Combining rules

Last updated

In computational chemistry and molecular dynamics, the combination rules or combining rules are equations that provide the interaction energy between two dissimilar non-bonded atoms, usually for the part of the potential representing the van der Waals interaction. [1] In the simulation of mixtures, the choice of combining rules can sometimes affect the outcome of the simulation. [2]

Contents

Combining rules for the Lennard-Jones potential

The Lennard-Jones Potential is a mathematically simple model for the interaction between a pair of atoms or molecules. One of the most common forms is

where ε is the depth of the potential well, σ is the finite distance at which the inter-particle potential is zero, r is the distance between the particles. The potential reaches a minimum, of depth ε, when r = 21/6σ ≈ 1.122σ.

Lorentz-Berthelot rules

The Lorentz rule was proposed by H. A. Lorentz in 1881: [3]

The Lorentz rule is only analytically correct for hard sphere systems. Intuitively, since loosely reflect the radii of particle i and j respectively, their averages can be said to be the effective radii between the two particles at which point repulsive interactions become severe.

The Berthelot rule (Daniel Berthelot, 1898) is given by: [4]

.

Physically, this arises from the fact that is related to the induced dipole interactions between two particles. Given two particles with instantaneous dipole respectively, their interactions correspond to the products of . An arithmetic average of and will not however, result in the average of the two dipole products, but the average of their logarithms would be.

These rules are the most widely used and are the default in many molecular simulation packages, but are not without failings. [5] [6] [7]

Waldman-Hagler rules

The Waldman-Hagler rules are given by: [8]

and

Fender-Halsey

The Fender-Halsey combining rule is given by [9]

Kong rules

The Kong rules for the Lennard-Jones potential are given by: [10]

Others

Many others have been proposed, including those of Tang and Toennies [11] Pena, [12] [13] Hudson and McCoubrey [14] and Sikora (1970). [15]

Combining rules for other potentials

Good-Hope rule

The Good-Hope rule for MieLennard‐Jones or Buckingham potentials is given by: [16]

Hogervorst rules

The Hogervorst rules for the Exp-6 potential are: [17]

and

Kong-Chakrabarty rules

The Kong-Chakrabarty rules for the Exp-6 potential are: [18]

and

Other rules for that have been proposed for the Exp-6 potential are the Mason-Rice rules [19] and the Srivastava and Srivastava rules (1956). [20]

Equations of state

Industrial equations of state have similar mixing and combining rules. These include the van der Waals one-fluid mixing rules

and the van der Waals combining rule, which introduces a binary interaction parameter ,

.

There is also the Huron-Vidal mixing rule, and the more complex Wong-Sandler mixing rule, which equates excess Helmholtz free energy at infinite pressure between an equation of state and an activity coefficient model (and thus with liquid excess Gibbs free energy).

Related Research Articles

<span class="mw-page-title-main">Lennard-Jones potential</span> Model of intermolecular interactions

In computational chemistry, molecular physics, and physical chemistry the Lennard-Jones potential is an intermolecular pair potential. Out of all the intermolecular potentials, the Lennard-Jones potential is probably the one that has been the most extensively studied. It is considered an archetype model for simple yet realistic intermolecular interactions.

<span class="mw-page-title-main">Poisson's ratio</span> Measure of material deformation perpendicular to loading

In materials science and solid mechanics, Poisson's ratio (nu) is a measure of the Poisson effect, the deformation of a material in directions perpendicular to the specific direction of loading. The value of Poisson's ratio is the negative of the ratio of transverse strain to axial strain. For small values of these changes, is the amount of transversal elongation divided by the amount of axial compression. Most materials have Poisson's ratio values ranging between 0.0 and 0.5. For soft materials, such as rubber, where the bulk modulus is much higher than the shear modulus, Poisson's ratio is near 0.5. For open-cell polymer foams, Poisson's ratio is near zero, since the cells tend to collapse in compression. Many typical solids have Poisson's ratios in the range of 0.2–0.3. The ratio is named after the French mathematician and physicist Siméon Poisson.

Linear elasticity is a mathematical model of how solid objects deform and become internally stressed due to prescribed loading conditions. It is a simplification of the more general nonlinear theory of elasticity and a branch of continuum mechanics.

In differential geometry, the Einstein tensor is used to express the curvature of a pseudo-Riemannian manifold. In general relativity, it occurs in the Einstein field equations for gravitation that describe spacetime curvature in a manner that is consistent with conservation of energy and momentum.

<span class="mw-page-title-main">Chiral model</span> Model of mesons in the massless quark limit

In nuclear physics, the chiral model, introduced by Feza Gürsey in 1960, is a phenomenological model describing effective interactions of mesons in the chiral limit (where the masses of the quarks go to zero), but without necessarily mentioning quarks at all. It is a nonlinear sigma model with the principal homogeneous space of a Lie group as its target manifold. When the model was originally introduced, this Lie group was the SU(N), where N is the number of quark flavors. The Riemannian metric of the target manifold is given by a positive constant multiplied by the Killing form acting upon the Maurer–Cartan form of SU(N).

In general relativity, the Gibbons–Hawking–York boundary term is a term that needs to be added to the Einstein–Hilbert action when the underlying spacetime manifold has a boundary.

The J-integral represents a way to calculate the strain energy release rate, or work (energy) per unit fracture surface area, in a material. The theoretical concept of J-integral was developed in 1967 by G. P. Cherepanov and independently in 1968 by James R. Rice, who showed that an energetic contour path integral was independent of the path around a crack.

<span class="mw-page-title-main">Electromagnetic stress–energy tensor</span>

In relativistic physics, the electromagnetic stress–energy tensor is the contribution to the stress–energy tensor due to the electromagnetic field. The stress–energy tensor describes the flow of energy and momentum in spacetime. The electromagnetic stress–energy tensor contains the negative of the classical Maxwell stress tensor that governs the electromagnetic interactions.

<span class="mw-page-title-main">Intraclass correlation</span> Descriptive statistic

In statistics, the intraclass correlation, or the intraclass correlation coefficient (ICC), is a descriptive statistic that can be used when quantitative measurements are made on units that are organized into groups. It describes how strongly units in the same group resemble each other. While it is viewed as a type of correlation, unlike most other correlation measures, it operates on data structured as groups rather than data structured as paired observations.

<span class="mw-page-title-main">Viscoplasticity</span> Theory in continuum mechanics

Viscoplasticity is a theory in continuum mechanics that describes the rate-dependent inelastic behavior of solids. Rate-dependence in this context means that the deformation of the material depends on the rate at which loads are applied. The inelastic behavior that is the subject of viscoplasticity is plastic deformation which means that the material undergoes unrecoverable deformations when a load level is reached. Rate-dependent plasticity is important for transient plasticity calculations. The main difference between rate-independent plastic and viscoplastic material models is that the latter exhibit not only permanent deformations after the application of loads but continue to undergo a creep flow as a function of time under the influence of the applied load.

In statistical mechanics, the eight-vertex model is a generalisation of the ice-type (six-vertex) models; it was discussed by Sutherland, and Fan & Wu, and solved by Baxter in the zero-field case.

Chapman–Enskog theory provides a framework in which equations of hydrodynamics for a gas can be derived from the Boltzmann equation. The technique justifies the otherwise phenomenological constitutive relations appearing in hydrodynamical descriptions such as the Navier–Stokes equations. In doing so, expressions for various transport coefficients such as thermal conductivity and viscosity are obtained in terms of molecular parameters. Thus, Chapman–Enskog theory constitutes an important step in the passage from a microscopic, particle-based description to a continuum hydrodynamical one.

In theoretical chemistry, the Buckingham potential is a formula proposed by Richard Buckingham which describes the Pauli exclusion principle and van der Waals energy for the interaction of two atoms that are not directly bonded as a function of the interatomic distance . It is a variety of interatomic potentials.

<span class="mw-page-title-main">Riemann–Silberstein vector</span> Complex vector of electromagnetic fields

In mathematical physics, in particular electromagnetism, the Riemann–Silberstein vector or Weber vector named after Bernhard Riemann, Heinrich Martin Weber and Ludwik Silberstein, is a complex vector that combines the electric field E and the magnetic field B.

In the Newman–Penrose (NP) formalism of general relativity, independent components of the Ricci tensors of a four-dimensional spacetime are encoded into seven Ricci scalars which consist of three real scalars , three complex scalars and the NP curvature scalar . Physically, Ricci-NP scalars are related with the energy–momentum distribution of the spacetime due to Einstein's field equation.

Reynolds stress equation model (RSM), also referred to as second moment closures are the most complete classical turbulence model. In these models, the eddy-viscosity hypothesis is avoided and the individual components of the Reynolds stress tensor are directly computed. These models use the exact Reynolds stress transport equation for their formulation. They account for the directional effects of the Reynolds stresses and the complex interactions in turbulent flows. Reynolds stress models offer significantly better accuracy than eddy-viscosity based turbulence models, while being computationally cheaper than Direct Numerical Simulations (DNS) and Large Eddy Simulations.

Charge transport mechanisms are theoretical models that aim to quantitatively describe the electric current flow through a given medium.

<span class="mw-page-title-main">Dual photon</span> Hypothetical particle dual to the photon

In theoretical physics, the dual photon is a hypothetical elementary particle that is a dual of the photon under electric–magnetic duality which is predicted by some theoretical models, including M-theory.

The Galilei-covariant tensor formulation is a method for treating non-relativistic physics using the extended Galilei group as the representation group of the theory. It is constructed in the light cone of a five dimensional manifold.

<span class="mw-page-title-main">Mie potential</span>

The Mie potential is an interaction potential describing the interactions between particles on the atomic level. It is mostly used for describing intermolecular interactions, but at times also for modeling intramolecular interaction, i.e. bonds.

References

  1. Halgren, Thomas A. (September 1992). "The representation of van der Waals (vdW) interactions in molecular mechanics force fields: potential form, combination rules, and vdW parameters". Journal of the American Chemical Society. 114 (20): 7827–7843. doi:10.1021/ja00046a032.
  2. Desgranges, Caroline; Delhommelle, Jerome (14 March 2014). "Evaluation of the grand-canonical partition function using expanded Wang-Landau simulations. III. Impact of combining rules on mixtures properties". The Journal of Chemical Physics. 140 (10): 104109. Bibcode:2014JChPh.140j4109D. doi:10.1063/1.4867498. PMID   24628154.
  3. Lorentz, H. A. (1881). "Ueber die Anwendung des Satzes vom Virial in der kinetischen Theorie der Gase". Annalen der Physik. 248 (1): 127–136. Bibcode:1881AnP...248..127L. doi:10.1002/andp.18812480110.
  4. Daniel Berthelot "Sur le mélange des gaz", Comptes rendus hebdomadaires des séances de l’Académie des Sciences, 126 pp. 1703-1855 (1898)
  5. DELHOMMELLE, JÉRÔME; MILLIÉ, PHILIPPE (20 April 2001). "Inadequacy of the Lorentz-Berthelot combining rules for accurate predictions of equilibrium properties by molecular simulation". Molecular Physics. 99 (8): 619–625. Bibcode:2001MolPh..99..619D. doi:10.1080/00268970010020041. S2CID   94931352.
  6. Boda, Dezső; Henderson, Douglas (20 October 2008). "The effects of deviations from Lorentz–Berthelot rules on the properties of a simple mixture". Molecular Physics. 106 (20): 2367–2370. Bibcode:2008MolPh.106.2367B. doi:10.1080/00268970802471137. S2CID   94289505.
  7. Song, W.; Rossky, P. J.; Maroncelli, M. (2003). "Modeling alkane+perfluoroalkane interactions using all-atom potentials: Failure of the usual combining rules". The Journal of Chemical Physics. 119 (17): 9145–9162. Bibcode:2003JChPh.119.9145S. doi:10.1063/1.1610435.
  8. Waldman, Marvin; Hagler, A.T. (September 1993). "New combining rules for rare gas van der waals parameters". Journal of Computational Chemistry. 14 (9): 1077–1084. doi:10.1002/jcc.540140909. S2CID   16732612.
  9. Fender, B. E. F.; Halsey, G. D. (1962). "Second Virial Coefficients of Argon, Krypton, and Argon-Krypton Mixtures at Low Temperatures". The Journal of Chemical Physics. 36 (7): 1881–1888. Bibcode:1962JChPh..36.1881F. doi:10.1063/1.1701284.
  10. Kong, Chang Lyoul (1973). "Combining rules for intermolecular potential parameters. II. Rules for the Lennard-Jones (12–6) potential and the Morse potential". The Journal of Chemical Physics. 59 (5): 2464–2467. Bibcode:1973JChPh..59.2464K. doi:10.1063/1.1680358.
  11. Tang, K. T.; Toennies, J. Peter (March 1986). "New combining rules for well parameters and shapes of the van der Waals potential of mixed rare gas systems". Zeitschrift für Physik D. 1 (1): 91–101. Bibcode:1986ZPhyD...1...91T. doi:10.1007/BF01384663. S2CID   122224768.
  12. Diaz Peña, M. (1982). "Combination rules for intermolecular potential parameters. I. Rules based on approximations for the long-range dispersion energy". The Journal of Chemical Physics. 76 (1): 325–332. Bibcode:1982JChPh..76..325D. doi:10.1063/1.442726.
  13. Diaz Peña, M. (1982). "Combination rules for intermolecular potential parameters. II. Rules based on approximations for the long-range dispersion energy and an atomic distortion model for the repulsive interactions". The Journal of Chemical Physics. 76 (1): 333–339. Bibcode:1982JChPh..76..333D. doi:10.1063/1.442727.
  14. Hudson, G. H.; McCoubrey, J. C. (1960). "Intermolecular forces between unlike molecules. A more complete form of the combining rules". Transactions of the Faraday Society. 56: 761. doi:10.1039/TF9605600761.
  15. Sikora, P T (November 1970). "Combining rules for spherically symmetric intermolecular potentials". Journal of Physics B. 3 (11): 1475–1482. Bibcode:1970JPhB....3.1475S. doi:10.1088/0022-3700/3/11/008.
  16. Good, Robert J. (1970). "New Combining Rule for Intermolecular Distances in Intermolecular Potential Functions". The Journal of Chemical Physics. 53 (2): 540–543. Bibcode:1970JChPh..53..540G. doi: 10.1063/1.1674022 .
  17. Hogervorst, W. (January 1971). "Transport and equilibrium properties of simple gases and forces between like and unlike atoms". Physica. 51 (1): 77–89. Bibcode:1971Phy....51...77H. doi:10.1016/0031-8914(71)90138-8.
  18. Kong, Chang Lyoul; Chakrabarty, Manoj R. (October 1973). "Combining rules for intermolecular potential parameters. III. Application to the exp 6 potential". The Journal of Physical Chemistry. 77 (22): 2668–2670. doi:10.1021/j100640a019.
  19. Mason, Edward A.; Rice, William E. (1954). "The Intermolecular Potentials of Helium and Hydrogen". The Journal of Chemical Physics. 22 (3): 522. Bibcode:1954JChPh..22..522M. doi:10.1063/1.1740100.
  20. Srivastava, B. N.; Srivastava, K. P. (1956). "Combination Rules for Potential Parameters of Unlike Molecules on Exp-Six Model". The Journal of Chemical Physics. 24 (6): 1275. Bibcode:1956JChPh..24.1275S. doi:10.1063/1.1742786.