influence of the Lennard-Jones Combination Rules on the ...
    Mar 15, 2023 · There is extensive literature discussing the accuracy and limitations of commonly used combination rules, (64−68) including comparisons with ...Introduction · Methodology · Results · Supporting Information
  2. [2]
    Classical and reactive molecular dynamics: Principles and ...
    Typically, the parameters of ε and σ for different pairs of atoms are based on combining rules of Lorentz-Berthelot [85,86]. For example, the corresponding ...Classical And Reactive... · 2. Methodology Of Molecular... · 2.3. Force Fields
  3. [3]
    Influence of the Lennard-Jones Combination Rules on the ... - NIH
    The three combination rules perform comparatively well, with the GM and LB results being more similar to each other and slightly more accurate compared to ...
  4. [4]
    Recent developments of first‐principles force fields - Sun - 2017
    Sep 7, 2016 · Therefore, the combining rule is used to reduce the number of fitting parameters. ... To improve the accuracy, MP2 has been combined with SAPT for ...
  5. [5]
    Test of Combining Rules for Intermolecular Distances. Potential ...
    Jul 1, 1971 · A new combining rule for equilibrium distances in intermolecular potential functions, d12 = (d11d22)1/2, is tested by means of experimental ...Missing: definition | Show results with:definition
  6. [6]
  7. [7]
    Evaluation of the grand-canonical partition function using expanded ...
    Mar 11, 2014 · We also find the Kong and Waldman-Hagler to be much better alternatives than the LB rules. The use of the LB rules results in much stronger ...
  8. [8]
    Improved Models for Short-Range Repulsion in ab Initio Force Fields
    Jun 23, 2016 · However, a Waldman–Hagler style analysis (58) (Supporting Information) suggests instead that a more suitable exponent is given by the ...
  9. [9]
  10. [10]
    Combining rules page on SklogWiki - a wiki for statistical mechanics ...
    Jan 16, 2015 · The combining rules are geometric expressions designed to provide the interaction energy between two dissimilar non-bonded atoms.
  11. [11]
  12. [12]
    [PDF] Investigating Local Structuring of Organic Semiconductors via ...
    The Good-hope rule (also known as geometric mixing rule) has been employed in this thesis. The LJ potential can be divided into two parts of an attractive ...
  13. [13]
  14. [14]
  15. [15]
    Combining rules for intermolecular potential parameters. III ...
    Combining rules for intermolecular potential parameters. ... Solubility of NaCl in water and its melting point by molecular dynamics in the slab geometry and a ...
  16. [16]
    Study of oxygen-rare gas collisions by depolarized Raman scattering
    ... Kong-Chakrabarty rules to determine the parameters of potentials leads to a very good overall agreement between theory and all the present experimental data.
  17. [17]
    Ar-Ne - Interatomic Potentials Repository
    Notes: The exp-6 pair potential has an exponential repulsive term and 1/6 power attractive term. Parameters due to W. Hogervorst for pure Ar and Ne are used. A ...
  18. [18]
    Chapter 5: Mixing and Combining Rules - Books
    Nov 1, 2010 · It is to the discussion of combining rules that we now turn by way of a digression regarding intermolecular potentials. In the absence of ...
  19. [19]
    Chapter 3: The Virial Equation of State - Books
    Nov 1, 2010 · The first of these relations is the Berthelot rule, and the second is the Lorentz rule. Combining rules are also required for other parameters ...
  20. [20]
    Taking Another Look at the van der Waals Equation of State–Almost ...
    Aug 2, 2019 · Thus, the Lorentz–Berthelot combining rules are far less successful compared to the conventional combining rules for real fluids in the context ...
  21. [21]
    Quadratic mixing rules for equations of state - NASA ADS
    In this report we relate the quadratic mixing rules to the rigorous mixing rules for virial coefficients of mixtures. The virial equation of state having a ...
  22. [22]
    Fundamental equation of state for mixtures of nitrogen, oxygen, and ...
    May 1, 2024 · A fundamental equation of state in terms of the Helmholtz energy is presented for mixtures of nitrogen, oxygen, and argon at any composition. ...
  23. [23]
    Toward Advanced, Predictive Mixing Rules in SAFT Equations of State
    A novel mixing rule that bridges the Statistical Associating Fluid Theory (SAFT)-type equations of state and activity coefficient models is proposed.
  24. [24]
    Development of OPLS-AA Force Field Parameters for 68 Unique ...
    Mar 12, 2009 · The geometric combining rules regularly used for the Lennard−Jones coefficients are employed: σij = (σiiσjj)1/2 and εij = (εiiεjj)1/2. (28) ...Introduction · Computational Methods · Results and Discussion · Conclusions
  25. [25]
    Non-bonded interactions - GROMACS 2025.3 documentation
    The non-bonded interactions contain a repulsion term, a dispersion term, and a Coulomb term. The repulsion and dispersion term are combined in either the ...
  26. [26]
    pair_modify command - LAMMPS documentation
    Aug 3, 2022 · Modify the parameters of the currently defined pair style. If the pair style is hybrid or hybrid/overlay, then the specified parameters are by default modified.
  27. [27]
    Practical considerations for Molecular Dynamics - GitHub Pages
    This combining rule was developed specifically for simulation of noble gases. Hybrid (the Lorentz–Berthelot for H and the Waldman–Hagler for other elements).
  28. [28]
    Building, Simulating, and Analyzing Protein–Ligand Systems in ...
    Feb 11, 2025 · An open-source toolkit that enables the preparation and simulation of protein–ligand complexes in OpenMM, along with the subsequent analysis of protein–ligand ...
  29. [29]
    Data Generation for Machine Learning Interatomic Potentials and ...
    Nov 21, 2024 · in a consistent manner using the Lorentz-Berthelot combining rules as for other parts of the TraPPE force field. The TraPPE-zeo force field ...
  30. [30]
  31. [31]
    [PDF] arXiv:0904.4436v1 [physics.chem-ph] 28 Apr 2009
    Apr 28, 2009 · parameters from the combining rules differ only very litte from the Lorentz rule. ... Lorentz, Annalen der Physik 12 (1881) 127–136. [23] D.
  32. [32]
  33. [33]
    Harnessing Deep Learning for Optimization of Lennard-Jones ...
    Apr 1, 2022 · A deep learning (DL)-based parametrization framework is developed, allowing for sampling of wide ranges of LJ parameters targeting experimental condensed phase ...