ART

The Lennard-Jones potential (also termed the LJ potential, 6-12 potential, or 12-6 potential) is a mathematically simple model that approximates the intermolecular potential energy between a pair of neutral atoms or molecules as metals or cyclic alkanes.[1][2] This interatomic potential emerged from Max Born's treatise on lattice enthalpies[3] and was elaborated in 1924 by John Lennard-Jones.[4] A common expression of the LJ potential is

\( {\displaystyle V_{\text{LJ}}=4\varepsilon \left[\left({\frac {\sigma }{r}}\right)^{12}-\left({\frac {\sigma }{r}}\right)^{6}\right]=\varepsilon \left[\left({\frac {r_{\text{m}}}{r}}\right)^{12}-2\left({\frac {r_{\text{m}}}{r}}\right)^{6}\right],} \)

where \( \varepsilon \) is the depth of the potential well, σ {\displaystyle \sigma } \sigma is the finite distance at which the inter-particle potential is zero, r is the distance between the particles, and r \( r_m\) is the distance at which the potential reaches its minimum. At r m {\displaystyle r_{m}} r_m, the potential function has the value − ε {\displaystyle \varepsilon } \varepsilon . The distances are related as \( {\displaystyle r_{m}=2^{1/6}\sigma \approx 1.122\sigma } \) . These parameters can be fitted to reproduce experimental data on lattice parameters ( \( r_m \)), surface energies ( \( \varepsilon \) ), or accurate quantum chemistry calculations.[2] Due to its computational simplicity and interpretability, the Lennard-Jones potential is used extensively in computer simulations. It is also part of more detailed interatomic potentials such as CHARMM,[5] AMBER,[6] OPLS-AA,[7] CVFF,[8] DREIDING,[9] and IFF.[10]

The Lennard-Jones potential is also used with different exponents, for example, in 9-6 form:

\( {\displaystyle V_{\text{LJ}}=\varepsilon \left[2\left({\frac {r_{\text{m}}}{r}}\right)^{9}-3\left({\frac {r_{\text{m}}}{r}}\right)^{6}\right]}, \)

in force fields such as CFF,[11] PCFF,[12] and IFF.[10] The interpretation of \( r_m \) as atomic diameter and \( \varepsilon \) as polarizability and a measure for internal cohesion is the same as for the 12-6 Lennard-Jones potential.[2]

Explanation

The \( {\displaystyle r^{12}} \)term, which is the repulsive term, describes Pauli repulsion at short ranges due to overlapping electron orbitals, and the \( {\displaystyle r^{6}} \) term, which is the attractive long-range term, describes attraction at long ranges (van der Waals force, or dispersion force).

Differentiating the LJ potential with respect to r gives an expression for the net inter-molecular force between 2 molecules. This inter-molecular force may be attractive or repulsive, depending on the value of r. When r is very small, the molecules repel each other.

Whereas the functional form of the attractive term has a clear physical justification, the repulsive term has no theoretical justification. It is used because it approximates the Pauli repulsion well and is more convenient due to the relative computing efficiency of calculating \( {\displaystyle r^{12}} \) as the square of \( {\displaystyle r^{6}} \).

The LJ potential is a relatively good approximation. Due to its simplicity, it is often used to describe the properties of gases and to model dispersion and overlap interactions in molecular models. It is especially accurate for noble gas atoms and is a good approximation at long and short distances for neutral atoms and molecules.

The lowest-energy arrangement of an infinite number of atoms described by a Lennard-Jones potential is a hexagonal close-packing. On raising temperature, the lowest-free-energy arrangement becomes cubic close packing, and then liquid. Under pressure, the lowest-energy structure switches between cubic and hexagonal close packing.[13] Real materials include body-centered cubic structures also.[14]

The Lennard-Jones (12,6) potential was improved by the Buckingham potential (exp-6) later proposed by Richard Buckingham, incorporating an extra parameter and the repulsive part is replaced by an exponential function:[15]

\( {\displaystyle V_{\text{B}}=\gamma \left[e^{-r/r_{1}}-\left({\frac {r_{0}}{r}}\right)^{6}\right].} \)

Other more recent methods, such as the Stockmayer potential, describe the interaction of molecules more accurately. Quantum chemistry methods, Møller–Plesset perturbation theory, coupled cluster method, or full configuration interaction can give extremely accurate results, but require large computing cost.
Alternative expressions

There are many different ways to formulate the Lennard-Jones potential. Some common forms follow.
AB form

This form is a simplified formulation that is used by some simulation software:

\( {\displaystyle V_{\text{LJ}}(r)={\frac {A}{r^{12}}}-{\frac {B}{r^{6}}},} \)

where, \( A=4\varepsilon \sigma ^{12} \) and \( B=4\varepsilon \sigma ^{6 \)}. Conversely, \( \sigma ={\sqrt[{6}]{\frac {A}{B}}} \) and \( \varepsilon ={\frac {B^{2}}{4A}} \). This is the form in which Lennard-Jones wrote the 12-6 potential.[16]

A more mathematically general form, which contains an extra variable n is

\( {\displaystyle V_{\text{LJ}}(r)=\varepsilon \left(\left({\frac {r_{0}}{r}}\right)^{2n}-2\left({\frac {r_{0}}{r}}\right)^{n}\right),} \)

where ε {\displaystyle \varepsilon } \varepsilon is the bonding energy of the molecule (the energy required to separate the atoms). The exponent n could be related to the spring constant k (at \( r_{0} \), where \( {\displaystyle V=-\varepsilon } \) ) as

\({\displaystyle k=2\varepsilon \left({\frac {n}{r_{0}}}\right)^{2},} \)

from where n {\displaystyle n} n can be calculated if k is known. Normally the harmonic states are known, \( {\displaystyle \Delta E=\hbar \omega } \), where \( \omega ={\sqrt {k/m}} \) . n can also be related to the group velocity in a crystal, v \( {\displaystyle v_{\text{g}}} \)

\( {\displaystyle v_{\text{g}}={\frac {a\cdot n}{r_{0}}}{\sqrt {\frac {\varepsilon }{m}}},} \)

where a is the lattice distance, and m is the mass of a particle.

Truncated and shifted form

To save computing time and satisfy the minimum image convention when using periodic boundary conditions, the Lennard-Jones potential is often truncated at a cut-off distance of \( {\displaystyle r_{\mathrm {c} }=2.5\sigma } \), where

\( {\displaystyle \displaystyle V_{\text{LJ}}(r_{\text{c}})=V_{\text{LJ}}(2.5\sigma )=4\varepsilon \left[\left({\frac {\sigma }{2.5\sigma }}\right)^{12}-\left({\frac {\sigma }{2.5\sigma }}\right)^{6}\right]\approx -0.0163\varepsilon ,} \) (1)

i.e., at \( {\displaystyle r_{\mathrm {c} }=2.5\sigma } \), the Lennard-Jones potential VLJ is about 1/60 of its minimum value, \( \varepsilon \) (the depth of the potential well). Beyond \( {\displaystyle r_{\text{c}}} \), the truncated potential is set to zero.

To avoid a jump discontinuity at \( {\displaystyle r_{\text{c}}} \), the LJ potential must be shifted upward a little, so that the truncated potential would be zero exactly at the cut-off distance, \( {\displaystyle r_{\text{c}}} \).

For clarity, let \( {\displaystyle V_{\text{LJ}}} \) denote the LJ potential as defined above, i.e.,

\({\displaystyle \displaystyle V_{\text{LJ}}(r)=4\varepsilon \left[\left({\frac {\sigma }{r}}\right)^{12}-\left({\frac {\sigma }{r}}\right)^{6}\right].} \)
(2)

Then the truncated Lennard-Jones potential \( {\displaystyle V_{{\text{LJ}}_{\text{trunc}}}} \) is defined as follows[17]

\( {\displaystyle \displaystyle V_{{\text{LJ}}_{\text{trunc}}}(r):={\begin{cases}V_{\text{LJ}}(r)-V_{\text{LJ}}(r_{\text{c}})&{\text{for }}r\leq r_{\text{c}}\\0&{\text{for }}r>r_{\text{c}}.\end{cases}}} \) (3)

It can be easily verified that VLJtrunc(rc) = 0, thus eliminating the jump discontinuity at \( {\displaystyle r=r_{\mathrm {c} }} \). Although the value of the (unshifted) Lennard Jones potential at \( {\displaystyle r=r_{\mathrm {c} }=2.5\sigma } \) is rather small, the effect of the truncation can be significant; for instance, on the gas–liquid critical point.[18][19] The potential energy can often be corrected for this effect in a mean-field manner by adding so-called tail corrections.[20] The stability of crystal structure is extremely sensitive to the truncation: a wide variety of close packed stackings may have the lowest energy.[21]

Extensions for polarizability

Lennard-Jones potentials have been combined with bonded potentials to describe virtual electrons and image charges in metals.[22] The combined potential reproduces lattice parameters, surface energies, the complete image potentials, hydration energies, and mechanical properties without additional fit parameters in higher accuracy than density functional theory calculations.
Dimensionless (reduced) units
LJ dimensionless (reduced) units Property Symbol Reduced form
Length \( r^{{*}} \) \( {\displaystyle {\frac {r}{\sigma }}} \)
Time \({\displaystyle t^{*}} \) \( {\displaystyle {\frac {t}{\tau }}=t{\sqrt {\frac {\epsilon }{m\sigma ^{2}}}}} \)
Temperature \( T^{*} \) \( {\displaystyle {\frac {k_{B}T}{\epsilon }}} \)
Force \( f^{*} \) \( {\displaystyle {\frac {f\sigma }{\epsilon }}} \)
Energy \( {\displaystyle \phi ^{*}} \) \( {\displaystyle {\frac {\phi }{\epsilon }}} \)
Pressure \( P^{*} \) \( {\displaystyle {\frac {P\sigma ^{3}}{\epsilon }}} \)
Number density \( {\displaystyle N^{*}} \) \( {\displaystyle N\sigma ^{3}} \)
Density \( \rho^{*} \) \( {\displaystyle \rho \sigma ^{3}} \)
Surface tension \( {\displaystyle \gamma ^{*}} \) \( {\displaystyle {\frac {\gamma \sigma ^{2}}{\epsilon }}} \)

Dimensionless units can be defined based on the Lennard-Jones potential, which are convenient for molecular dynamics simulations. From a numerical point, the advantages of dimensionless units include computing values which are closer to unity, using simplified equations and being able to easily scale the results.[23]

When the Lennard-Jones potential is used for molecular dynamics simulations, the most convenient dimensionless units are obtained by choosing length σ, mass m and energy ε as the scaling factors for the various physical properties.[23]
Limitations

The LJ potential has only two parameters (A and B), which determine the length and energy scales. The potential is therefore limited in how accurately it can be fitted to the properties of any real material.
The bonding of the LJ potential has no directionality: the potential is spherically symmetric.
The sixth-power term models effectively the dipole–dipole interactions due to electron dispersion in noble gases (London dispersion forces), but it does not represent other kinds of bonding well. The twelfth-power term appearing in the potential is chosen for its ease of calculation for simulations (by squaring the sixth-power term) and is not theoretically or physically based.
The potential diverges when two atoms approach one another. This may create instabilities that require special treatment in molecular dynamics simulations.

12-6-4 Lennard-Jones potential for charged particles

For simulations of charged particles, e.g. ions, an additional \( r^{4} \) term has been proposed to account for dipole interaction,[24][25] giving rise to a so-called 12-6-4 potential.

\( {\displaystyle V_{\text{12-6-4-LJ}}=\varepsilon \left[\left({\frac {r_{\text{m}}}{r}}\right)^{12}-2\left({\frac {r_{\text{m}}}{r}}\right)^{6}\right]-{\frac {C}{r^{4}}},} \)

where C is a system-dependent constant.
See also

Morse/Long-range potential
Molecular mechanics
Embedded atom model
Morse potential
Force field (chemistry)
Comparison of force field implementations
Heterogenous catalysis
Hydrogen spillover
Physisorption
Virial expansion

References

Muñoz-Muñoz, Y. Mauricio; Guevara-Carrion, Gabriela; Llano-Restrepo, Mario; Vrabec, Jadran (2015), "Lennard-Jones force field parameters for cyclic alkanes from cyclopropane to cyclohexane", Fluid Phase Equilibria, 404: 150–160, doi:10.1016/j.fluid.2015.06.033
Heinz, Hendrik; Vaia, R. A.; Farmer, B. L.; Naik, R. R. (2008-11-06). "Accurate Simulation of Surfaces and Interfaces of Face-Centered Cubic Metals Using 12−6 and 9−6 Lennard-Jones Potentials". The Journal of Physical Chemistry C. 112 (44): 17281–17290. doi:10.1021/jp801931d. ISSN 1932-7447.
Born, Max (1919). "Eine Thermochemische Anwendung der Gittertheorie". Verhandlungen der Deutschen Physikalischen Gesellschaft. 21: 13–24.
Lennard-Jones, J. E. (1924), "On the Determination of Molecular Fields", Proc. R. Soc. Lond. A, 106 (738): 463–477, Bibcode:1924RSPSA.106..463J, doi:10.1098/rspa.1924.0082
Huang, Jing; MacKerell, Alexander D. (2013-09-30). "CHARMM36 all-atom additive protein force field: Validation based on comparison to NMR data". Journal of Computational Chemistry. 34 (25): 2135–2145. doi:10.1002/jcc.23354. PMC 3800559. PMID 23832629.
Wang, Junmei; Wolf, Romain M.; Caldwell, James W.; Kollman, Peter A.; Case, David A. (2004-07-15). "Development and testing of a general amber force field". Journal of Computational Chemistry. 25 (9): 1157–1174. doi:10.1002/jcc.20035. ISSN 0192-8651. PMID 15116359.
Jorgensen, William L.; Maxwell, David S.; Tirado-Rives, Julian (January 1996). "Development and Testing of the OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic Liquids". Journal of the American Chemical Society. 118 (45): 11225–11236. doi:10.1021/ja9621760. ISSN 0002-7863.
Dauber-Osguthorpe, Pnina; Roberts, Victoria A.; Osguthorpe, David J.; Wolff, Jon; Genest, Moniqe; Hagler, Arnold T. (1988). "Structure and energetics of ligand binding to proteins:Escherichia coli dihydrofolate reductase-trimethoprim, a drug-receptor system". Proteins: Structure, Function, and Genetics. 4 (1): 31–47. doi:10.1002/prot.340040106. ISSN 0887-3585. PMID 3054871.
Mayo, Stephen L.; Olafson, Barry D.; Goddard, William A. (December 1990). "DREIDING: a generic force field for molecular simulations". The Journal of Physical Chemistry. 94 (26): 8897–8909. doi:10.1021/j100389a010. ISSN 0022-3654.
Heinz, Hendrik; Lin, Tzu-Jen; Kishore Mishra, Ratan; Emami, Fateme S. (2013-02-12). "Thermodynamically Consistent Force Fields for the Assembly of Inorganic, Organic, and Biological Nanostructures: The INTERFACE Force Field". Langmuir. 29 (6): 1754–1765. doi:10.1021/la3038846. ISSN 0743-7463. PMID 23276161.
Sun, Huai; Mumby, Stephen J.; Maple, Jon R.; Hagler, Arnold T. (April 1994). "An ab Initio CFF93 All-Atom Force Field for Polycarbonates". Journal of the American Chemical Society. 116 (7): 2978–2987. doi:10.1021/ja00086a030. ISSN 0002-7863.
Sun, H. (May 1995). "Ab initio calculations and force field development for computer simulation of polysilanes". Macromolecules. 28 (3): 701–712. doi:10.1021/ma00107a006. ISSN 0024-9297.
Barron, T. H. K.; Domb, C. (1955), "On the Cubic and Hexagonal Close-Packed Lattices", Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, 227 (1171): 447–465, Bibcode:1955RSPSA.227..447B, doi:10.1098/rspa.1955.0023, S2CID 95146429
Zhen, Shu; Davies, G. J. (16 August 1983). "Calculation of the Lennard-Jones n–m potential energy parameters for metals". Physica Status Solidi A. 78 (2): 595–605. Bibcode:1983PSSAR..78..595Z. doi:10.1002/pssa.2210780226.
Peter Atkins and Julio de Paula, "Atkins' Physical Chemistry" (8th edn, W. H. Freeman), p. 637.
Lennard-Jones, J. E. (1931). "Cohesion". Proceedings of the Physical Society. 43 (5): 461–482. Bibcode:1931PPS....43..461L. doi:10.1088/0959-5309/43/5/301.
softmatter:Lennard-Jones Potential Archived 2007-10-07 at the Wayback Machine, Soft matter Archived 2007-12-08 at the Wayback Machine, Materials Digital Library Pathway , NIST-Lennard-Jones-fluid-properties.
Smit, B. (1992), "Phase diagrams of Lennard-Jones fluids" (PDF), Journal of Chemical Physics, 96 (11): 8639–8640, Bibcode:1992JChPh..96.8639S, doi:10.1063/1.462271
Di Pierro, M.; Elber, R.; Leimkuhler, B. (2015), "A Stochastic Algorithm for the Isobaric-Isothermal Ensemble with Ewald Summations for all Long Range Forces", Journal of Chemical Theory and Computation, 11 (12): 5624–5637, doi:10.1021/acs.jctc.5b00648, PMC 4890727, PMID 26616351
Frenkel, D. & Smit, B. (2002), Understanding Molecular Simulation (Second ed.), San Diego: Academic Press, ISBN 0-12-267351-4
Pártay, Lívia B.; Ortner, Christoph; Bartók, Albert P.; Pickard, Chris J.; Csányi, Gábor (2017). "Polytypism in the ground state structure of the Lennard-Jonesium". Physical Chemistry Chemical Physics. 19 (29): 19369–19376. arXiv:1705.01751. Bibcode:2017PCCP...1919369P. doi:10.1039/C7CP02923C. PMID 28707687. S2CID 40385351.
Geada, Isidro Lorenzo; Ramezani-Dakhel, Hadi; Jamil, Tariq; Sulpizi, Marialore; Heinz, Hendrik (December 2018). "Insight into induced charges at metal surfaces and biointerfaces using a polarizable Lennard–Jones potential". Nature Communications. 9 (1): 716. doi:10.1038/s41467-018-03137-8. ISSN 2041-1723. PMC 5818522. PMID 29459638.
D. C. Rapaport (1 April 2004). The Art of Molecular Dynamics Simulation. Cambridge University Press. ISBN 978-0-521-82568-9.
Pengfei, Li; Merz Jr., Kenneth M. (2014), "Taking into Account the Ion-induced Dipole Interaction in the Nonbonded Model of Ions", Journal of Chemical Theory and Computation, 10 (1): 289–297, doi:10.1021/ct400751u, PMC 3960013, PMID 24659926

Pengfei, Li; Son, Lin Frank; Merz Jr., Kenneth M. (2015), "Parameterization of Highly Charged Metal Ions Using the 12-6‑4 LJ-Type Nonbonded Model in Explicit Water", Journal of Physical Chemistry B, 119 (3): 883–895, doi:10.1021/jp505875v, PMC 4306492, PMID 25145273

Physics Encyclopedia

World

Index

Hellenica World - Scientific Library

Retrieved from "http://en.wikipedia.org/"
All text is available under the terms of the GNU Free Documentation License