fbpx
Wikipedia

Lennard-Jones potential

In computational chemistry, molecular physics, and physical chemistry the Lennard-Jones potential (also termed the LJ potential or 12-6 potential; named for John Lennard-Jones) 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.[1][2] It is considered an archetype model for simple yet realistic intermolecular interactions.

Figure 1. Graph of the Lennard-Jones potential function: Intermolecular potential energy VLJ as a function of the distance of a pair of particles. The potential minimum is at

The Lennard-Jones potential models soft repulsive and attractive (van der Waals) interactions. Hence, the Lennard-Jones potential describes electronically neutral atoms or molecules.[3][4][5] The commonly used expression for the Lennard-Jones potential is

where r is the distance between two interacting particles, ε is the depth of the potential well (usually referred to as 'dispersion energy'), and σ is the distance at which the particle-particle potential energy V is zero (often referred to as 'size of the particle'). The Lennard-Jones potential has its minimum at a distance of where the potential energy has the value

The Lennard-Jones potential is a simplified model that yet describes the essential features of interactions between simple atoms and molecules: Two interacting particles repel each other at very close distance, attract each other at moderate distance, and do not interact at infinite distance, as shown in Figure 1. The Lennard-Jones potential is a pair potential, i.e. no three- or multi-body interactions are covered by the potential.

Statistical mechanics[6] and computer simulations[7][8] can be used to study the Lennard-Jones potential and to obtain thermophysical properties of the 'Lennard-Jones substance'. The Lennard-Jones substance is often referred to as 'Lennard-Jonesium,'[2] suggesting that it is viewed as a (fictive) chemical element.[9] Moreover, its energy and length parameters can be adjusted to fit many different real substances. Both the Lennard-Jones potential and, accordingly, the Lennard-Jones substance are simplified yet realistic models, such as they accurately capture essential physical principles like the presence of a critical and a triple point, condensation and freezing. Due in part to its mathematical simplicity, the Lennard-Jones potential has been extensively used in studies on matter since the early days of computer simulation.[10][11][12][13] The Lennard-Jones potential is probably still the most frequently studied model potential.[14][9]

The Lennard-Jones potential is usually the standard choice for the development of theories for matter (especially soft-matter) as well as for the development and testing of computational methods and algorithms. Upon adjusting the model parameters ε and σ to real substance properties, the Lennard-Jones potential can be used to describe simple substance (like noble gases) with good accuracy. Furthermore, the Lennard-Jones potential is often used as a building block in molecular models (a.k.a. force fields) for more complex substances.[15][16][17][18][19]

Physical background and mathematical details edit

The Lennard-Jones potential models the two most important and fundamental molecular interactions: The repulsive term (  term) describes the Pauli repulsion at short distances of the interacting particles due to overlapping electron orbitals and the attractive term (  term) describes attraction at long ranged interactions (London dispersion force), which vanish at infinite distance between two particles. The steep repulsive interactions at short distances yield the low compressibility of the solid and liquid phase; the attractive dispersive interactions act stabilizing for the condensed phase, especially the vapor–liquid equilibrium.

The functional form of the attractive term, the exponent '6', has a physical justification, which does not hold as rigorously for the repulsive term with the exponent '12'. The attractive dispersive interactions between simple atoms and molecules are a result of fluctuating partial charges. It has been shown by quantum-chemical calculations that this dispersive contribution has to decay with  .[20][2]

The   term is mainly used because it can be implemented computationally very efficiently as the square of  , which does not hold to the same extent for values other than '12'. Also,   approximates the Pauli repulsion reasonably well. If needed, the Lennard-Jones potential can be generalized using arbitrary exponents instead of 12 and 6; the resulting model is called the Mie potential. The present article exclusively discusses the original (12-6) Lennard-Jones potential.

The Lennard-Jones potential exhibits a pole at  , i.e. the potential energy diverges to  , which can cause instabilities in molecular simulations, e.g. for the sampling of the chemical potential. The Lennard-Jones potential converges to   for  . Hence, from a mathematical standpoint, attractive interactions stay present for infinitely distanced particles. These dispersive 'long-range' interactions have an important influence on several properties of the Lennard-Jones substance, e.g. the pressure or heat capacity in the vicinity of the critical point and the critical point itself. The importance of the long-range interactions were noticed already in the early stages of statistical mechanics.[21] For computer simulations, only finite numbers of particles can be used, which leads to the fact that the potential can only be evaluated up to a finite radius  , which is a so-called finite size effect. There are well-established methods to implicitly consider the thereby neglected long-range contribution for a given observable (details are given below).

It is often claimed that multiple Lennard-Jones potentials and corresponding substances exist depending on the handling of the long-range interactions. This is misleading. There is only one 'Lennard-Jones potential', which is exactly defined by Eq. (1). The Lennard-Jones potential requires the consideration and evaluation of long-range interactions up to very long (actually infinite) distances – at least so that the influence of the truncation has no influence on the observable of interest for the reported decimal places.

The Lennard-Jones potential implies that the particles are point masses with a mass  . Even though the parameter   is often referred to as 'size of the particle', particles interacting with the Lennard-Jones potential have no uniquely defined 'size' – opposite to the hard sphere potential. Particles interacting with the Lennard-Jones potential rather have soft repulsive cores.

The Lennard-Jones model describes the potential intermolecular energy   between two particles based on the outlined principles. Following Newton's mechanics, the actual force   between two interacting particles is simply obtained by negating and differentiating the Lennard-Jones potential with respect to  , i.e.  . Depending on the distance between the two particles, the net force can be either attractive or repulsive.

The Lennard-Jones potential yields a good approximation of intermolecular interactions for many applications: The macroscopic properties computed using the Lennard-Jones potential are in good agreement with experimental data for simple substances like argon on one side and the potential function   is in fair agreement with results from quantum chemistry on the other side. The Lennard-Jones potential gives a good description of molecular interactions in fluid phases, whereas molecular interactions in solid phases are only roughly well described. This is mainly due to the fact that multi-body interactions play a significant role in solid phases, which are not comprised in the Lennard-Jones potential. Therefore, the Lennard-Jones potential is extensively used in soft-matter physics and associated fields, whereas it is less frequently used in solid-state physics. Due to its simplicity, the Lennard-Jones potential is often used to describe the properties of gases and simple fluids and to model dispersive and repulsive interactions in molecular models. It is especially accurate for noble gas atoms and methane. It is furthermore a good approximation for molecular interactions at long and short distances for neutral atoms and molecules. Therefore, the Lennard-Jones potential is very often used as a building block of molecular models of complex molecules, e.g. alkanes or water.[18][22][17] The Lennard-Jones potential can also be used to model the adsorption interactions at solid–fluid interfaces, i.e. physisorption or chemisorption.

It is well accepted, that the main limitations of the Lennard-Jones potential lie in the fact the potential is a pair potential (does not cover multi-body interactions) and that the   exponent term is used for the repulsion. Results from quantum chemistry suggest that a higher exponent than 12 has to be used, i.e. a steeper potential. Furthermore, the Lennard-Jones potential has a limited flexibility, i.e. only the two model parameters   and   can be used for the fitting to describe a real substance.

Numerous intermolecular potentials have been proposed in the past for the modeling of simple soft repulsive and attractive interactions between spherically symmetric particles, i.e. the general shape shown in Figure 1. Examples for other potentials are the Morse potential, the Mie potential,[23] the Buckingham potential and the Tang-Tönnies potential.[24] While some of these may be more suited to modelling real fluids,[25] the simplicity of the Lennard-Jones potential, as well as it's often surprising ability to accurately capture real fluid behaviour, has historically made it the pair-potential of greatest general importance.[26]

Application of the Lennard-Jones potential edit

The Lennard-Jones potential is not only of fundamental importance in computational chemistry and soft-matter physics, but also for the modeling of real substances. The Lennard-Jones potential is frequently used for fundamental studies on the behavior of matter and for elucidating atomistic phenomena. It is also often used for somewhat special use cases, e.g. for studying thermophysical properties of two- or four-dimensional substances[27][28][29] (instead of the classical three spatial directions of our universe).

The Lennard-Jones potential is extensively used for molecular modeling. There are essentially two ways the Lennard-Jones potential can be used for molecular modeling: (1) A real substance atom or molecule is modeled directly by the Lennard-Jones potential, which yields very good results for noble gases and methane, i.e. dispersively interacting spherical particles. In the case of methane, the molecule is assumed to be spherically symmetric and the hydrogen atoms are fused with the carbon atom to a common unit. This simplification can in general also be applied to more complex molecules, but yields usually poor results. (2) A real substance molecule is built of multiple Lennard-Jones interactions sites, which can be connected either by rigid bonds or flexible additional potentials (and eventually also consists of other potential types, e.g. partial charges). Molecular models (often referred to as 'force fields') for practically all molecular and ionic particles can be constructed using this scheme for example for alkanes.

Upon using the first outlined approach, the molecular model has only the two parameters of the Lennard-Jones potential   and   that can be used for the fitting, e.g.   and   are frequently used for argon. Evidently, this approach is only a good approximation for spherical and simply dispersively interacting molecules and atoms. The direct use of the Lennard-Jones potential has the great advantage that simulation results and theories for the Lennard-Jones potential can be used directly. Hence, available results for the Lennard-Jones potential and substance can be directly scaled using the appropriate   and   (see reduced units). The Lennard-Jones potential parameters   and   can in general be fitted to any desired real substance property. In soft-matter physics, usually experimental data for the vapor–liquid phase equilibrium or the critical point are used for the parametrization; in solid-state physics, rather the compressibility, heat capacity or lattice constants are employed.[30][31]

The second outlined approach of using the Lennard-Jones potential as a building block of elongated and complex molecules is far more sophisticated. Molecular models are thereby tailor-made in a sense that simulation results are only applicable for that particular model. This development approach for molecular force fields is today mainly performed in soft-matter physics and associated fields such as chemical engineering, chemistry, and computational biology. A large number of force fields are based on the Lennard-Jones potential, e.g. the TraPPE force field,[18] the OPLS force field,[32] and the MolMod force field[17] (an overview of molecular force fields is out of the scope of the present article). For the state-of-the-art modeling of solid-state materials, more elaborate multi-body potentials (e.g. EAM potentials[33]) are used.

Alternative notations of the Lennard-Jones potential edit

There are several different ways to formulate the Lennard-Jones potential besides Eq. (1). Alternatives are:

AB form

The AB form is frequently used in implementations of simulation software as it is computationally favorable. The Lennard-Jones potential can be written as

 

where,   and  . Conversely,   and  .
This is the form in which Lennard-Jones originally presented the potential named after him.[34]

n-exp form

The n-exp form is a mathematically more general form (see Mie potential) and can be written as

 

where   is the bonding energy of the molecule (the energy required to separate the atoms). Applying a harmonic approximation to the potential minimum (at  ), the exponent   and the energy parameter   can be related to the harmonic spring constant:

 

from which   can be calculated if   is known. If changes in the harmonic states   are known from experiment (for example, from Raman spectroscopy) then  , where   and   is the reduced mass and   is the particle mass, may be used to estimate the spring constant. Alternatively   can be related to  , the group velocity in a crystal, using

 

where   is the lattice distance.[citation needed]

Dimensionless (reduced units) edit

dimensionless (reduced) units
Property Symbol Reduced form
Length    
Time    
Temperature    
Force    
Energy    
Pressure    
Density    
Surface tension    

Dimensionless reduced units can be defined based on the Lennard-Jones potential parameters, which is convenient for molecular simulations. From a numerical point of view, the advantages of this unit system include computing values which are closer to unity, using simplified equations and being able to easily scale the results.[35][7] This reduced units system requires the specification of the size parameter   and the energy parameter   of the Lennard-Jones potential and the mass of the particle  . All physical properties can be converted straightforwardly taking the respective dimension into account, see table. The reduced units are often abbreviated and indicated by an asterisk.

In general, reduced units can also be built up on other molecular interaction potentials that consist of a length parameter and an energy parameter.

Thermophysical properties of the Lennard-Jones substance edit

 
Figure 2. Phase diagram of the Lennard-Jones substance. Correlations and numeric values for the critical point and triple point(s) are taken from Refs.[9][36][14] The star indicates the critical point.[9] The circle indicates the vapor–liquid–solid triple point and the triangle indicates the vapor–solid (fcc)–solid (hcp) triple point.[36][37] The solid lines indicate coexistence lines of two phases.[9][36] The dashed lines indicate the vapor–liquid spinodal.[14]

Thermophysical properties of the Lennard-Jones substance,[2] i.e. particles interacting with the Lennard-Jones potential can be obtained using statistical mechanics. Some properties can be computed analytically, i.e. with machine precision, whereas most properties can only be obtained by performing molecular simulations.[7] The latter will in general be superimposed by both statistical and systematic uncertainties.[38][9][39][40] The virial coefficients can for example be computed directly from the Lennard-potential using algebraic expressions[6] and reported data has therefore no uncertainty. Molecular simulation results, e.g. the pressure at a given temperature and density has both statistical and systematic uncertainties.[38][40] Molecular simulations of the Lennard-Jones potential can in general be performed using either molecular dynamics (MD) simulations or Monte Carlo (MC) simulation. For MC simulations, the Lennard-Jones potential   is directly used, whereas MD simulations are always based on the derivative of the potential, i.e. the force  . These differences in combination with differences in the treatment of the long-range interactions (see below) can influence computed thermophysical properties.[41][42]

Since the Lennard-Jonesium is the archetype for the modeling of simple yet realistic intermolecular interactions, a large number of thermophysical properties were studied and reported in the literature.[9] Computer experiment data of the Lennard-Jones potential is presently considered the most accurately known data in classical mechanics computational chemistry. Hence, such data is also mostly used as benchmark for the validation and testing of new algorithms and theories. The Lennard-Jones potential has been constantly used since the early days of molecular simulations. The first results from computer experiments for the Lennard-Jones potential were reported by Rosenbluth and Rosenbluth[11] and Wood and Parker[10] after molecular simulations on "fast computing machines" became available in 1953.[43] Since then many studies reported data of the Lennard-Jones substance;[9] approximately 50,000 data points are publicly available. The current state of research of thermophysical properties of the Lennard-Jones substance is summarized in the following. The most comprehensive summary and digital database was given by Stephan et al.[9] Presently, no data repository covers and maintains this database (or any other model potential) – the concise data selection stated by the NIST website should be treated with caution regarding referencing[44] and coverage (it contains a small fraction of the available data). Most of the data on NIST website provides non-peer-reviewed data generated in-house by NIST.

Figure 2 shows the phase diagram of the Lennard-Jones fluid. Phase equilibria of the Lennard-Jones potential have been studied numerous times and are accordingly known today with good precision.[36][9][45] Figure 2 shows results correlations derived from computer experiment results (hence, lines instead of data points are shown).

The mean intermolecular interaction of a Lennard-Jones particle strongly depends on the thermodynamic state, i.e. temperature and pressure (or density). For solid states, the attractive Lennard-Jones interaction plays a dominant role – especially at low temperatures. For liquid states, no ordered structure is present compared to solid states. The mean potential energy per particle is negative. For gaseous states, attractive interactions of the Lennard-Jones potential play a minor role – since they are far distanced. The main part of the internal energy is stored as kinetic energy for gaseous states. At supercritical states, the attractive Lennard-Jones interaction plays a minor role. With increasing temperature, the mean kinetic energy of the particles increases and exceeds the energy well of the Lennard-Jones potential. Hence, the particles mainly interact by the potentials' soft repulsive interactions and the mean potential energy per particle is accordingly positive.

Overall, due to the large timespan the Lennard-Jones potential has been studied and thermophysical property data has been reported in the literature and computational resources were insufficient for accurate simulations (to modern standards), a noticeable amount of data is known to be dubious.[9] Nevertheless, in many studies such data is used as reference. The lack of data repositories and data assessment is a crucial element for future work in the long-going field of Lennard-Jones potential research.

Characteristic points and curves edit

The most important characteristic points of the Lennard-Jones potential are the critical point and the vapor–liquid–solid triple point. They were studied numerous times in the literature and compiled in Ref.[9] The critical point was thereby assessed to be located at

  •  
  •  
  •  

The given uncertainties were calculated from the standard deviation of the critical parameters derived from the most reliable available vapor–liquid equilibrium data sets.[9] These uncertainties can be assumed as a lower limit to the accuracy with which the critical point of fluid can be obtained from molecular simulation results.

 
Figure 3. Characteristic curves of the Lennard-Jones substance. The thick black line indicates the vapor–liquid equilibrium; the star indicates the critical point. The brown line indicates the solid–fluid equilibrium. Other black solid lines and symbols indicate Brown's characteristic curves (see text for details) of the Lennard-Jones substance: lines are results from an equation of state, symbols from molecular simulations and triangles exact data in the ideal gas limit obtained from the virial coefficients. Data taken from.[46][47][48]

The triple point is presently assumed to be located at

  •  
  •  
  •  
  •  
  •  

The uncertainties represent the scattering of data from different authors.[36] The critical point of the Lennard-Jones substance has been studied far more often than the triple point. For both the critical point and the vapor–liquid–solid triple point, several studies reported results out of the above stated ranges. The above stated data is the presently assumed correct and reliable data. Nevertheless, the determinateness of the critical temperature and the triple point temperature is still unsatisfactory.

Evidently, the phase coexistence curves (cf. figure 2) are of fundamental importance to characterize the Lennard-Jones potential. Furthermore, Brown's characteristic curves[49] yield an illustrative description of essential features of the Lennard-Jones potential. Brown's characteristic curves are defined as curves on which a certain thermodynamic property of the substance matches that of an ideal gas. For a real fluid,   and its derivatives can match the values of the ideal gas for special  ,   combinations only as a result of Gibbs' phase rule. The resulting points collectively constitute a characteristic curve. Four main characteristic curves are defined: One 0th-order (named Zeno curve) and three 1st-order curves (named Amagat, Boyle, and Charles curve). The characteristic curve are required to have a negative or zero curvature throughout and a single maximum in a double-logarithmic pressure-temperature diagram. Furthermore, Brown's characteristic curves and the virial coefficients are directly linked in the limit of the ideal gas and are therefore known exactly at  . Both computer simulation results and equation of state results have been reported in the literature for the Lennard-Jones potential.[47][9][46][50][51]

Points on the Zeno curve Z have a compressibility factor of unity  . The Zeno curve originates at the Boyle temperature  , surrounds the critical point, and has a slope of unity in the low temperature limit.[46] Points on the Boyle curve B have  . The Boyle curve originates with the Zeno curve at the Boyle temperature, faintly surrounds the critical point, and ends on the vapor pressure curve. Points on the Charles curve (a.k.a. Joule-Thomson inversion curve) have   and more importantly  , i.e. no temperature change upon isenthalpic throttling. It originates at   in the ideal gas limit, crosses the Zeno curve, and terminates on the vapor pressure curve. Points on the Amagat curve A have  . It also starts in the ideal gas limit at  , surrounds the critical point and the other three characteristic curves and passes into the solid phase region. A comprehensive discussion of the characteristic curves of the Lennard-Jones potential is given by Stephan and Deiters.[46]

 
Figure 4. Virial coefficients from the Lennard-Jones potential as a function of the temperature: Second virial coefficient   (top) and third virial coefficient   (bottom). The circle indicates the Boyle temperature  . Results taken from.[46]

Properties of the Lennard-Jones fluid edit

 
Figure 5. Vapor–liquid equilibrium of the Lennard-Jones substance: Vapor pressure (top), saturated densities (middle) and interfacial tension (bottom). Symbols indicate molecular simulation results.[52][9] Lines indicate results from equation of state (and square gradient theory for the interfacial tension).[52][14]

Properties of the Lennard-Jones fluid have been studied extensively in the literature due to the outstanding importance of the Lennard-Jones potential in soft-matter physics and related fields.[2] About 50 datasets of computer experiment data for the vapor–liquid equilibrium have been published to date.[9] Furthermore, more than 35,000 data points at homogeneous fluid states have been published over the years and recently been compiled and assessed for outliers in an open access database.[9]

The vapor–liquid equilibrium of the Lennard-Jones substance is presently known with a precision, i.e. mutual agreement of thermodynamically consistent data, of   for the vapor pressure,   for the saturated liquid density,   for the saturated vapor density,   for the enthalpy of vaporization, and   for the surface tension.[9] This status quo can not be considered satisfactory considering the fact that statistical uncertainties usually reported for single data sets are significantly below the above stated values (even for far more complex molecular force fields).

Both phase equilibrium properties and homogeneous state properties at arbitrary density can in general only be obtained from molecular simulations, whereas virial coefficients can be computed directly from the Lennard-Jones potential.[6] Numerical data for the second and third virial coefficient is available in a wide temperature range.[53][46][9] For higher virial coefficients (up to the sixteenth), the number of available data points decreases with increasing number of the virial coefficient.[54][55] Also transport properties (viscosity, heat conductivity, and self diffusion coefficient) of the Lennard-Jones fluid have been studied frequently,[56][57] but the database is significantly less dense than for homogeneous equilibrium properties like   – or internal energy data. Moreover, a large number of analytical models (equations of state) have been developed for the description of the Lennard-Jones fluid (see below for details).

Properties of the Lennard-Jones solid edit

The database and knowledge for the Lennard-Jones solid is significantly poorer than for the fluid phases, which is mainly due to the fact that the Lennard-Jones potential is less frequently used in applications for the modeling of solid substances. It was realized early that the interactions in solid phases should not be approximated to be pair-wise additive – especially for metals.[30][31]

Nevertheless, the Lennard-Jones potential is still frequently used in solid-state physics due to its simplicity and computational efficiency. Hence, the basic properties of the solid phases and the solid–fluid phase equilibria have been investigated several times, e.g. Refs.[45][36][37][58][59][48]

The Lennard-Jones substance form fcc (face centered cubic), hcp (hexagonal close-packed) and other close-packed polytype lattices – depending on temperature and pressure, cf. figure 2. At low temperature and up to moderate pressure, the hcp lattice is energetically favored and therefore the equilibrium structure. The fcc lattice structure is energetically favored at both high temperature and high pressure and therefore overall the equilibrium structure in a wider state range. The coexistence line between the fcc and hcp phase starts at   at approximately  , passes through a temperature maximum at approximately  , and then ends on the vapor–solid phase boundary at approximately  , which thereby forms a triple point.[58][36] Hence, only the fcc solid phase exhibits phase equilibria with the liquid and supercritical phase, cf. figure 2.

The triple point of the two solid phases (fcc and hcp) and the vapor phase is reported to be located at:[58][36]

  •  
  •   not reported yet
  •  
  •  
  •  

Note, that other and significantly differing values have also been reported in the literature. Hence, the database for the fcc-hcp–vapor triple point should be further solidified in the future.

 
Figure 6. Vapor–liquid equilibria of binary Lennard-Jones mixtures. In all shown cases, component 2 is the more volatile component (enriching in the vapor phase). The units are given in   and   of component 1, which is the same in all four shown mixtures. The temperature is  . Symbols are molecular simulation results and lines are results from an equation of state. Data taken from Ref.[52]

Mixtures of Lennard-Jones substances edit

Mixtures of Lennard-Jones particles are mostly used as a prototype for the development of theories and methods of solutions, but also to study properties of solutions in general. This dates back to the fundamental work of conformal solution theory of Longuet-Higgins[60] and Leland and Rowlinson and co-workers.[61][62] Those are today the basis of most theories for mixtures.[63][64]

Mixtures of two or more Lennard-Jones components are set up by changing at least one potential interaction parameter (  or  ) of one of the components with respect to the other. For a binary mixture, this yields three types of pair interactions that are all modeled by the Lennard-Jones potential: 1-1, 2-2, and 1-2 interactions. For the cross interactions 1–2, additional assumptions are required for the specification of parameters   or   from  ,   and  ,  . Various choices (all more or less empirical and not rigorously based on physical arguments) can be used for these co-called combination rules.[65] The by far most frequently used combination rule is the one of Lorentz and Berthelot[66]

 
 

The parameter   is an additional state-independent interaction parameter for the mixture. The parameter   is usually set to unity since the arithmetic mean can be considered physically plausible for the cross-interaction size parameter. The parameter   on the other hand is often used to adjust the geometric mean so as to reproduce the phase behavior of the model mixture. For analytical models, e.g. equations of state, the deviation parameter is usually written as  . For  , the cross-interaction dispersion energy and accordingly the attractive force between unlike particles is intensified, and the attractive forces between unlike particles are diminished for  .

For Lennard-Jones mixtures, both fluid and solid phase equilibria can be studied, i.e. vapor–liquid, liquid–liquid, gas–gas, solid–vapor, solid–liquid, and solid–solid. Accordingly, different types of triple points (three-phase equilibria) and critical points can exist as well as different eutectic and azeotropic points.[67][64] Binary Lennard-Jones mixtures in the fluid region (various types of equilibria of liquid and gas phases)[52][68][69][70][71] have been studied more comprehensively then phase equilibria comprising solid phases.[72][73][74][75][76] A large number of different Lennard-Jones mixtures have been studied in the literature. To date, no standard for such has been established. Usually, the binary interaction parameters and the two component parameters are chosen such that a mixture with properties convenient for a given task are obtained. Yet, this often makes comparisons tricky.

For the fluid phase behavior, mixtures exhibit practically ideal behavior (in the sense of Raoult's law) for  . For   attractive interactions prevail and the mixtures tend to form high-boiling azeotropes, i.e. a lower pressure than pure components' vapor pressures is required to stabilize the vapor–liquid equilibrium. For   repulsive interactions prevail and mixtures tend to form low-boiling azeotropes, i.e. a higher pressure than pure components' vapor pressures is required to stabilize the vapor–liquid equilibrium since the mean dispersive forces are decreased. Particularly low values of   furthermore will result in liquid–liquid miscibility gaps. Also various types of phase equilibria comprising solid phases have been studied in the literature, e.g. by Carol and co-workers.[74][76][73][72] Also, cases exist where the solid phase boundaries interrupt fluid phase equilibria. However, for phase equilibria that comprise solid phases, the amount of published data is sparse.

Equations of state for the Lennard-Jones potential edit

A large number of equations of state (EOS) for the Lennard-Jones potential/ substance have been proposed since its characterization and evaluation became available with the first computer simulations.[43] Due to the fundamental importance of the Lennard-Jones potential, most currently available molecular-based EOS are built around the Lennard-Jones fluid. They have been comprehensively reviewed by Stephan et al.[14][46]

Equations of state for the Lennard-Jones fluid are of particular importance in soft-matter physics and physical chemistry since those are frequently used as starting point for the development of EOS for complex fluids, e.g. polymers and associating fluids. The monomer units of these models are usually directly adapted from Lennard-Jones EOS as a building block, e.g. the PHC EOS,[77] the BACKONE EOS,[78][79] and SAFT type EOS.[80][81][82][83]

More than 30 Lennard-Jones EOS have been proposed in the literature. A comprehensive evaluation[14][46] of such EOS showed that several EOS[84][85][86][87] describe the Lennard-Jones potential with good and similar accuracy, but none of them is outstanding. Three of those EOS show an unacceptable unphysical behavior in some fluid region, e.g. multiple van der Waals loops, while being elsewise reasonably precise. Only the Lennard-Jones EOS of Kolafa and Nezbeda[85] was found to be robust and precise for most thermodynamic properties of the Lennard-Jones fluid.[46][14] Furthermore, the Lennard-Jones EOS of Johnson et al.[88] (today the most frequently used/ cited) was found to be less precise for practically all available reference data[9][14] than the Kolafa and Nezbeda EOS.[85]

Long-range interactions of the Lennard-Jones potential edit

 
Figure 7. Illustrative example of the convergence of a correction scheme to account for the long-range interactions of the Lennard-Jones potential. Therein,   indicates an exemplaric observable and   the applied cut-off radius. The long-range corrected value is indicated as   (symbols and line as a guide for the eye); the hypothetical 'true' value as   (dashed line).

The Lennard-Jones potential, cf. Eq. (1) and Figure 1, has an infinite range. Only under its consideration, the 'true' and 'full' Lennard-Jones potential is examined. For the evaluation of an observable of an ensemble of particles interacting by the Lennard-Jones potential using molecular simulations, the interactions can only be evaluated explicitly up to a certain distance – simply due to the fact that the number of particles will always be finite. The maximum distance applied in a simulation is usually referred to as 'cut-off' radius   (because the Lennard-Jones potential is radially symmetric). To obtain thermophysical properties (both macroscopic or microscopic) of the 'true' and 'full' Lennard-Jones (LJ) potential, the contribution of the potential beyond the cut-off radius has to be accounted for.

Different corrections schemes have been developed to account for the influence of the long-range interactions in simulations and to sustain a sufficiently good approximation of the 'full' potential.[8][35] They are based on simplifying assumptions regarding the structure of the fluid. For simple cases, such as in studies of the equilibrium of homogeneous fluids, simple correction terms yield excellent results. In other cases, such as in studies of inhomogeneous systems with different phases, accounting for the long-range interactions is more tedious. These corrections are usually referred to as 'long-range corrections'. For most properties, simple analytical expressions are known and well established. For a given observable  , the 'corrected' simulation result   is then simply computed from the actually sampled value   and the long-range correction value  , e.g. for the internal energy  .[35] The hypothetical true value of the observable of the Lennard-Jones potential at truly infinite cut-off distance (thermodynamic limit)   can in general only be estimated.

Furthermore, the quality of the long-range correction scheme depends on the cut-off radius. The assumptions made with the correction schemes are usually not justified at (very) short cut-off radii. This is illustrated in the example shown in figure 7. The long-range correction scheme is said to be converged, if the remaining error of the correction scheme is sufficiently small at a given cut-off distance, cf. figure 7.

Lennard-Jones truncated & shifted (LJTS) potential edit

 
Figure 8. Comparison of the vapor–liquid equilibrium of the 'full' Lennard-Jones potential (black) and the 'Lennard-Jones truncated & shifted' potential (blue). The symbols indicate molecular simulation results;[9][89] the lines indicate results from equations of state.[14][90]

The Lennard-Jones truncated & shifted (LJTS) potential is an often used alternative to the 'full' Lennard-Jones potential (see Eq. (1)). The 'full' and the 'truncated & shifted' Lennard-Jones potential have to be kept strictly separate. They are simply two different intermolecular potentials yielding different thermophysical properties. The Lennard-Jones truncated & shifted potential is defined as

 
with
 

Hence, the LJTS potential is truncated at   and shifted by the corresponding energy value  . The latter is applied to avoid a discontinuity jump of the potential at  . For the LJTS potential, no long-range interactions beyond   are required – neither explicitly nor implicitly. The most frequently used version of the Lennard-Jones truncated & shifted potential is the one with  . Nevertheless, different   values have been used in the literature.[91][92][93][94] Each LJTS potential with a given truncation radius   has to be considered as a potential and accordingly a substance of its own.

The LJTS potential is computationally significantly cheaper than the 'full' Lennard-Jones potential, but still covers the essential physical features of matter (the presence of a critical and a triple point, soft repulsive and attractive interactions, phase equilibria etc.). Therefore, the LJTS potential is very frequently used for the testing of new algorithms, simulation methods, and new physical theories.[95][96][97]

Interestingly, for homogeneous systems, the intermolecular forces that are calculated from the LJ and the LJTS potential at a given distance are the same (since   is the same), whereas the potential energy and the pressure are affected by the shifting. Also, the properties of the LJTS substance may furthermore be affected by the chosen simulation algorithm, i.e. MD or MC sampling (this is in general not the case for the 'full' Lennard-Jones potential).

For the LJTS potential with  , the potential energy shift is approximately 1/60 of the dispersion energy at the potential well:  . The figure 8 shows the comparison of the vapor–liquid equilibrium of the 'full' Lennard-Jones potential and the 'Lennard-Jones truncated & shifted' potential. The 'full' Lennard-Jones potential results prevail a significantly higher critical temperature and pressure compared to the LJTS potential results, but the critical density is very similar.[52][42][93] The vapor pressure and the enthalpy of vaporization are influenced more strongly by the long-range interactions than the saturated densities. This is due to the fact that the potential is manipulated mainly energetically by the truncation and shifting.

Extensions and modifications of the Lennard-Jones potential edit

The Lennard-Jones potential – as an archetype for intermolecular potentials – has been used numerous times as starting point for the development of more elaborate or more generalized intermolecular potentials. Various extensions and modifications of the Lennard-Jones potential have been proposed in the literature; a more extensive list is given in the 'interatomic potential' article. The following list refers only to several example potentials that are directly related to the Lennard-Jones potential and are of both historic importance and still relevant for present research.

  • Mie potential The Mie potential is the generalized version of the Lennard-Jones potential, i.e. the exponents 12 and 6 are introduced as parameters   and  . Especially thermodynamic derivative properties, e.g. the compressibility and the speed of sound, are known to be very sensitive to the steepness of the repulsive part of the intermolecular potential, which can therefore be modeled more sophisticated by the Mie potential.[80] The first explicit formulation of the Mie potential is attributed to Eduard Grüneisen.[98][99] Hence, the Mie potential was actually proposed before the Lennard-Jones potential. The Mie potential is named after Gustav Mie.[23]
  • Buckingham potential The Buckingham potential was proposed by Richard Buckingham. The repulsive part of the Lennard-Jones potential is therein replaced by an exponential function and it incorporates an additional parameter.
  • Stockmayer potential The Stockmayer potential is named after W.H. Stockmayer.[100] The Stockmayer potential is a combination of a Lennard-Jones potential superimposed by a dipole. Hence, Stockmayer particles are not spherically symmetric, but rather have an important orientational structure.
  • Two center Lennard-Jones potential The two center Lennard-Jones potential consists of two identical Lennard-Jones interaction sites (same  ,  ,  ) that are bonded as a rigid body. It is often abbreviated as 2CLJ. Usually, the elongation (distance between the Lennard-Jones sites) is significantly smaller than the size parameter  . Hence, the two interaction sites are significantly fused.
  • Lennard-Jones truncated & splined potential The Lennard-Jones truncated & splined potential is a rarely used yet useful potential. Similar to the more popular LJTS potential, it is sturdily truncated at a certain 'end' distance   and no long-range interactions are considered beyond. Opposite to the LJTS potential, which is shifted such that the potential is continuous, the Lennard-Jones truncated & splined potential is made continuous by using an arbitrary but favorable spline function.

See also edit

References edit

  1. ^ Fischer, Johann; Wendland, Martin (October 2023). "On the history of key empirical intermolecular potentials". Fluid Phase Equilibria. 573: 113876. doi:10.1016/j.fluid.2023.113876. ISSN 0378-3812.
  2. ^ a b c d e Lenhard, Johannes; Stephan, Simon; Hasse, Hans (February 2024). "A child of prediction. On the History, Ontology, and Computation of the Lennard-Jonesium". Studies in History and Philosophy of Science. 103: 105–113. doi:10.1016/j.shpsa.2023.11.007. PMID 38128443. S2CID 266440296.
  3. ^ Jones, J. E. (1924). "On the determination of molecular fields.—I. From the variation of the viscosity of a gas with temperature". Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character. 106 (738): 441–462. Bibcode:1924RSPSA.106..441J. doi:10.1098/rspa.1924.0081. ISSN 0950-1207.
  4. ^ Jones, J. E. (1924). "On the determination of molecular fields. —II. From the equation of state of a gas". Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character. 106 (738): 463–477. Bibcode:1924RSPSA.106..463J. doi:10.1098/rspa.1924.0082. ISSN 0950-1207.
  5. ^ Lennard-Jones, J E (1931-09-01). "Cohesion". Proceedings of the Physical Society. 43 (5): 461–482. Bibcode:1931PPS....43..461L. doi:10.1088/0959-5309/43/5/301. ISSN 0959-5309.
  6. ^ a b c Hill, Terrell L. (1956). Statistical mechanics: principles and selected applications. New York: Dover Publications. ISBN 0-486-65390-0. OCLC 15163657.
  7. ^ a b c D. C. Rapaport (1 April 2004). The Art of Molecular Dynamics Simulation. Cambridge University Press. ISBN 978-0-521-82568-9.
  8. ^ a b Frenkel, D.; Smit, B. (2002), Understanding Molecular Simulation (Second ed.), San Diego: Academic Press, ISBN 0-12-267351-4
  9. ^ a b c d e f g h i j k l m n o p q r s t u Stephan, Simon; Thol, Monika; Vrabec, Jadran; Hasse, Hans (2019-10-28). "Thermophysical Properties of the Lennard-Jones Fluid: Database and Data Assessment". Journal of Chemical Information and Modeling. 59 (10): 4248–4265. doi:10.1021/acs.jcim.9b00620. ISSN 1549-9596. PMID 31609113. S2CID 204545481.
  10. ^ a b Wood, W. W.; Parker, F. R. (1957). "Monte Carlo Equation of State of Molecules Interacting with the Lennard-Jones Potential. I. A Supercritical Isotherm at about Twice the Critical Temperature". The Journal of Chemical Physics. 27 (3): 720–733. Bibcode:1957JChPh..27..720W. doi:10.1063/1.1743822. ISSN 0021-9606.
  11. ^ a b Rosenbluth, Marshall N.; Rosenbluth, Arianna W. (1954). "Further Results on Monte Carlo Equations of State". The Journal of Chemical Physics. 22 (5): 881–884. Bibcode:1954JChPh..22..881R. doi:10.1063/1.1740207. ISSN 0021-9606.
  12. ^ Alder, B. J.; Wainwright, T. E. (1959). "Studies in Molecular Dynamics. I. General Method". The Journal of Chemical Physics. 31 (2): 459–466. Bibcode:1959JChPh..31..459A. doi:10.1063/1.1730376. ISSN 0021-9606.
  13. ^ Rahman, A. (1964-10-19). "Correlations in the Motion of Atoms in Liquid Argon". Physical Review. 136 (2A): A405–A411. Bibcode:1964PhRv..136..405R. doi:10.1103/PhysRev.136.A405. ISSN 0031-899X.
  14. ^ a b c d e f g h i Stephan, Simon; Staubach, Jens; Hasse, Hans (2020). "Review and comparison of equations of state for the Lennard-Jones fluid". Fluid Phase Equilibria. 523: 112772. doi:10.1016/j.fluid.2020.112772. S2CID 224844789.
  15. ^ 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.
  16. ^ 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. S2CID 18734898.
  17. ^ a b c Stephan, Simon; Horsch, Martin T.; Vrabec, Jadran; Hasse, Hans (2019-07-03). "MolMod – an open access database of force fields for molecular simulations of fluids". Molecular Simulation. 45 (10): 806–814. arXiv:1904.05206. doi:10.1080/08927022.2019.1601191. ISSN 0892-7022. S2CID 119199372.
  18. ^ a b c Eggimann, Becky L.; Sunnarborg, Amara J.; Stern, Hudson D.; Bliss, Andrew P.; Siepmann, J. Ilja (2014-01-02). "An online parameter and property database for the TraPPE force field". Molecular Simulation. 40 (1–3): 101–105. doi:10.1080/08927022.2013.842994. ISSN 0892-7022. S2CID 95716947.
  19. ^ Zhen, Shu; Davies, G. J. (16 August 1983). "Calculation of the Lennard-Jones nm potential energy parameters for metals". Physica Status Solidi A. 78 (2): 595–605. Bibcode:1983PSSAR..78..595Z. doi:10.1002/pssa.2210780226.
  20. ^ Eisenschitz, R.; London, F. (1930-07-01). "Über das Verhältnis der van der Waalsschen Kräfte zu den homöopolaren Bindungskräften". Zeitschrift für Physik (in German). 60 (7): 491–527. Bibcode:1930ZPhy...60..491E. doi:10.1007/BF01341258. ISSN 0044-3328. S2CID 125644826.
  21. ^ Rowlinson, J. S. (2006-11-20). "The evolution of some statistical mechanical ideas". Molecular Physics. 104 (22–24): 3399–3410. Bibcode:2006MolPh.104.3399R. doi:10.1080/00268970600965835. ISSN 0026-8976. S2CID 119942778.
  22. ^ Abascal, J. L. F.; Vega, C. (2005-12-15). "A general purpose model for the condensed phases of water: TIP4P/2005". The Journal of Chemical Physics. 123 (23): 234505. Bibcode:2005JChPh.123w4505A. doi:10.1063/1.2121687. ISSN 0021-9606. PMID 16392929.
  23. ^ a b Mie, Gustav (1903). "Zur kinetischen Theorie der einatomigen Körper". Annalen der Physik (in German). 316 (8): 657–697. Bibcode:1903AnP...316..657M. doi:10.1002/andp.19033160802.
  24. ^ Tang, K. T.; Toennies, J. Peter (1984-04-15). "An improved simple model for the van der Waals potential based on universal damping functions for the dispersion coefficients". The Journal of Chemical Physics. 80 (8): 3726–3741. Bibcode:1984JChPh..80.3726T. doi:10.1063/1.447150. ISSN 0021-9606.
  25. ^ Lafitte, Thomas; Apostolakou, Anastasia; Avendaño, Carlos; Galindo, Amparo; Adjiman, Claire S.; Müller, Erich A.; Jackson, George (2013-10-21). "Accurate statistical associating fluid theory for chain molecules formed from Mie segments". The Journal of Chemical Physics. 139 (15). Bibcode:2013JChPh.139o4504L. doi:10.1063/1.4819786. hdl:10044/1/12859. ISSN 0021-9606. PMID 24160524.
  26. ^ Stephan, Simon; Staubach, Jens; Hasse, Hans (November 2020). "Review and comparison of equations of state for the Lennard-Jones fluid". Fluid Phase Equilibria. 523: 112772. doi:10.1016/j.fluid.2020.112772. S2CID 224844789.
  27. ^ Smit, B.; Frenkel, D. (1991-04-15). "Vapor–liquid equilibria of the two-dimensional Lennard-Jones fluid(s)". The Journal of Chemical Physics. 94 (8): 5663–5668. Bibcode:1991JChPh..94.5663S. doi:10.1063/1.460477. ISSN 0021-9606. S2CID 1580499.
  28. ^ Scalise, Osvaldo H (June 2001). "Type I gas–liquid equilibria of a two-dimensional Lennard–Jones binary mixture". Fluid Phase Equilibria. 182 (1–2): 59–64. doi:10.1016/s0378-3812(01)00380-6. ISSN 0378-3812.
  29. ^ Hloucha, M.; Sandler, S. I. (November 1999). "Phase diagram of the four-dimensional Lennard-Jones fluid". The Journal of Chemical Physics. 111 (17): 8043–8047. Bibcode:1999JChPh.111.8043H. doi:10.1063/1.480138. ISSN 0021-9606.
  30. ^ a b Zhen, Shu; Davies, G. J. (1983-08-16). "Calculation of the Lennard-Jonesn–m potential energy parameters for metals". Physica Status Solidi A (in German). 78 (2): 595–605. Bibcode:1983PSSAR..78..595Z. doi:10.1002/pssa.2210780226.
  31. ^ a b Halicioglu, T.; Pound, G. M. (1975-08-16). "Calculation of potential energy parameters form crystalline state properties". Physica Status Solidi A. 30 (2): 619–623. Bibcode:1975PSSAR..30..619H. doi:10.1002/pssa.2210300223.
  32. ^ 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.
  33. ^ Mendelev, M. I.; Han, S.; Srolovitz, D. J.; Ackland, G. J.; Sun, D. Y.; Asta, M. (2003). "Development of new interatomic potentials appropriate for crystalline and liquid iron". Philosophical Magazine. 83 (35): 3977–3994. Bibcode:2003PMag...83.3977A. doi:10.1080/14786430310001613264. ISSN 1478-6435. S2CID 4119718.
  34. ^ 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.
  35. ^ a b c Allen, Michael P.; Tildesley, Dominic J. (2017-11-23). "Computer Simulation of Liquids". Oxford Scholarship Online. doi:10.1093/oso/9780198803195.001.0001. ISBN 9780198803195.
  36. ^ a b c d e f g h Schultz, Andrew J.; Kofke, David A. (2018-11-28). "Comprehensive high-precision high-accuracy equation of state and coexistence properties for classical Lennard-Jones crystals and low-temperature fluid phases". The Journal of Chemical Physics. 149 (20): 204508. Bibcode:2018JChPh.149t4508S. doi:10.1063/1.5053714. ISSN 0021-9606. PMID 30501268. S2CID 54629914.
  37. ^ a b Schultz, Andrew J.; Kofke, David A. (2020-08-07). "Erratum: "Comprehensive high-precision high-accuracy equation of state and coexistence properties for classical Lennard-Jones crystals and low-temperature fluid phases" [J. Chem. Phys. 149, 204508 (2018)]". The Journal of Chemical Physics. 153 (5): 059901. doi:10.1063/5.0021283. ISSN 0021-9606. PMID 32770918.
  38. ^ a b Schappals, Michael; Mecklenfeld, Andreas; Kröger, Leif; Botan, Vitalie; Köster, Andreas; Stephan, Simon; García, Edder J.; Rutkai, Gabor; Raabe, Gabriele; Klein, Peter; Leonhard, Kai (2017-09-12). "Round Robin Study: Molecular Simulation of Thermodynamic Properties from Models with Internal Degrees of Freedom". Journal of Chemical Theory and Computation. 13 (9): 4270–4280. doi:10.1021/acs.jctc.7b00489. ISSN 1549-9618. PMID 28738147.
  39. ^ Loeffler, Hannes H.; Bosisio, Stefano; Duarte Ramos Matos, Guilherme; Suh, Donghyuk; Roux, Benoit; Mobley, David L.; Michel, Julien (2018-11-13). "Reproducibility of Free Energy Calculations across Different Molecular Simulation Software Packages". Journal of Chemical Theory and Computation. 14 (11): 5567–5582. doi:10.1021/acs.jctc.8b00544. hdl:20.500.11820/52d85d71-d3df-468b-8f88-9c52e83da1f1. ISSN 1549-9618. PMID 30289712. S2CID 52923832.
  40. ^ a b Lenhard, Johannes; Küster, Uwe (2019). "Reproducibility and the Concept of Numerical Solution". Minds and Machines. 29 (1): 19–36. doi:10.1007/s11023-019-09492-9. ISSN 0924-6495. S2CID 59159685.
  41. ^ Shi, Wei; Johnson, J. Karl (2001-09-15). "Histogram reweighting and finite-size scaling study of the Lennard–Jones fluids". Fluid Phase Equilibria. 187–188: 171–191. doi:10.1016/S0378-3812(01)00534-9. ISSN 0378-3812.
  42. ^ a b 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
  43. ^ a b Metropolis, Nicholas; Rosenbluth, Arianna W.; Rosenbluth, Marshall N.; Teller, Augusta H.; Teller, Edward (1953). "Equation of State Calculations by Fast Computing Machines". The Journal of Chemical Physics. 21 (6): 1087–1092. Bibcode:1953JChPh..21.1087M. doi:10.1063/1.1699114. ISSN 0021-9606. OSTI 4390578. S2CID 1046577.
  44. ^ Stephan, Simon; Thol, Monika; Vrabec, Jadran; Hasse, Hans (2019-10-28). "Thermophysical Properties of the Lennard-Jones Fluid: Database and Data Assessment". Journal of Chemical Information and Modeling. 59 (10): 4248–4265. doi:10.1021/acs.jcim.9b00620. ISSN 1549-9596. PMID 31609113. S2CID 204545481.
  45. ^ a b Köster, Andreas; Mausbach, Peter; Vrabec, Jadran (2017-10-10). "Premelting, solid–fluid equilibria, and thermodynamic properties in the high density region based on the Lennard-Jones potential". The Journal of Chemical Physics. 147 (14): 144502. Bibcode:2017JChPh.147n4502K. doi:10.1063/1.4990667. ISSN 0021-9606. PMID 29031254.
  46. ^ a b c d e f g h i Stephan, Simon; Deiters, Ulrich K. (2020-08-20). "Characteristic Curves of the Lennard-Jones Fluid". International Journal of Thermophysics. 41 (10): 147. Bibcode:2020IJT....41..147S. doi:10.1007/s10765-020-02721-9. ISSN 1572-9567. PMC 7441092. PMID 32863513.
  47. ^ a b Deiters, Ulrich K.; Neumaier, Arnold (2016-08-11). "Computer Simulation of the Characteristic Curves of Pure Fluids". Journal of Chemical & Engineering Data. 61 (8): 2720–2728. doi:10.1021/acs.jced.6b00133. ISSN 0021-9568.
  48. ^ a b Agrawal, Rupal; Kofke, David A. (1995). "Thermodynamic and structural properties of model systems at solid–fluid coexistence: II. Melting and sublimation of the Lennard-Jones system". Molecular Physics. 85 (1): 43–59. doi:10.1080/00268979500100921. ISSN 0026-8976.
  49. ^ Brown, E.H. (1960). "On the thermodynamic properties of fluids". Bulletin de l'Institut International du Froid. Annexe 1960-1: 169–178.
  50. ^ Apfelbaum, E. M.; Vorob’ev, V. S. (2020-06-18). "The Line of the Unit Compressibility Factor (Zeno-Line) for Crystal States". The Journal of Physical Chemistry B. 124 (24): 5021–5027. doi:10.1021/acs.jpcb.0c02749. ISSN 1520-6106. PMID 32437611. S2CID 218835048.
  51. ^ Apfelbaum, E. M.; Vorob’ev, V. S.; Martynov, G. A. (2008). "Regarding the Theory of the Zeno Line". The Journal of Physical Chemistry A. 112 (26): 6042–6044. Bibcode:2008JPCA..112.6042A. doi:10.1021/jp802999z. ISSN 1089-5639. PMID 18543889.
  52. ^ a b c d e Stephan, Simon; Hasse, Hans (2020-06-01). "Influence of dispersive long-range interactions on properties of vapour–liquid equilibria and interfaces of binary Lennard-Jones mixtures". Molecular Physics. 118 (9–10): e1699185. Bibcode:2020MolPh.11899185S. doi:10.1080/00268976.2019.1699185. ISSN 0026-8976. S2CID 214174102.
  53. ^ Nicolas, J.J.; Gubbins, K.E.; Streett, W.B.; Tildesley, D.J. (1979). "Equation of state for the Lennard-Jones fluid". Molecular Physics. 37 (5): 1429–1454. Bibcode:1979MolPh..37.1429N. doi:10.1080/00268977900101051. ISSN 0026-8976.
  54. ^ Feng, Chao; Schultz, Andrew J.; Chaudhary, Vipin; Kofke, David A. (2015-07-28). "Eighth to sixteenth virial coefficients of the Lennard-Jones model". The Journal of Chemical Physics. 143 (4): 044504. Bibcode:2015JChPh.143d4504F. doi:10.1063/1.4927339. ISSN 0021-9606. PMID 26233142.
  55. ^ Schultz, Andrew J.; Kofke, David A. (2009-11-10). "Sixth, seventh and eighth virial coefficients of the Lennard-Jones model". Molecular Physics. 107 (21): 2309–2318. Bibcode:2009MolPh.107.2309S. doi:10.1080/00268970903267053. ISSN 0026-8976. S2CID 94811614.
  56. ^ Bell, Ian H.; Messerly, Richard; Thol, Monika; Costigliola, Lorenzo; Dyre, Jeppe C. (2019-07-25). "Modified Entropy Scaling of the Transport Properties of the Lennard-Jones Fluid". The Journal of Physical Chemistry B. 123 (29): 6345–6363. doi:10.1021/acs.jpcb.9b05808. ISSN 1520-6106. PMC 7147083. PMID 31241958.
  57. ^ Lautenschlaeger, Martin P.; Hasse, Hans (2019). "Transport properties of the Lennard-Jones truncated and shifted fluid from non-equilibrium molecular dynamics simulations". Fluid Phase Equilibria. 482: 38–47. doi:10.1016/j.fluid.2018.10.019. S2CID 106113718.
  58. ^ a b c Travesset, Alex (2014-10-28). "Phase diagram of power law and Lennard-Jones systems: Crystal phases". The Journal of Chemical Physics. 141 (16): 164501. Bibcode:2014JChPh.141p4501T. doi:10.1063/1.4898371. ISSN 0021-9606. PMID 25362319.
  59. ^ Hansen, Jean-Pierre; Verlet, Loup (1969-08-05). "Phase Transitions of the Lennard-Jones System". Physical Review. 184 (1): 151–161. Bibcode:1969PhRv..184..151H. doi:10.1103/PhysRev.184.151. ISSN 0031-899X.
  60. ^ Longuet-Higgins, H.C. (1951-02-07). "The statistical thermodynamics of multicomponent systems". Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences. 205 (1081): 247–269. Bibcode:1951RSPSA.205..247L. doi:10.1098/rspa.1951.0028. ISSN 0080-4630. S2CID 202575459.
  61. ^ Leland, T. W.; Rowlinson, J. S.; Sather, G. A. (1968). "Statistical thermodynamics of mixtures of molecules of different sizes". Transactions of the Faraday Society. 64: 1447. doi:10.1039/tf9686401447. ISSN 0014-7672.
  62. ^ Mansoori, G. Ali; Leland, Thomas W. (1972). "Statistical thermodynamics of mixtures. A new version for the theory of conformal solution". Journal of the Chemical Society, Faraday Transactions 2. 68: 320. doi:10.1039/f29726800320. ISSN 0300-9238.
  63. ^ Rowlinson, J.S.; Swinton, F.L. (1982). Liquids and liquid mixtures (Third ed.). London: Butterworth.
  64. ^ a b Deiters, Ulrich K.; Kraska, Thomas (2012). High-pressure fluid phase equilibria: phenomenology and computation (1st ed.). Amsterdam: Elsevier. ISBN 978-0-444-56354-5. OCLC 787847134.
  65. ^ Schnabel, Thorsten; Vrabec, Jadran; Hasse, Hans (2007). "Unlike Lennard–Jones parameters for vapor–liquid equilibria". Journal of Molecular Liquids. 135 (1–3): 170–178. arXiv:0904.4436. doi:10.1016/j.molliq.2006.12.024. S2CID 16111477.
  66. ^ Lorentz, H. A. (1881). "Ueber die Anwendung des Satzes vom Virial in der kinetischen Theorie der Gase". Annalen der Physik (in German). 248 (1): 127–136. Bibcode:1881AnP...248..127L. doi:10.1002/andp.18812480110.
  67. ^ van Konynenburg, P.H.; Scott, R.L. (1980-12-18). "Critical lines and phase equilibria in binary van der Waals mixtures". Philosophical Transactions of the Royal Society of London. Series A, Mathematical and Physical Sciences. 298 (1442): 495–540. Bibcode:1980RSPTA.298..495K. doi:10.1098/rsta.1980.0266. ISSN 0080-4614. S2CID 122538015.
  68. ^ Potoff, Jeffrey J.; Panagiotopoulos, Athanassios Z. (1998-12-22). "Critical point and phase behavior of the pure fluid and a Lennard-Jones mixture". The Journal of Chemical Physics. 109 (24): 10914–10920. Bibcode:1998JChPh.10910914P. doi:10.1063/1.477787. ISSN 0021-9606.
  69. ^ Protsenko, Sergey P.; Baidakov, Vladimir G. (2016). "Binary Lennard-Jones mixtures with highly asymmetric interactions of the components. 1. Effect of the energy parameters on phase equilibria and properties of liquid–gas interfaces". Fluid Phase Equilibria. 429: 242–253. doi:10.1016/j.fluid.2016.09.009.
  70. ^ Protsenko, Sergey P.; Baidakov, Vladimir G.; Bryukhanov, Vasiliy M. (2016). "Binary Lennard-Jones mixtures with highly asymmetric interactions of the components. 2. Effect of the particle size on phase equilibria and properties of liquid–gas interfaces". Fluid Phase Equilibria. 430: 67–74. doi:10.1016/j.fluid.2016.09.022.
  71. ^ Stephan, Simon; Hasse, Hans (2020-01-23). "Molecular interactions at vapor–liquid interfaces: Binary mixtures of simple fluids". Physical Review E. 101 (1): 012802. Bibcode:2020PhRvE.101a2802S. doi:10.1103/PhysRevE.101.012802. ISSN 2470-0045. PMID 32069593. S2CID 211192904.
  72. ^ a b Lamm, Monica H.; Hall, Carol K. (2002). "Equilibria between solid, liquid, and vapor phases in binary Lennard–Jones mixtures". Fluid Phase Equilibria. 194–197: 197–206. doi:10.1016/S0378-3812(01)00650-1.
  73. ^ a b Lamm, Monica H.; Hall, Carol K. (2001). "Monte Carlo simulations of complete phase diagrams for binary Lennard–Jones mixtures". Fluid Phase Equilibria. 182 (1–2): 37–46. doi:10.1016/S0378-3812(01)00378-8.
  74. ^ a b Hitchcock, Monica R.; Hall, Carol K. (1999-06-15). "Solid–liquid phase equilibrium for binary Lennard-Jones mixtures". The Journal of Chemical Physics. 110 (23): 11433–11444. Bibcode:1999JChPh.11011433H. doi:10.1063/1.479084. ISSN 0021-9606.
  75. ^ Jungblut, Swetlana; Dellago, Christoph (2011-03-14). "Crystallization of a binary Lennard-Jones mixture". The Journal of Chemical Physics. 134 (10): 104501. Bibcode:2011JChPh.134j4501J. doi:10.1063/1.3556664. ISSN 0021-9606. PMID 21405169.
  76. ^ a b Lamm, Monica H.; Hall, Carol K. (2004). "Effect of pressure on the complete phase behavior of binary mixtures". AIChE Journal. 50 (1): 215–225. Bibcode:2004AIChE..50..215L. doi:10.1002/aic.10020. ISSN 0001-1541.
  77. ^ Cotterman, R. L.; Prausnitz, J. M. (1986). "Molecular thermodynamics for fluids at low and high densities. Part II: Phase equilibria for mixtures containing components with large differences in molecular size or potential energy". AIChE Journal. 32 (11): 1799–1812. Bibcode:1986AIChE..32.1799C. doi:10.1002/aic.690321105. ISSN 0001-1541. S2CID 96417239.
  78. ^ Müller, Andreas; Winkelmann, Jochen; Fischer, Johann (1996). "Backone family of equations of state: 1. Nonpolar and polar pure fluids". AIChE Journal. 42 (4): 1116–1126. Bibcode:1996AIChE..42.1116M. doi:10.1002/aic.690420423. ISSN 0001-1541.
  79. ^ Weingerl, Ulrike; Wendland, Martin; Fischer, Johann; Müller, Andreas; Winkelmann, Jochen (2001). "Backone family of equations of state: 2. Nonpolar and polar fluid mixtures". AIChE Journal. 47 (3): 705–717. Bibcode:2001AIChE..47..705W. doi:10.1002/aic.690470317.
  80. ^ a b Lafitte, Thomas; Apostolakou, Anastasia; Avendaño, Carlos; Galindo, Amparo; Adjiman, Claire S.; Müller, Erich A.; Jackson, George (2013-10-16). "Accurate statistical associating fluid theory for chain molecules formed from Mie segments". The Journal of Chemical Physics. 139 (15): 154504. Bibcode:2013JChPh.139o4504L. doi:10.1063/1.4819786. hdl:10044/1/12859. ISSN 0021-9606. PMID 24160524.
  81. ^ Blas, F.J.; Vega, L.F. (1997). "Thermodynamic behaviour of homonuclear and heteronuclear Lennard-Jones chains with association sites from simulation and theory". Molecular Physics. 92 (1): 135–150. Bibcode:1997MolPh..92..135F. doi:10.1080/002689797170707. ISSN 0026-8976.
  82. ^ Kraska, Thomas; Gubbins, Keith E. (1996). "Phase Equilibria Calculations with a Modified SAFT Equation of State. 1. Pure Alkanes, Alkanols, and Water". Industrial & Engineering Chemistry Research. 35 (12): 4727–4737. doi:10.1021/ie9602320. ISSN 0888-5885.
  83. ^ Ghonasgi, D.; Chapman, Walter G. (1994). "Prediction of the properties of model polymer solutions and blends". AIChE Journal. 40 (5): 878–887. Bibcode:1994AIChE..40..878G. doi:10.1002/aic.690400514. ISSN 0001-1541.
  84. ^ Mecke, M.; Müller, A.; Winkelmann, J.; Vrabec, J.; Fischer, J.; Span, R.; Wagner, W. (1996-03-01). "An accurate Van der Waals-type equation of state for the Lennard-Jones fluid". International Journal of Thermophysics. 17 (2): 391–404. Bibcode:1996IJT....17..391M. doi:10.1007/BF01443399. ISSN 1572-9567. S2CID 123304062.
  85. ^ a b c Kolafa, Jiří; Nezbeda, Ivo (1994). "The Lennard-Jones fluid: an accurate analytic and theoretically-based equation of state". Fluid Phase Equilibria. 100: 1–34. doi:10.1016/0378-3812(94)80001-4.
  86. ^ Thol, Monika; Rutkai, Gabor; Köster, Andreas; Lustig, Rolf; Span, Roland; Vrabec, Jadran (2016). "Equation of State for the Lennard-Jones Fluid". Journal of Physical and Chemical Reference Data. 45 (2): 023101. Bibcode:2016JPCRD..45b3101T. doi:10.1063/1.4945000. ISSN 0047-2689.
  87. ^ Gottschalk, Matthias (2019-12-01). "An EOS for the Lennard-Jones fluid: A virial expansion approach". AIP Advances. 9 (12): 125206. Bibcode:2019AIPA....9l5206G. doi:10.1063/1.5119761. ISSN 2158-3226.
  88. ^ Johnson, J. Karl; Zollweg, John A.; Gubbins, Keith E. (1993-02-20). "The Lennard-Jones equation of state revisited". Molecular Physics. 78 (3): 591–618. Bibcode:1993MolPh..78..591J. doi:10.1080/00268979300100411. ISSN 0026-8976.
  89. ^ Vrabec, Jadran; Kedia, Gaurav Kumar; Fuchs, Guido; Hasse, Hans (2006-05-10). "Comprehensive study of the vapour–liquid coexistence of the truncated and shifted Lennard–Jones fluid including planar and spherical interface properties". Molecular Physics. 104 (9): 1509–1527. Bibcode:2006MolPh.104.1509V. doi:10.1080/00268970600556774. ISSN 0026-8976. S2CID 96606562.
  90. ^ Heier, Michaela; Stephan, Simon; Liu, Jinlu; Chapman, Walter G.; Hasse, Hans; Langenbach, Kai (2018-08-18). "Equation of state for the Lennard-Jones truncated and shifted fluid with a cut-off radius of 2.5 σ based on perturbation theory and its applications to interfacial thermodynamics". Molecular Physics. 116 (15–16): 2083–2094. Bibcode:2018MolPh.116.2083H. doi:10.1080/00268976.2018.1447153. ISSN 0026-8976. S2CID 102956189.
  91. ^ Shaul, Katherine R. S.; Schultz, Andrew J.; Kofke, David A. (2010). "The effect of truncation and shift on virial coefficients of Lennard–Jones potentials". Collection of Czechoslovak Chemical Communications. 75 (4): 447–462. doi:10.1135/cccc2009113. ISSN 1212-6950.
  92. ^ Shi, Wei; Johnson, J.Karl (2001). "Histogram reweighting and finite-size scaling study of the Lennard–Jones fluids". Fluid Phase Equilibria. 187–188: 171–191. doi:10.1016/S0378-3812(01)00534-9.
  93. ^ a b Dunikov, D. O.; Malyshenko, S. P.; Zhakhovskii, V. V. (2001-10-08). "Corresponding states law and molecular dynamics simulations of the Lennard-Jones fluid". The Journal of Chemical Physics. 115 (14): 6623–6631. Bibcode:2001JChPh.115.6623D. doi:10.1063/1.1396674. ISSN 0021-9606.
  94. ^ Lívia B. Pártay, Christoph Ortner, Albert P. Bartók, Chris J. Pickard, and Gábor Csányi "Polytypism in the ground state structure of the Lennard-Jonesium", Physical Chemistry Chemical Physics 19 19369 (2017)
  95. ^ Tchipev, Nikola; Seckler, Steffen; Heinen, Matthias; Vrabec, Jadran; Gratl, Fabio; Horsch, Martin; Bernreuther, Martin; Glass, Colin W; Niethammer, Christoph; Hammer, Nicolay; Krischok, Bernd (2019). "TweTriS: Twenty trillion-atom simulation". The International Journal of High Performance Computing Applications. 33 (5): 838–854. doi:10.1177/1094342018819741. ISSN 1094-3420. S2CID 59345875.
  96. ^ Stephan, Simon; Liu, Jinlu; Langenbach, Kai; Chapman, Walter G.; Hasse, Hans (2018). "Vapor−Liquid Interface of the Lennard-Jones Truncated and Shifted Fluid: Comparison of Molecular Simulation, Density Gradient Theory, and Density Functional Theory". The Journal of Physical Chemistry C. 122 (43): 24705–24715. doi:10.1021/acs.jpcc.8b06332. ISSN 1932-7447. S2CID 105759822.
  97. ^ Kob, Walter; Andersen, Hans C. (1995-05-01). "Testing mode-coupling theory for a supercooled binary Lennard-Jones mixture I: The van Hove correlation function". Physical Review E. 51 (5): 4626–4641. arXiv:cond-mat/9501102. Bibcode:1995PhRvE..51.4626K. doi:10.1103/PhysRevE.51.4626. PMID 9963176. S2CID 17662741.
  98. ^ Grüneisen, Edward (1911). "Das Verhältnis der thermischen Ausdehnung zur spezifischen Wärme fester Elemente". Zeitschrift für Elektrochemie und angewandte physikalische Chemie. 17 (17): 737–739. doi:10.1002/bbpc.191100004. S2CID 178760389.
  99. ^ Grüneisen, E. (1912). "Theorie des festen Zustandes einatomiger Elemente". Annalen der Physik (in German). 344 (12): 257–306. Bibcode:1912AnP...344..257G. doi:10.1002/andp.19123441202.
  100. ^ Stockmayer, W. H. (1941-05-01). "Second Virial Coefficients of Polar Gases". The Journal of Chemical Physics. 9 (5): 398–402. Bibcode:1941JChPh...9..398S. doi:10.1063/1.1750922. ISSN 0021-9606.

External links edit

  • Lennard-Jones model on SklogWiki.

lennard, jones, potential, computational, chemistry, molecular, physics, physical, chemistry, also, termed, potential, potential, named, john, lennard, jones, intermolecular, pair, potential, intermolecular, potentials, probably, that, been, most, extensively,. In computational chemistry molecular physics and physical chemistry the Lennard Jones potential also termed the LJ potential or 12 6 potential named for John Lennard Jones 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 1 2 It is considered an archetype model for simple yet realistic intermolecular interactions Figure 1 Graph of the Lennard Jones potential function Intermolecular potential energy VLJ as a function of the distance of a pair of particles The potential minimum is at r rmin 21 6s displaystyle r r rm min 2 1 6 sigma The Lennard Jones potential models soft repulsive and attractive van der Waals interactions Hence the Lennard Jones potential describes electronically neutral atoms or molecules 3 4 5 The commonly used expression for the Lennard Jones potential isVLJ r 4e sr 12 sr 6 displaystyle V text LJ r 4 varepsilon left left frac sigma r right 12 left frac sigma r right 6 right where r is the distance between two interacting particles e is the depth of the potential well usually referred to as dispersion energy and s is the distance at which the particle particle potential energy V is zero often referred to as size of the particle The Lennard Jones potential has its minimum at a distance of r rmin 21 6s displaystyle r r rm min 2 1 6 sigma where the potential energy has the value V e displaystyle V varepsilon The Lennard Jones potential is a simplified model that yet describes the essential features of interactions between simple atoms and molecules Two interacting particles repel each other at very close distance attract each other at moderate distance and do not interact at infinite distance as shown in Figure 1 The Lennard Jones potential is a pair potential i e no three or multi body interactions are covered by the potential Statistical mechanics 6 and computer simulations 7 8 can be used to study the Lennard Jones potential and to obtain thermophysical properties of the Lennard Jones substance The Lennard Jones substance is often referred to as Lennard Jonesium 2 suggesting that it is viewed as a fictive chemical element 9 Moreover its energy and length parameters can be adjusted to fit many different real substances Both the Lennard Jones potential and accordingly the Lennard Jones substance are simplified yet realistic models such as they accurately capture essential physical principles like the presence of a critical and a triple point condensation and freezing Due in part to its mathematical simplicity the Lennard Jones potential has been extensively used in studies on matter since the early days of computer simulation 10 11 12 13 The Lennard Jones potential is probably still the most frequently studied model potential 14 9 The Lennard Jones potential is usually the standard choice for the development of theories for matter especially soft matter as well as for the development and testing of computational methods and algorithms Upon adjusting the model parameters e and s to real substance properties the Lennard Jones potential can be used to describe simple substance like noble gases with good accuracy Furthermore the Lennard Jones potential is often used as a building block in molecular models a k a force fields for more complex substances 15 16 17 18 19 Contents 1 Physical background and mathematical details 2 Application of the Lennard Jones potential 3 Alternative notations of the Lennard Jones potential 4 Dimensionless reduced units 5 Thermophysical properties of the Lennard Jones substance 5 1 Characteristic points and curves 5 2 Properties of the Lennard Jones fluid 5 3 Properties of the Lennard Jones solid 6 Mixtures of Lennard Jones substances 7 Equations of state for the Lennard Jones potential 8 Long range interactions of the Lennard Jones potential 9 Lennard Jones truncated amp shifted LJTS potential 10 Extensions and modifications of the Lennard Jones potential 11 See also 12 References 13 External linksPhysical background and mathematical details editThe Lennard Jones potential models the two most important and fundamental molecular interactions The repulsive term 1 r12 displaystyle 1 r 12 nbsp term describes the Pauli repulsion at short distances of the interacting particles due to overlapping electron orbitals and the attractive term 1 r6 displaystyle 1 r 6 nbsp term describes attraction at long ranged interactions London dispersion force which vanish at infinite distance between two particles The steep repulsive interactions at short distances yield the low compressibility of the solid and liquid phase the attractive dispersive interactions act stabilizing for the condensed phase especially the vapor liquid equilibrium The functional form of the attractive term the exponent 6 has a physical justification which does not hold as rigorously for the repulsive term with the exponent 12 The attractive dispersive interactions between simple atoms and molecules are a result of fluctuating partial charges It has been shown by quantum chemical calculations that this dispersive contribution has to decay with 1 r6 displaystyle 1 r 6 nbsp 20 2 The 1 r12 displaystyle 1 r 12 nbsp term is mainly used because it can be implemented computationally very efficiently as the square of 1 r6 displaystyle 1 r 6 nbsp which does not hold to the same extent for values other than 12 Also 1 r12 displaystyle 1 r 12 nbsp approximates the Pauli repulsion reasonably well If needed the Lennard Jones potential can be generalized using arbitrary exponents instead of 12 and 6 the resulting model is called the Mie potential The present article exclusively discusses the original 12 6 Lennard Jones potential The Lennard Jones potential exhibits a pole at r 0 displaystyle r rightarrow 0 nbsp i e the potential energy diverges to V displaystyle V rightarrow infty nbsp which can cause instabilities in molecular simulations e g for the sampling of the chemical potential The Lennard Jones potential converges to V 0 displaystyle V rightarrow 0 nbsp for r displaystyle r rightarrow infty nbsp Hence from a mathematical standpoint attractive interactions stay present for infinitely distanced particles These dispersive long range interactions have an important influence on several properties of the Lennard Jones substance e g the pressure or heat capacity in the vicinity of the critical point and the critical point itself The importance of the long range interactions were noticed already in the early stages of statistical mechanics 21 For computer simulations only finite numbers of particles can be used which leads to the fact that the potential can only be evaluated up to a finite radius r displaystyle r nbsp which is a so called finite size effect There are well established methods to implicitly consider the thereby neglected long range contribution for a given observable details are given below It is often claimed that multiple Lennard Jones potentials and corresponding substances exist depending on the handling of the long range interactions This is misleading There is only one Lennard Jones potential which is exactly defined by Eq 1 The Lennard Jones potential requires the consideration and evaluation of long range interactions up to very long actually infinite distances at least so that the influence of the truncation has no influence on the observable of interest for the reported decimal places The Lennard Jones potential implies that the particles are point masses with a mass m displaystyle m nbsp Even though the parameter s displaystyle sigma nbsp is often referred to as size of the particle particles interacting with the Lennard Jones potential have no uniquely defined size opposite to the hard sphere potential Particles interacting with the Lennard Jones potential rather have soft repulsive cores The Lennard Jones model describes the potential intermolecular energy V displaystyle V nbsp between two particles based on the outlined principles Following Newton s mechanics the actual force F displaystyle F nbsp between two interacting particles is simply obtained by negating and differentiating the Lennard Jones potential with respect to r displaystyle r nbsp i e F dV dr displaystyle F mathrm d V mathrm d r nbsp Depending on the distance between the two particles the net force can be either attractive or repulsive The Lennard Jones potential yields a good approximation of intermolecular interactions for many applications The macroscopic properties computed using the Lennard Jones potential are in good agreement with experimental data for simple substances like argon on one side and the potential function VLJ r displaystyle V mathrm LJ r nbsp is in fair agreement with results from quantum chemistry on the other side The Lennard Jones potential gives a good description of molecular interactions in fluid phases whereas molecular interactions in solid phases are only roughly well described This is mainly due to the fact that multi body interactions play a significant role in solid phases which are not comprised in the Lennard Jones potential Therefore the Lennard Jones potential is extensively used in soft matter physics and associated fields whereas it is less frequently used in solid state physics Due to its simplicity the Lennard Jones potential is often used to describe the properties of gases and simple fluids and to model dispersive and repulsive interactions in molecular models It is especially accurate for noble gas atoms and methane It is furthermore a good approximation for molecular interactions at long and short distances for neutral atoms and molecules Therefore the Lennard Jones potential is very often used as a building block of molecular models of complex molecules e g alkanes or water 18 22 17 The Lennard Jones potential can also be used to model the adsorption interactions at solid fluid interfaces i e physisorption or chemisorption It is well accepted that the main limitations of the Lennard Jones potential lie in the fact the potential is a pair potential does not cover multi body interactions and that the 1 r12 displaystyle 1 r 12 nbsp exponent term is used for the repulsion Results from quantum chemistry suggest that a higher exponent than 12 has to be used i e a steeper potential Furthermore the Lennard Jones potential has a limited flexibility i e only the two model parameters e displaystyle varepsilon nbsp and s displaystyle sigma nbsp can be used for the fitting to describe a real substance Numerous intermolecular potentials have been proposed in the past for the modeling of simple soft repulsive and attractive interactions between spherically symmetric particles i e the general shape shown in Figure 1 Examples for other potentials are the Morse potential the Mie potential 23 the Buckingham potential and the Tang Tonnies potential 24 While some of these may be more suited to modelling real fluids 25 the simplicity of the Lennard Jones potential as well as it s often surprising ability to accurately capture real fluid behaviour has historically made it the pair potential of greatest general importance 26 Application of the Lennard Jones potential editThe Lennard Jones potential is not only of fundamental importance in computational chemistry and soft matter physics but also for the modeling of real substances The Lennard Jones potential is frequently used for fundamental studies on the behavior of matter and for elucidating atomistic phenomena It is also often used for somewhat special use cases e g for studying thermophysical properties of two or four dimensional substances 27 28 29 instead of the classical three spatial directions of our universe The Lennard Jones potential is extensively used for molecular modeling There are essentially two ways the Lennard Jones potential can be used for molecular modeling 1 A real substance atom or molecule is modeled directly by the Lennard Jones potential which yields very good results for noble gases and methane i e dispersively interacting spherical particles In the case of methane the molecule is assumed to be spherically symmetric and the hydrogen atoms are fused with the carbon atom to a common unit This simplification can in general also be applied to more complex molecules but yields usually poor results 2 A real substance molecule is built of multiple Lennard Jones interactions sites which can be connected either by rigid bonds or flexible additional potentials and eventually also consists of other potential types e g partial charges Molecular models often referred to as force fields for practically all molecular and ionic particles can be constructed using this scheme for example for alkanes Upon using the first outlined approach the molecular model has only the two parameters of the Lennard Jones potential e displaystyle varepsilon nbsp and s displaystyle sigma nbsp that can be used for the fitting e g e kB 120K displaystyle varepsilon k mathrm B 120 mathrm K nbsp and s 0 34nm displaystyle sigma 0 34 mathrm nm nbsp are frequently used for argon Evidently this approach is only a good approximation for spherical and simply dispersively interacting molecules and atoms The direct use of the Lennard Jones potential has the great advantage that simulation results and theories for the Lennard Jones potential can be used directly Hence available results for the Lennard Jones potential and substance can be directly scaled using the appropriate e displaystyle varepsilon nbsp and s displaystyle sigma nbsp see reduced units The Lennard Jones potential parameters e displaystyle varepsilon nbsp and s displaystyle sigma nbsp can in general be fitted to any desired real substance property In soft matter physics usually experimental data for the vapor liquid phase equilibrium or the critical point are used for the parametrization in solid state physics rather the compressibility heat capacity or lattice constants are employed 30 31 The second outlined approach of using the Lennard Jones potential as a building block of elongated and complex molecules is far more sophisticated Molecular models are thereby tailor made in a sense that simulation results are only applicable for that particular model This development approach for molecular force fields is today mainly performed in soft matter physics and associated fields such as chemical engineering chemistry and computational biology A large number of force fields are based on the Lennard Jones potential e g the TraPPE force field 18 the OPLS force field 32 and the MolMod force field 17 an overview of molecular force fields is out of the scope of the present article For the state of the art modeling of solid state materials more elaborate multi body potentials e g EAM potentials 33 are used Alternative notations of the Lennard Jones potential editThere are several different ways to formulate the Lennard Jones potential besides Eq 1 Alternatives are AB formThe AB form is frequently used in implementations of simulation software as it is computationally favorable The Lennard Jones potential can be written asVLJ r Ar12 Br6 displaystyle V text LJ r frac A r 12 frac B r 6 nbsp where A 4es12 displaystyle A 4 varepsilon sigma 12 nbsp and B 4es6 displaystyle B 4 varepsilon sigma 6 nbsp Conversely s AB6 displaystyle sigma sqrt 6 frac A B nbsp and e B24A displaystyle varepsilon frac B 2 4A nbsp This is the form in which Lennard Jones originally presented the potential named after him 34 n exp formThe n exp form is a mathematically more general form see Mie potential and can be written asVLJ r 4e r0r 2n r0r n displaystyle V text LJ r 4 varepsilon left left frac r 0 r right 2n left frac r 0 r right n right nbsp where e displaystyle varepsilon nbsp is the bonding energy of the molecule the energy required to separate the atoms Applying a harmonic approximation to the potential minimum at V rm e displaystyle V r m varepsilon nbsp the exponent n displaystyle n nbsp and the energy parameter e displaystyle varepsilon nbsp can be related to the harmonic spring constant k 2e nr0 2 displaystyle k 2 varepsilon left frac n r 0 right 2 nbsp from which n displaystyle n nbsp can be calculated if k displaystyle k nbsp is known If changes in the harmonic states E displaystyle E nbsp are known from experiment for example from Raman spectroscopy then DE ℏw displaystyle Delta E hbar omega nbsp where w k m textstyle omega sqrt k mu nbsp and m m 2 displaystyle mu m 2 nbsp is the reduced mass and m displaystyle m nbsp is the particle mass may be used to estimate the spring constant Alternatively n displaystyle n nbsp can be related to vg displaystyle v text g nbsp the group velocity in a crystal usingvg a nr0em displaystyle v text g frac a cdot n r 0 sqrt frac varepsilon m nbsp where a displaystyle a nbsp is the lattice distance citation needed Dimensionless reduced units editdimensionless reduced units Property Symbol Reduced formLength r displaystyle r nbsp rs displaystyle frac r sigma nbsp Time t displaystyle t nbsp tems2 displaystyle t sqrt frac varepsilon m sigma 2 nbsp Temperature T displaystyle T nbsp kBTe displaystyle frac k B T varepsilon nbsp Force F displaystyle F nbsp Fse displaystyle frac F sigma varepsilon nbsp Energy U displaystyle U nbsp Ue displaystyle frac U varepsilon nbsp Pressure p displaystyle p nbsp ps3e displaystyle frac p sigma 3 varepsilon nbsp Density r displaystyle rho nbsp rs3 displaystyle rho sigma 3 nbsp Surface tension g displaystyle gamma nbsp gs2e displaystyle frac gamma sigma 2 varepsilon nbsp Dimensionless reduced units can be defined based on the Lennard Jones potential parameters which is convenient for molecular simulations From a numerical point of view the advantages of this unit system include computing values which are closer to unity using simplified equations and being able to easily scale the results 35 7 This reduced units system requires the specification of the size parameter s displaystyle sigma nbsp and the energy parameter e displaystyle varepsilon nbsp of the Lennard Jones potential and the mass of the particle m displaystyle m nbsp All physical properties can be converted straightforwardly taking the respective dimension into account see table The reduced units are often abbreviated and indicated by an asterisk In general reduced units can also be built up on other molecular interaction potentials that consist of a length parameter and an energy parameter Thermophysical properties of the Lennard Jones substance edit nbsp Figure 2 Phase diagram of the Lennard Jones substance Correlations and numeric values for the critical point and triple point s are taken from Refs 9 36 14 The star indicates the critical point 9 The circle indicates the vapor liquid solid triple point and the triangle indicates the vapor solid fcc solid hcp triple point 36 37 The solid lines indicate coexistence lines of two phases 9 36 The dashed lines indicate the vapor liquid spinodal 14 Thermophysical properties of the Lennard Jones substance 2 i e particles interacting with the Lennard Jones potential can be obtained using statistical mechanics Some properties can be computed analytically i e with machine precision whereas most properties can only be obtained by performing molecular simulations 7 The latter will in general be superimposed by both statistical and systematic uncertainties 38 9 39 40 The virial coefficients can for example be computed directly from the Lennard potential using algebraic expressions 6 and reported data has therefore no uncertainty Molecular simulation results e g the pressure at a given temperature and density has both statistical and systematic uncertainties 38 40 Molecular simulations of the Lennard Jones potential can in general be performed using either molecular dynamics MD simulations or Monte Carlo MC simulation For MC simulations the Lennard Jones potential VLJ r displaystyle V mathrm LJ r nbsp is directly used whereas MD simulations are always based on the derivative of the potential i e the force F dV dr displaystyle F mathrm d V mathrm d r nbsp These differences in combination with differences in the treatment of the long range interactions see below can influence computed thermophysical properties 41 42 Since the Lennard Jonesium is the archetype for the modeling of simple yet realistic intermolecular interactions a large number of thermophysical properties were studied and reported in the literature 9 Computer experiment data of the Lennard Jones potential is presently considered the most accurately known data in classical mechanics computational chemistry Hence such data is also mostly used as benchmark for the validation and testing of new algorithms and theories The Lennard Jones potential has been constantly used since the early days of molecular simulations The first results from computer experiments for the Lennard Jones potential were reported by Rosenbluth and Rosenbluth 11 and Wood and Parker 10 after molecular simulations on fast computing machines became available in 1953 43 Since then many studies reported data of the Lennard Jones substance 9 approximately 50 000 data points are publicly available The current state of research of thermophysical properties of the Lennard Jones substance is summarized in the following The most comprehensive summary and digital database was given by Stephan et al 9 Presently no data repository covers and maintains this database or any other model potential the concise data selection stated by the NIST website should be treated with caution regarding referencing 44 and coverage it contains a small fraction of the available data Most of the data on NIST website provides non peer reviewed data generated in house by NIST Figure 2 shows the phase diagram of the Lennard Jones fluid Phase equilibria of the Lennard Jones potential have been studied numerous times and are accordingly known today with good precision 36 9 45 Figure 2 shows results correlations derived from computer experiment results hence lines instead of data points are shown The mean intermolecular interaction of a Lennard Jones particle strongly depends on the thermodynamic state i e temperature and pressure or density For solid states the attractive Lennard Jones interaction plays a dominant role especially at low temperatures For liquid states no ordered structure is present compared to solid states The mean potential energy per particle is negative For gaseous states attractive interactions of the Lennard Jones potential play a minor role since they are far distanced The main part of the internal energy is stored as kinetic energy for gaseous states At supercritical states the attractive Lennard Jones interaction plays a minor role With increasing temperature the mean kinetic energy of the particles increases and exceeds the energy well of the Lennard Jones potential Hence the particles mainly interact by the potentials soft repulsive interactions and the mean potential energy per particle is accordingly positive Overall due to the large timespan the Lennard Jones potential has been studied and thermophysical property data has been reported in the literature and computational resources were insufficient for accurate simulations to modern standards a noticeable amount of data is known to be dubious 9 Nevertheless in many studies such data is used as reference The lack of data repositories and data assessment is a crucial element for future work in the long going field of Lennard Jones potential research Characteristic points and curves edit The most important characteristic points of the Lennard Jones potential are the critical point and the vapor liquid solid triple point They were studied numerous times in the literature and compiled in Ref 9 The critical point was thereby assessed to be located at Tc 1 321 0 007ekB 1 displaystyle T mathrm c 1 321 pm 0 007 varepsilon k mathrm B 1 nbsp rc 0 316 0 005s 3 displaystyle rho mathrm c 0 316 pm 0 005 sigma 3 nbsp pc 0 129 0 005es 3 displaystyle p mathrm c 0 129 pm 0 005 varepsilon sigma 3 nbsp The given uncertainties were calculated from the standard deviation of the critical parameters derived from the most reliable available vapor liquid equilibrium data sets 9 These uncertainties can be assumed as a lower limit to the accuracy with which the critical point of fluid can be obtained from molecular simulation results nbsp Figure 3 Characteristic curves of the Lennard Jones substance The thick black line indicates the vapor liquid equilibrium the star indicates the critical point The brown line indicates the solid fluid equilibrium Other black solid lines and symbols indicate Brown s characteristic curves see text for details of the Lennard Jones substance lines are results from an equation of state symbols from molecular simulations and triangles exact data in the ideal gas limit obtained from the virial coefficients Data taken from 46 47 48 The triple point is presently assumed to be located at Ttr 0 69 0 005ekB 1 displaystyle T mathrm tr 0 69 pm 0 005 varepsilon k mathrm B 1 nbsp rtr gas 0 0017 0 004s 3 displaystyle rho mathrm tr gas 0 0017 pm 0 004 sigma 3 nbsp rtr liq 0 845 0 009s 3 displaystyle rho mathrm tr liq 0 845 pm 0 009 sigma 3 nbsp rtr sol 0 961 0 007s 3 displaystyle rho mathrm tr sol 0 961 pm 0 007 sigma 3 nbsp ptr 0 0012 0 0007es 3 displaystyle p mathrm tr 0 0012 pm 0 0007 varepsilon sigma 3 nbsp The uncertainties represent the scattering of data from different authors 36 The critical point of the Lennard Jones substance has been studied far more often than the triple point For both the critical point and the vapor liquid solid triple point several studies reported results out of the above stated ranges The above stated data is the presently assumed correct and reliable data Nevertheless the determinateness of the critical temperature and the triple point temperature is still unsatisfactory Evidently the phase coexistence curves cf figure 2 are of fundamental importance to characterize the Lennard Jones potential Furthermore Brown s characteristic curves 49 yield an illustrative description of essential features of the Lennard Jones potential Brown s characteristic curves are defined as curves on which a certain thermodynamic property of the substance matches that of an ideal gas For a real fluid Z displaystyle Z nbsp and its derivatives can match the values of the ideal gas for special T displaystyle T nbsp r displaystyle rho nbsp combinations only as a result of Gibbs phase rule The resulting points collectively constitute a characteristic curve Four main characteristic curves are defined One 0th order named Zeno curve and three 1st order curves named Amagat Boyle and Charles curve The characteristic curve are required to have a negative or zero curvature throughout and a single maximum in a double logarithmic pressure temperature diagram Furthermore Brown s characteristic curves and the virial coefficients are directly linked in the limit of the ideal gas and are therefore known exactly at r 0 displaystyle rho rightarrow 0 nbsp Both computer simulation results and equation of state results have been reported in the literature for the Lennard Jones potential 47 9 46 50 51 Points on the Zeno curve Z have a compressibility factor of unity Z p rT 1 displaystyle Z p rho T 1 nbsp The Zeno curve originates at the Boyle temperature TB 3 417927982ekB 1 displaystyle T mathrm B 3 417927982 varepsilon k mathrm B 1 nbsp surrounds the critical point and has a slope of unity in the low temperature limit 46 Points on the Boyle curve B have dZd 1 r T 0 displaystyle left frac mathrm d Z mathrm d 1 rho right T 0 nbsp The Boyle curve originates with the Zeno curve at the Boyle temperature faintly surrounds the critical point and ends on the vapor pressure curve Points on the Charles curve a k a Joule Thomson inversion curve have dZdT p 0 displaystyle left frac mathrm d Z mathrm d T right p 0 nbsp and more importantly dTdp h 0 displaystyle left frac mathrm d T mathrm d p right h 0 nbsp i e no temperature change upon isenthalpic throttling It originates at T 6 430798418ekB 1 displaystyle T 6 430798418 varepsilon k mathrm B 1 nbsp in the ideal gas limit crosses the Zeno curve and terminates on the vapor pressure curve Points on the Amagat curve A have dZdT r 0 displaystyle left frac mathrm d Z mathrm d T right rho 0 nbsp It also starts in the ideal gas limit at T 25 15242837ekB 1 displaystyle T 25 15242837 varepsilon k mathrm B 1 nbsp surrounds the critical point and the other three characteristic curves and passes into the solid phase region A comprehensive discussion of the characteristic curves of the Lennard Jones potential is given by Stephan and Deiters 46 nbsp Figure 4 Virial coefficients from the Lennard Jones potential as a function of the temperature Second virial coefficient B displaystyle B nbsp top and third virial coefficient C displaystyle C nbsp bottom The circle indicates the Boyle temperature TB displaystyle T mathrm B nbsp Results taken from 46 Properties of the Lennard Jones fluid edit nbsp Figure 5 Vapor liquid equilibrium of the Lennard Jones substance Vapor pressure top saturated densities middle and interfacial tension bottom Symbols indicate molecular simulation results 52 9 Lines indicate results from equation of state and square gradient theory for the interfacial tension 52 14 Properties of the Lennard Jones fluid have been studied extensively in the literature due to the outstanding importance of the Lennard Jones potential in soft matter physics and related fields 2 About 50 datasets of computer experiment data for the vapor liquid equilibrium have been published to date 9 Furthermore more than 35 000 data points at homogeneous fluid states have been published over the years and recently been compiled and assessed for outliers in an open access database 9 The vapor liquid equilibrium of the Lennard Jones substance is presently known with a precision i e mutual agreement of thermodynamically consistent data of 1 displaystyle pm 1 nbsp for the vapor pressure 0 2 displaystyle pm 0 2 nbsp for the saturated liquid density 1 displaystyle pm 1 nbsp for the saturated vapor density 0 75 displaystyle pm 0 75 nbsp for the enthalpy of vaporization and 4 displaystyle pm 4 nbsp for the surface tension 9 This status quo can not be considered satisfactory considering the fact that statistical uncertainties usually reported for single data sets are significantly below the above stated values even for far more complex molecular force fields Both phase equilibrium properties and homogeneous state properties at arbitrary density can in general only be obtained from molecular simulations whereas virial coefficients can be computed directly from the Lennard Jones potential 6 Numerical data for the second and third virial coefficient is available in a wide temperature range 53 46 9 For higher virial coefficients up to the sixteenth the number of available data points decreases with increasing number of the virial coefficient 54 55 Also transport properties viscosity heat conductivity and self diffusion coefficient of the Lennard Jones fluid have been studied frequently 56 57 but the database is significantly less dense than for homogeneous equilibrium properties like pvT displaystyle pvT nbsp or internal energy data Moreover a large number of analytical models equations of state have been developed for the description of the Lennard Jones fluid see below for details Properties of the Lennard Jones solid edit The database and knowledge for the Lennard Jones solid is significantly poorer than for the fluid phases which is mainly due to the fact that the Lennard Jones potential is less frequently used in applications for the modeling of solid substances It was realized early that the interactions in solid phases should not be approximated to be pair wise additive especially for metals 30 31 Nevertheless the Lennard Jones potential is still frequently used in solid state physics due to its simplicity and computational efficiency Hence the basic properties of the solid phases and the solid fluid phase equilibria have been investigated several times e g Refs 45 36 37 58 59 48 The Lennard Jones substance form fcc face centered cubic hcp hexagonal close packed and other close packed polytype lattices depending on temperature and pressure cf figure 2 At low temperature and up to moderate pressure the hcp lattice is energetically favored and therefore the equilibrium structure The fcc lattice structure is energetically favored at both high temperature and high pressure and therefore overall the equilibrium structure in a wider state range The coexistence line between the fcc and hcp phase starts at T 0 displaystyle T 0 nbsp at approximately p 878 5es 3 displaystyle p 878 5 varepsilon sigma 3 nbsp passes through a temperature maximum at approximately T 0 4ekB 1 displaystyle T 0 4 varepsilon k mathrm B 1 nbsp and then ends on the vapor solid phase boundary at approximately T 0 32ekB 1 displaystyle T 0 32 varepsilon k mathrm B 1 nbsp which thereby forms a triple point 58 36 Hence only the fcc solid phase exhibits phase equilibria with the liquid and supercritical phase cf figure 2 The triple point of the two solid phases fcc and hcp and the vapor phase is reported to be located at 58 36 Ttr 0 32 0 001ekB 1 displaystyle T mathrm tr 0 32 pm 0 001 varepsilon k mathrm B 1 nbsp rtr gas displaystyle rho mathrm tr gas nbsp not reported yet rtr fcc 1 03859 0 0008s 3 displaystyle rho mathrm tr fcc 1 03859 pm 0 0008 sigma 3 nbsp rtr hcp 1 03861 0 0007s 3 displaystyle rho mathrm tr hcp 1 03861 pm 0 0007 sigma 3 nbsp ptr 0 96 10 9es 3 displaystyle p mathrm tr 0 96 cdot 10 9 varepsilon sigma 3 nbsp Note that other and significantly differing values have also been reported in the literature Hence the database for the fcc hcp vapor triple point should be further solidified in the future nbsp Figure 6 Vapor liquid equilibria of binary Lennard Jones mixtures In all shown cases component 2 is the more volatile component enriching in the vapor phase The units are given in e displaystyle varepsilon nbsp and s displaystyle sigma nbsp of component 1 which is the same in all four shown mixtures The temperature is T 0 92ekB 1 displaystyle T 0 92 varepsilon k mathrm B 1 nbsp Symbols are molecular simulation results and lines are results from an equation of state Data taken from Ref 52 Mixtures of Lennard Jones substances editMixtures of Lennard Jones particles are mostly used as a prototype for the development of theories and methods of solutions but also to study properties of solutions in general This dates back to the fundamental work of conformal solution theory of Longuet Higgins 60 and Leland and Rowlinson and co workers 61 62 Those are today the basis of most theories for mixtures 63 64 Mixtures of two or more Lennard Jones components are set up by changing at least one potential interaction parameter e displaystyle varepsilon nbsp or s displaystyle sigma nbsp of one of the components with respect to the other For a binary mixture this yields three types of pair interactions that are all modeled by the Lennard Jones potential 1 1 2 2 and 1 2 interactions For the cross interactions 1 2 additional assumptions are required for the specification of parameters e12 displaystyle varepsilon mathrm 12 nbsp or s12 displaystyle sigma mathrm 12 nbsp from e11 displaystyle varepsilon mathrm 11 nbsp s11 displaystyle sigma mathrm 11 nbsp and e22 displaystyle varepsilon mathrm 22 nbsp s22 displaystyle sigma mathrm 22 nbsp Various choices all more or less empirical and not rigorously based on physical arguments can be used for these co called combination rules 65 The by far most frequently used combination rule is the one of Lorentz and Berthelot 66 s12 h12s11 s222 displaystyle sigma 12 eta 12 frac sigma 11 sigma 22 2 nbsp e12 312e11e22 displaystyle varepsilon 12 xi 12 sqrt varepsilon 11 varepsilon 22 nbsp The parameter 312 displaystyle xi 12 nbsp is an additional state independent interaction parameter for the mixture The parameter h12 displaystyle eta 12 nbsp is usually set to unity since the arithmetic mean can be considered physically plausible for the cross interaction size parameter The parameter 312 displaystyle xi 12 nbsp on the other hand is often used to adjust the geometric mean so as to reproduce the phase behavior of the model mixture For analytical models e g equations of state the deviation parameter is usually written as k12 1 312 displaystyle k 12 1 xi 12 nbsp For 312 gt 1 displaystyle xi 12 gt 1 nbsp the cross interaction dispersion energy and accordingly the attractive force between unlike particles is intensified and the attractive forces between unlike particles are diminished for 312 lt 1 displaystyle xi 12 lt 1 nbsp For Lennard Jones mixtures both fluid and solid phase equilibria can be studied i e vapor liquid liquid liquid gas gas solid vapor solid liquid and solid solid Accordingly different types of triple points three phase equilibria and critical points can exist as well as different eutectic and azeotropic points 67 64 Binary Lennard Jones mixtures in the fluid region various types of equilibria of liquid and gas phases 52 68 69 70 71 have been studied more comprehensively then phase equilibria comprising solid phases 72 73 74 75 76 A large number of different Lennard Jones mixtures have been studied in the literature To date no standard for such has been established Usually the binary interaction parameters and the two component parameters are chosen such that a mixture with properties convenient for a given task are obtained Yet this often makes comparisons tricky For the fluid phase behavior mixtures exhibit practically ideal behavior in the sense of Raoult s law for 312 1 displaystyle xi 12 1 nbsp For 312 gt 1 displaystyle xi 12 gt 1 nbsp attractive interactions prevail and the mixtures tend to form high boiling azeotropes i e a lower pressure than pure components vapor pressures is required to stabilize the vapor liquid equilibrium For 312 lt 1 displaystyle xi 12 lt 1 nbsp repulsive interactions prevail and mixtures tend to form low boiling azeotropes i e a higher pressure than pure components vapor pressures is required to stabilize the vapor liquid equilibrium since the mean dispersive forces are decreased Particularly low values of 312 displaystyle xi 12 nbsp furthermore will result in liquid liquid miscibility gaps Also various types of phase equilibria comprising solid phases have been studied in the literature e g by Carol and co workers 74 76 73 72 Also cases exist where the solid phase boundaries interrupt fluid phase equilibria However for phase equilibria that comprise solid phases the amount of published data is sparse Equations of state for the Lennard Jones potential editA large number of equations of state EOS for the Lennard Jones potential substance have been proposed since its characterization and evaluation became available with the first computer simulations 43 Due to the fundamental importance of the Lennard Jones potential most currently available molecular based EOS are built around the Lennard Jones fluid They have been comprehensively reviewed by Stephan et al 14 46 Equations of state for the Lennard Jones fluid are of particular importance in soft matter physics and physical chemistry since those are frequently used as starting point for the development of EOS for complex fluids e g polymers and associating fluids The monomer units of these models are usually directly adapted from Lennard Jones EOS as a building block e g the PHC EOS 77 the BACKONE EOS 78 79 and SAFT type EOS 80 81 82 83 More than 30 Lennard Jones EOS have been proposed in the literature A comprehensive evaluation 14 46 of such EOS showed that several EOS 84 85 86 87 describe the Lennard Jones potential with good and similar accuracy but none of them is outstanding Three of those EOS show an unacceptable unphysical behavior in some fluid region e g multiple van der Waals loops while being elsewise reasonably precise Only the Lennard Jones EOS of Kolafa and Nezbeda 85 was found to be robust and precise for most thermodynamic properties of the Lennard Jones fluid 46 14 Furthermore the Lennard Jones EOS of Johnson et al 88 today the most frequently used cited was found to be less precise for practically all available reference data 9 14 than the Kolafa and Nezbeda EOS 85 Long range interactions of the Lennard Jones potential edit nbsp Figure 7 Illustrative example of the convergence of a correction scheme to account for the long range interactions of the Lennard Jones potential Therein X displaystyle X nbsp indicates an exemplaric observable and rc displaystyle r mathrm c nbsp the applied cut off radius The long range corrected value is indicated as Xcorr displaystyle X mathrm corr nbsp symbols and line as a guide for the eye the hypothetical true value as Xtrue displaystyle X mathrm true nbsp dashed line The Lennard Jones potential cf Eq 1 and Figure 1 has an infinite range Only under its consideration the true and full Lennard Jones potential is examined For the evaluation of an observable of an ensemble of particles interacting by the Lennard Jones potential using molecular simulations the interactions can only be evaluated explicitly up to a certain distance simply due to the fact that the number of particles will always be finite The maximum distance applied in a simulation is usually referred to as cut off radius rc displaystyle r mathrm c nbsp because the Lennard Jones potential is radially symmetric To obtain thermophysical properties both macroscopic or microscopic of the true and full Lennard Jones LJ potential the contribution of the potential beyond the cut off radius has to be accounted for Different corrections schemes have been developed to account for the influence of the long range interactions in simulations and to sustain a sufficiently good approximation of the full potential 8 35 They are based on simplifying assumptions regarding the structure of the fluid For simple cases such as in studies of the equilibrium of homogeneous fluids simple correction terms yield excellent results In other cases such as in studies of inhomogeneous systems with different phases accounting for the long range interactions is more tedious These corrections are usually referred to as long range corrections For most properties simple analytical expressions are known and well established For a given observable X displaystyle X nbsp the corrected simulation result Xcorr displaystyle X mathrm corr nbsp is then simply computed from the actually sampled value Xsampled displaystyle X mathrm sampled nbsp and the long range correction value Xlrc displaystyle X mathrm lrc nbsp e g for the internal energy Ucorr Usampled Ulrc displaystyle U mathrm corr U mathrm sampled U mathrm lrc nbsp 35 The hypothetical true value of the observable of the Lennard Jones potential at truly infinite cut off distance thermodynamic limit Xtrue displaystyle X mathrm true nbsp can in general only be estimated Furthermore the quality of the long range correction scheme depends on the cut off radius The assumptions made with the correction schemes are usually not justified at very short cut off radii This is illustrated in the example shown in figure 7 The long range correction scheme is said to be converged if the remaining error of the correction scheme is sufficiently small at a given cut off distance cf figure 7 Lennard Jones truncated amp shifted LJTS potential edit nbsp Figure 8 Comparison of the vapor liquid equilibrium of the full Lennard Jones potential black and the Lennard Jones truncated amp shifted potential blue The symbols indicate molecular simulation results 9 89 the lines indicate results from equations of state 14 90 The Lennard Jones truncated amp shifted LJTS potential is an often used alternative to the full Lennard Jones potential see Eq 1 The full and the truncated amp shifted Lennard Jones potential have to be kept strictly separate They are simply two different intermolecular potentials yielding different thermophysical properties The Lennard Jones truncated amp shifted potential is defined asVLJTS r VLJ r VLJ rend r rend0 r gt rend displaystyle V text LJTS r begin cases V text LJ r V text LJ r text end amp r leq r text end 0 amp r gt r text end end cases nbsp with VLJ r 4e sr 12 sr 6 displaystyle V text LJ r 4 varepsilon left left frac sigma r right 12 left frac sigma r right 6 right nbsp Hence the LJTS potential is truncated at rend displaystyle r mathrm end nbsp and shifted by the corresponding energy value VLJ rend displaystyle V mathrm LJ r mathrm end nbsp The latter is applied to avoid a discontinuity jump of the potential at rend displaystyle r mathrm end nbsp For the LJTS potential no long range interactions beyond rend displaystyle r mathrm end nbsp are required neither explicitly nor implicitly The most frequently used version of the Lennard Jones truncated amp shifted potential is the one with rend 2 5s displaystyle r mathrm end 2 5 sigma nbsp Nevertheless different rend displaystyle r mathrm end nbsp values have been used in the literature 91 92 93 94 Each LJTS potential with a given truncation radius rend displaystyle r mathrm end nbsp has to be considered as a potential and accordingly a substance of its own The LJTS potential is computationally significantly cheaper than the full Lennard Jones potential but still covers the essential physical features of matter the presence of a critical and a triple point soft repulsive and attractive interactions phase equilibria etc Therefore the LJTS potential is very frequently used for the testing of new algorithms simulation methods and new physical theories 95 96 97 Interestingly for homogeneous systems the intermolecular forces that are calculated from the LJ and the LJTS potential at a given distance are the same since dV dr displaystyle text d V text d r nbsp is the same whereas the potential energy and the pressure are affected by the shifting Also the properties of the LJTS substance may furthermore be affected by the chosen simulation algorithm i e MD or MC sampling this is in general not the case for the full Lennard Jones potential For the LJTS potential with rend 2 5s displaystyle r mathrm end 2 5 sigma nbsp the potential energy shift is approximately 1 60 of the dispersion energy at the potential well VLJ rend 2 5s 0 0163e displaystyle V mathrm LJ r mathrm end 2 5 sigma 0 0163 varepsilon nbsp The figure 8 shows the comparison of the vapor liquid equilibrium of the full Lennard Jones potential and the Lennard Jones truncated amp shifted potential The full Lennard Jones potential results prevail a significantly higher critical temperature and pressure compared to the LJTS potential results but the critical density is very similar 52 42 93 The vapor pressure and the enthalpy of vaporization are influenced more strongly by the long range interactions than the saturated densities This is due to the fact that the potential is manipulated mainly energetically by the truncation and shifting Extensions and modifications of the Lennard Jones potential editThe Lennard Jones potential as an archetype for intermolecular potentials has been used numerous times as starting point for the development of more elaborate or more generalized intermolecular potentials Various extensions and modifications of the Lennard Jones potential have been proposed in the literature a more extensive list is given in the interatomic potential article The following list refers only to several example potentials that are directly related to the Lennard Jones potential and are of both historic importance and still relevant for present research Mie potential The Mie potential is the generalized version of the Lennard Jones potential i e the exponents 12 and 6 are introduced as parameters lrep displaystyle lambda mathrm rep nbsp and lattr displaystyle lambda mathrm attr nbsp Especially thermodynamic derivative properties e g the compressibility and the speed of sound are known to be very sensitive to the steepness of the repulsive part of the intermolecular potential which can therefore be modeled more sophisticated by the Mie potential 80 The first explicit formulation of the Mie potential is attributed to Eduard Gruneisen 98 99 Hence the Mie potential was actually proposed before the Lennard Jones potential The Mie potential is named after Gustav Mie 23 Buckingham potential The Buckingham potential was proposed by Richard Buckingham The repulsive part of the Lennard Jones potential is therein replaced by an exponential function and it incorporates an additional parameter Stockmayer potential The Stockmayer potential is named after W H Stockmayer 100 The Stockmayer potential is a combination of a Lennard Jones potential superimposed by a dipole Hence Stockmayer particles are not spherically symmetric but rather have an important orientational structure Two center Lennard Jones potential The two center Lennard Jones potential consists of two identical Lennard Jones interaction sites same e displaystyle varepsilon nbsp s displaystyle sigma nbsp m displaystyle m nbsp that are bonded as a rigid body It is often abbreviated as 2CLJ Usually the elongation distance between the Lennard Jones sites is significantly smaller than the size parameter s displaystyle sigma nbsp Hence the two interaction sites are significantly fused Lennard Jones truncated amp splined potential The Lennard Jones truncated amp splined potential is a rarely used yet useful potential Similar to the more popular LJTS potential it is sturdily truncated at a certain end distance rend displaystyle r mathrm end nbsp and no long range interactions are considered beyond Opposite to the LJTS potential which is shifted such that the potential is continuous the Lennard Jones truncated amp splined potential is made continuous by using an arbitrary but favorable spline function See also edit nbsp Wikimedia Commons has media related to Lennard Jones potentials Comparison of force field implementations Embedded atom model Force field chemistry Molecular mechanics Morse potential and Morse Long range potential Virial expansionReferences edit Fischer Johann Wendland Martin October 2023 On the history of key empirical intermolecular potentials Fluid Phase Equilibria 573 113876 doi 10 1016 j fluid 2023 113876 ISSN 0378 3812 a b c d e Lenhard Johannes Stephan Simon Hasse Hans February 2024 A child of prediction On the History Ontology and Computation of the Lennard Jonesium Studies in History and Philosophy of Science 103 105 113 doi 10 1016 j shpsa 2023 11 007 PMID 38128443 S2CID 266440296 Jones J E 1924 On the determination of molecular fields I From the variation of the viscosity of a gas with temperature Proceedings of the Royal Society of London Series A Containing Papers of a Mathematical and Physical Character 106 738 441 462 Bibcode 1924RSPSA 106 441J doi 10 1098 rspa 1924 0081 ISSN 0950 1207 Jones J E 1924 On the determination of molecular fields II From the equation of state of a gas Proceedings of the Royal Society of London Series A Containing Papers of a Mathematical and Physical Character 106 738 463 477 Bibcode 1924RSPSA 106 463J doi 10 1098 rspa 1924 0082 ISSN 0950 1207 Lennard Jones J E 1931 09 01 Cohesion Proceedings of the Physical Society 43 5 461 482 Bibcode 1931PPS 43 461L doi 10 1088 0959 5309 43 5 301 ISSN 0959 5309 a b c Hill Terrell L 1956 Statistical mechanics principles and selected applications New York Dover Publications ISBN 0 486 65390 0 OCLC 15163657 a b c D C Rapaport 1 April 2004 The Art of Molecular Dynamics Simulation Cambridge University Press ISBN 978 0 521 82568 9 a b Frenkel D Smit B 2002 Understanding Molecular Simulation Second ed San Diego Academic Press ISBN 0 12 267351 4 a b c d e f g h i j k l m n o p q r s t u Stephan Simon Thol Monika Vrabec Jadran Hasse Hans 2019 10 28 Thermophysical Properties of the Lennard Jones Fluid Database and Data Assessment Journal of Chemical Information and Modeling 59 10 4248 4265 doi 10 1021 acs jcim 9b00620 ISSN 1549 9596 PMID 31609113 S2CID 204545481 a b Wood W W Parker F R 1957 Monte Carlo Equation of State of Molecules Interacting with the Lennard Jones Potential I A Supercritical Isotherm at about Twice the Critical Temperature The Journal of Chemical Physics 27 3 720 733 Bibcode 1957JChPh 27 720W doi 10 1063 1 1743822 ISSN 0021 9606 a b Rosenbluth Marshall N Rosenbluth Arianna W 1954 Further Results on Monte Carlo Equations of State The Journal of Chemical Physics 22 5 881 884 Bibcode 1954JChPh 22 881R doi 10 1063 1 1740207 ISSN 0021 9606 Alder B J Wainwright T E 1959 Studies in Molecular Dynamics I General Method The Journal of Chemical Physics 31 2 459 466 Bibcode 1959JChPh 31 459A doi 10 1063 1 1730376 ISSN 0021 9606 Rahman A 1964 10 19 Correlations in the Motion of Atoms in Liquid Argon Physical Review 136 2A A405 A411 Bibcode 1964PhRv 136 405R doi 10 1103 PhysRev 136 A405 ISSN 0031 899X a b c d e f g h i Stephan Simon Staubach Jens Hasse Hans 2020 Review and comparison of equations of state for the Lennard Jones fluid Fluid Phase Equilibria 523 112772 doi 10 1016 j fluid 2020 112772 S2CID 224844789 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 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 S2CID 18734898 a b c Stephan Simon Horsch Martin T Vrabec Jadran Hasse Hans 2019 07 03 MolMod an open access database of force fields for molecular simulations of fluids Molecular Simulation 45 10 806 814 arXiv 1904 05206 doi 10 1080 08927022 2019 1601191 ISSN 0892 7022 S2CID 119199372 a b c Eggimann Becky L Sunnarborg Amara J Stern Hudson D Bliss Andrew P Siepmann J Ilja 2014 01 02 An online parameter and property database for the TraPPE force field Molecular Simulation 40 1 3 101 105 doi 10 1080 08927022 2013 842994 ISSN 0892 7022 S2CID 95716947 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 Eisenschitz R London F 1930 07 01 Uber das Verhaltnis der van der Waalsschen Krafte zu den homoopolaren Bindungskraften Zeitschrift fur Physik in German 60 7 491 527 Bibcode 1930ZPhy 60 491E doi 10 1007 BF01341258 ISSN 0044 3328 S2CID 125644826 Rowlinson J S 2006 11 20 The evolution of some statistical mechanical ideas Molecular Physics 104 22 24 3399 3410 Bibcode 2006MolPh 104 3399R doi 10 1080 00268970600965835 ISSN 0026 8976 S2CID 119942778 Abascal J L F Vega C 2005 12 15 A general purpose model for the condensed phases of water TIP4P 2005 The Journal of Chemical Physics 123 23 234505 Bibcode 2005JChPh 123w4505A doi 10 1063 1 2121687 ISSN 0021 9606 PMID 16392929 a b Mie Gustav 1903 Zur kinetischen Theorie der einatomigen Korper Annalen der Physik in German 316 8 657 697 Bibcode 1903AnP 316 657M doi 10 1002 andp 19033160802 Tang K T Toennies J Peter 1984 04 15 An improved simple model for the van der Waals potential based on universal damping functions for the dispersion coefficients The Journal of Chemical Physics 80 8 3726 3741 Bibcode 1984JChPh 80 3726T doi 10 1063 1 447150 ISSN 0021 9606 Lafitte Thomas Apostolakou Anastasia Avendano Carlos Galindo Amparo Adjiman Claire S Muller Erich A Jackson George 2013 10 21 Accurate statistical associating fluid theory for chain molecules formed from Mie segments The Journal of Chemical Physics 139 15 Bibcode 2013JChPh 139o4504L doi 10 1063 1 4819786 hdl 10044 1 12859 ISSN 0021 9606 PMID 24160524 Stephan Simon Staubach Jens Hasse Hans November 2020 Review and comparison of equations of state for the Lennard Jones fluid Fluid Phase Equilibria 523 112772 doi 10 1016 j fluid 2020 112772 S2CID 224844789 Smit B Frenkel D 1991 04 15 Vapor liquid equilibria of the two dimensional Lennard Jones fluid s The Journal of Chemical Physics 94 8 5663 5668 Bibcode 1991JChPh 94 5663S doi 10 1063 1 460477 ISSN 0021 9606 S2CID 1580499 Scalise Osvaldo H June 2001 Type I gas liquid equilibria of a two dimensional Lennard Jones binary mixture Fluid Phase Equilibria 182 1 2 59 64 doi 10 1016 s0378 3812 01 00380 6 ISSN 0378 3812 Hloucha M Sandler S I November 1999 Phase diagram of the four dimensional Lennard Jones fluid The Journal of Chemical Physics 111 17 8043 8047 Bibcode 1999JChPh 111 8043H doi 10 1063 1 480138 ISSN 0021 9606 a b Zhen Shu Davies G J 1983 08 16 Calculation of the Lennard Jonesn m potential energy parameters for metals Physica Status Solidi A in German 78 2 595 605 Bibcode 1983PSSAR 78 595Z doi 10 1002 pssa 2210780226 a b Halicioglu T Pound G M 1975 08 16 Calculation of potential energy parameters form crystalline state properties Physica Status Solidi A 30 2 619 623 Bibcode 1975PSSAR 30 619H doi 10 1002 pssa 2210300223 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 Mendelev M I Han S Srolovitz D J Ackland G J Sun D Y Asta M 2003 Development of new interatomic potentials appropriate for crystalline and liquid iron Philosophical Magazine 83 35 3977 3994 Bibcode 2003PMag 83 3977A doi 10 1080 14786430310001613264 ISSN 1478 6435 S2CID 4119718 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 a b c Allen Michael P Tildesley Dominic J 2017 11 23 Computer Simulation of Liquids Oxford Scholarship Online doi 10 1093 oso 9780198803195 001 0001 ISBN 9780198803195 a b c d e f g h Schultz Andrew J Kofke David A 2018 11 28 Comprehensive high precision high accuracy equation of state and coexistence properties for classical Lennard Jones crystals and low temperature fluid phases The Journal of Chemical Physics 149 20 204508 Bibcode 2018JChPh 149t4508S doi 10 1063 1 5053714 ISSN 0021 9606 PMID 30501268 S2CID 54629914 a b Schultz Andrew J Kofke David A 2020 08 07 Erratum Comprehensive high precision high accuracy equation of state and coexistence properties for classical Lennard Jones crystals and low temperature fluid phases J Chem Phys 149 204508 2018 The Journal of Chemical Physics 153 5 059901 doi 10 1063 5 0021283 ISSN 0021 9606 PMID 32770918 a b Schappals Michael Mecklenfeld Andreas Kroger Leif Botan Vitalie Koster Andreas Stephan Simon Garcia Edder J Rutkai Gabor Raabe Gabriele Klein Peter Leonhard Kai 2017 09 12 Round Robin Study Molecular Simulation of Thermodynamic Properties from Models with Internal Degrees of Freedom Journal of Chemical Theory and Computation 13 9 4270 4280 doi 10 1021 acs jctc 7b00489 ISSN 1549 9618 PMID 28738147 Loeffler Hannes H Bosisio Stefano Duarte Ramos Matos Guilherme Suh Donghyuk Roux Benoit Mobley David L Michel Julien 2018 11 13 Reproducibility of Free Energy Calculations across Different Molecular Simulation Software Packages Journal of Chemical Theory and Computation 14 11 5567 5582 doi 10 1021 acs jctc 8b00544 hdl 20 500 11820 52d85d71 d3df 468b 8f88 9c52e83da1f1 ISSN 1549 9618 PMID 30289712 S2CID 52923832 a b Lenhard Johannes Kuster Uwe 2019 Reproducibility and the Concept of Numerical Solution Minds and Machines 29 1 19 36 doi 10 1007 s11023 019 09492 9 ISSN 0924 6495 S2CID 59159685 Shi Wei Johnson J Karl 2001 09 15 Histogram reweighting and finite size scaling study of the Lennard Jones fluids Fluid Phase Equilibria 187 188 171 191 doi 10 1016 S0378 3812 01 00534 9 ISSN 0378 3812 a b 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 a b Metropolis Nicholas Rosenbluth Arianna W Rosenbluth Marshall N Teller Augusta H Teller Edward 1953 Equation of State Calculations by Fast Computing Machines The Journal of Chemical Physics 21 6 1087 1092 Bibcode 1953JChPh 21 1087M doi 10 1063 1 1699114 ISSN 0021 9606 OSTI 4390578 S2CID 1046577 Stephan Simon Thol Monika Vrabec Jadran Hasse Hans 2019 10 28 Thermophysical Properties of the Lennard Jones Fluid Database and Data Assessment Journal of Chemical Information and Modeling 59 10 4248 4265 doi 10 1021 acs jcim 9b00620 ISSN 1549 9596 PMID 31609113 S2CID 204545481 a b Koster Andreas Mausbach Peter Vrabec Jadran 2017 10 10 Premelting solid fluid equilibria and thermodynamic properties in the high density region based on the Lennard Jones potential The Journal of Chemical Physics 147 14 144502 Bibcode 2017JChPh 147n4502K doi 10 1063 1 4990667 ISSN 0021 9606 PMID 29031254 a b c d e f g h i Stephan Simon Deiters Ulrich K 2020 08 20 Characteristic Curves of the Lennard Jones Fluid International Journal of Thermophysics 41 10 147 Bibcode 2020IJT 41 147S doi 10 1007 s10765 020 02721 9 ISSN 1572 9567 PMC 7441092 PMID 32863513 a b Deiters Ulrich K Neumaier Arnold 2016 08 11 Computer Simulation of the Characteristic Curves of Pure Fluids Journal of Chemical amp Engineering Data 61 8 2720 2728 doi 10 1021 acs jced 6b00133 ISSN 0021 9568 a b Agrawal Rupal Kofke David A 1995 Thermodynamic and structural properties of model systems at solid fluid coexistence II Melting and sublimation of the Lennard Jones system Molecular Physics 85 1 43 59 doi 10 1080 00268979500100921 ISSN 0026 8976 Brown E H 1960 On the thermodynamic properties of fluids Bulletin de l Institut International du Froid Annexe 1960 1 169 178 Apfelbaum E M Vorob ev V S 2020 06 18 The Line of the Unit Compressibility Factor Zeno Line for Crystal States The Journal of Physical Chemistry B 124 24 5021 5027 doi 10 1021 acs jpcb 0c02749 ISSN 1520 6106 PMID 32437611 S2CID 218835048 Apfelbaum E M Vorob ev V S Martynov G A 2008 Regarding the Theory of the Zeno Line The Journal of Physical Chemistry A 112 26 6042 6044 Bibcode 2008JPCA 112 6042A doi 10 1021 jp802999z ISSN 1089 5639 PMID 18543889 a b c d e Stephan Simon Hasse Hans 2020 06 01 Influence of dispersive long range interactions on properties of vapour liquid equilibria and interfaces of binary Lennard Jones mixtures Molecular Physics 118 9 10 e1699185 Bibcode 2020MolPh 11899185S doi 10 1080 00268976 2019 1699185 ISSN 0026 8976 S2CID 214174102 Nicolas J J Gubbins K E Streett W B Tildesley D J 1979 Equation of state for the Lennard Jones fluid Molecular Physics 37 5 1429 1454 Bibcode 1979MolPh 37 1429N doi 10 1080 00268977900101051 ISSN 0026 8976 Feng Chao Schultz Andrew J Chaudhary Vipin Kofke David A 2015 07 28 Eighth to sixteenth virial coefficients of the Lennard Jones model The Journal of Chemical Physics 143 4 044504 Bibcode 2015JChPh 143d4504F doi 10 1063 1 4927339 ISSN 0021 9606 PMID 26233142 Schultz Andrew J Kofke David A 2009 11 10 Sixth seventh and eighth virial coefficients of the Lennard Jones model Molecular Physics 107 21 2309 2318 Bibcode 2009MolPh 107 2309S doi 10 1080 00268970903267053 ISSN 0026 8976 S2CID 94811614 Bell Ian H Messerly Richard Thol Monika Costigliola Lorenzo Dyre Jeppe C 2019 07 25 Modified Entropy Scaling of the Transport Properties of the Lennard Jones Fluid The Journal of Physical Chemistry B 123 29 6345 6363 doi 10 1021 acs jpcb 9b05808 ISSN 1520 6106 PMC 7147083 PMID 31241958 Lautenschlaeger Martin P Hasse Hans 2019 Transport properties of the Lennard Jones truncated and shifted fluid from non equilibrium molecular dynamics simulations Fluid Phase Equilibria 482 38 47 doi 10 1016 j fluid 2018 10 019 S2CID 106113718 a b c Travesset Alex 2014 10 28 Phase diagram of power law and Lennard Jones systems Crystal phases The Journal of Chemical Physics 141 16 164501 Bibcode 2014JChPh 141p4501T doi 10 1063 1 4898371 ISSN 0021 9606 PMID 25362319 Hansen Jean Pierre Verlet Loup 1969 08 05 Phase Transitions of the Lennard Jones System Physical Review 184 1 151 161 Bibcode 1969PhRv 184 151H doi 10 1103 PhysRev 184 151 ISSN 0031 899X Longuet Higgins H C 1951 02 07 The statistical thermodynamics of multicomponent systems Proceedings of the Royal Society of London Series A Mathematical and Physical Sciences 205 1081 247 269 Bibcode 1951RSPSA 205 247L doi 10 1098 rspa 1951 0028 ISSN 0080 4630 S2CID 202575459 Leland T W Rowlinson J S Sather G A 1968 Statistical thermodynamics of mixtures of molecules of different sizes Transactions of the Faraday Society 64 1447 doi 10 1039 tf9686401447 ISSN 0014 7672 Mansoori G Ali Leland Thomas W 1972 Statistical thermodynamics of mixtures A new version for the theory of conformal solution Journal of the Chemical Society Faraday Transactions 2 68 320 doi 10 1039 f29726800320 ISSN 0300 9238 Rowlinson J S Swinton F L 1982 Liquids and liquid mixtures Third ed London Butterworth a b Deiters Ulrich K Kraska Thomas 2012 High pressure fluid phase equilibria phenomenology and computation 1st ed Amsterdam Elsevier ISBN 978 0 444 56354 5 OCLC 787847134 Schnabel Thorsten Vrabec Jadran Hasse Hans 2007 Unlike Lennard Jones parameters for vapor liquid equilibria Journal of Molecular Liquids 135 1 3 170 178 arXiv 0904 4436 doi 10 1016 j molliq 2006 12 024 S2CID 16111477 Lorentz H A 1881 Ueber die Anwendung des Satzes vom Virial in der kinetischen Theorie der Gase Annalen der Physik in German 248 1 127 136 Bibcode 1881AnP 248 127L doi 10 1002 andp 18812480110 van Konynenburg P H Scott R L 1980 12 18 Critical lines and phase equilibria in binary van der Waals mixtures Philosophical Transactions of the Royal Society of London Series A Mathematical and Physical Sciences 298 1442 495 540 Bibcode 1980RSPTA 298 495K doi 10 1098 rsta 1980 0266 ISSN 0080 4614 S2CID 122538015 Potoff Jeffrey J Panagiotopoulos Athanassios Z 1998 12 22 Critical point and phase behavior of the pure fluid and a Lennard Jones mixture The Journal of Chemical Physics 109 24 10914 10920 Bibcode 1998JChPh 10910914P doi 10 1063 1 477787 ISSN 0021 9606 Protsenko Sergey P Baidakov Vladimir G 2016 Binary Lennard Jones mixtures with highly asymmetric interactions of the components 1 Effect of the energy parameters on phase equilibria and properties of liquid gas interfaces Fluid Phase Equilibria 429 242 253 doi 10 1016 j fluid 2016 09 009 Protsenko Sergey P Baidakov Vladimir G Bryukhanov Vasiliy M 2016 Binary Lennard Jones mixtures with highly asymmetric interactions of the components 2 Effect of the particle size on phase equilibria and properties of liquid gas interfaces Fluid Phase Equilibria 430 67 74 doi 10 1016 j fluid 2016 09 022 Stephan Simon Hasse Hans 2020 01 23 Molecular interactions at vapor liquid interfaces Binary mixtures of simple fluids Physical Review E 101 1 012802 Bibcode 2020PhRvE 101a2802S doi 10 1103 PhysRevE 101 012802 ISSN 2470 0045 PMID 32069593 S2CID 211192904 a b Lamm Monica H Hall Carol K 2002 Equilibria between solid liquid and vapor phases in binary Lennard Jones mixtures Fluid Phase Equilibria 194 197 197 206 doi 10 1016 S0378 3812 01 00650 1 a b Lamm Monica H Hall Carol K 2001 Monte Carlo simulations of complete phase diagrams for binary Lennard Jones mixtures Fluid Phase Equilibria 182 1 2 37 46 doi 10 1016 S0378 3812 01 00378 8 a b Hitchcock Monica R Hall Carol K 1999 06 15 Solid liquid phase equilibrium for binary Lennard Jones mixtures The Journal of Chemical Physics 110 23 11433 11444 Bibcode 1999JChPh 11011433H doi 10 1063 1 479084 ISSN 0021 9606 Jungblut Swetlana Dellago Christoph 2011 03 14 Crystallization of a binary Lennard Jones mixture The Journal of Chemical Physics 134 10 104501 Bibcode 2011JChPh 134j4501J doi 10 1063 1 3556664 ISSN 0021 9606 PMID 21405169 a b Lamm Monica H Hall Carol K 2004 Effect of pressure on the complete phase behavior of binary mixtures AIChE Journal 50 1 215 225 Bibcode 2004AIChE 50 215L doi 10 1002 aic 10020 ISSN 0001 1541 Cotterman R L Prausnitz J M 1986 Molecular thermodynamics for fluids at low and high densities Part II Phase equilibria for mixtures containing components with large differences in molecular size or potential energy AIChE Journal 32 11 1799 1812 Bibcode 1986AIChE 32 1799C doi 10 1002 aic 690321105 ISSN 0001 1541 S2CID 96417239 Muller Andreas Winkelmann Jochen Fischer Johann 1996 Backone family of equations of state 1 Nonpolar and polar pure fluids AIChE Journal 42 4 1116 1126 Bibcode 1996AIChE 42 1116M doi 10 1002 aic 690420423 ISSN 0001 1541 Weingerl Ulrike Wendland Martin Fischer Johann Muller Andreas Winkelmann Jochen 2001 Backone family of equations of state 2 Nonpolar and polar fluid mixtures AIChE Journal 47 3 705 717 Bibcode 2001AIChE 47 705W doi 10 1002 aic 690470317 a b Lafitte Thomas Apostolakou Anastasia Avendano Carlos Galindo Amparo Adjiman Claire S Muller Erich A Jackson George 2013 10 16 Accurate statistical associating fluid theory for chain molecules formed from Mie segments The Journal of Chemical Physics 139 15 154504 Bibcode 2013JChPh 139o4504L doi 10 1063 1 4819786 hdl 10044 1 12859 ISSN 0021 9606 PMID 24160524 Blas F J Vega L F 1997 Thermodynamic behaviour of homonuclear and heteronuclear Lennard Jones chains with association sites from simulation and theory Molecular Physics 92 1 135 150 Bibcode 1997MolPh 92 135F doi 10 1080 002689797170707 ISSN 0026 8976 Kraska Thomas Gubbins Keith E 1996 Phase Equilibria Calculations with a Modified SAFT Equation of State 1 Pure Alkanes Alkanols and Water Industrial amp Engineering Chemistry Research 35 12 4727 4737 doi 10 1021 ie9602320 ISSN 0888 5885 Ghonasgi D Chapman Walter G 1994 Prediction of the properties of model polymer solutions and blends AIChE Journal 40 5 878 887 Bibcode 1994AIChE 40 878G doi 10 1002 aic 690400514 ISSN 0001 1541 Mecke M Muller A Winkelmann J Vrabec J Fischer J Span R Wagner W 1996 03 01 An accurate Van der Waals type equation of state for the Lennard Jones fluid International Journal of Thermophysics 17 2 391 404 Bibcode 1996IJT 17 391M doi 10 1007 BF01443399 ISSN 1572 9567 S2CID 123304062 a b c Kolafa Jiri Nezbeda Ivo 1994 The Lennard Jones fluid an accurate analytic and theoretically based equation of state Fluid Phase Equilibria 100 1 34 doi 10 1016 0378 3812 94 80001 4 Thol Monika Rutkai Gabor Koster Andreas Lustig Rolf Span Roland Vrabec Jadran 2016 Equation of State for the Lennard Jones Fluid Journal of Physical and Chemical Reference Data 45 2 023101 Bibcode 2016JPCRD 45b3101T doi 10 1063 1 4945000 ISSN 0047 2689 Gottschalk Matthias 2019 12 01 An EOS for the Lennard Jones fluid A virial expansion approach AIP Advances 9 12 125206 Bibcode 2019AIPA 9l5206G doi 10 1063 1 5119761 ISSN 2158 3226 Johnson J Karl Zollweg John A Gubbins Keith E 1993 02 20 The Lennard Jones equation of state revisited Molecular Physics 78 3 591 618 Bibcode 1993MolPh 78 591J doi 10 1080 00268979300100411 ISSN 0026 8976 Vrabec Jadran Kedia Gaurav Kumar Fuchs Guido Hasse Hans 2006 05 10 Comprehensive study of the vapour liquid coexistence of the truncated and shifted Lennard Jones fluid including planar and spherical interface properties Molecular Physics 104 9 1509 1527 Bibcode 2006MolPh 104 1509V doi 10 1080 00268970600556774 ISSN 0026 8976 S2CID 96606562 Heier Michaela Stephan Simon Liu Jinlu Chapman Walter G Hasse Hans Langenbach Kai 2018 08 18 Equation of state for the Lennard Jones truncated and shifted fluid with a cut off radius of 2 5 s based on perturbation theory and its applications to interfacial thermodynamics Molecular Physics 116 15 16 2083 2094 Bibcode 2018MolPh 116 2083H doi 10 1080 00268976 2018 1447153 ISSN 0026 8976 S2CID 102956189 Shaul Katherine R S Schultz Andrew J Kofke David A 2010 The effect of truncation and shift on virial coefficients of Lennard Jones potentials Collection of Czechoslovak Chemical Communications 75 4 447 462 doi 10 1135 cccc2009113 ISSN 1212 6950 Shi Wei Johnson J Karl 2001 Histogram reweighting and finite size scaling study of the Lennard Jones fluids Fluid Phase Equilibria 187 188 171 191 doi 10 1016 S0378 3812 01 00534 9 a b Dunikov D O Malyshenko S P Zhakhovskii V V 2001 10 08 Corresponding states law and molecular dynamics simulations of the Lennard Jones fluid The Journal of Chemical Physics 115 14 6623 6631 Bibcode 2001JChPh 115 6623D doi 10 1063 1 1396674 ISSN 0021 9606 Livia B Partay Christoph Ortner Albert P Bartok Chris J Pickard and Gabor Csanyi Polytypism in the ground state structure of the Lennard Jonesium Physical Chemistry Chemical Physics 19 19369 2017 Tchipev Nikola Seckler Steffen Heinen Matthias Vrabec Jadran Gratl Fabio Horsch Martin Bernreuther Martin Glass Colin W Niethammer Christoph Hammer Nicolay Krischok Bernd 2019 TweTriS Twenty trillion atom simulation The International Journal of High Performance Computing Applications 33 5 838 854 doi 10 1177 1094342018819741 ISSN 1094 3420 S2CID 59345875 Stephan Simon Liu Jinlu Langenbach Kai Chapman Walter G Hasse Hans 2018 Vapor Liquid Interface of the Lennard Jones Truncated and Shifted Fluid Comparison of Molecular Simulation Density Gradient Theory and Density Functional Theory The Journal of Physical Chemistry C 122 43 24705 24715 doi 10 1021 acs jpcc 8b06332 ISSN 1932 7447 S2CID 105759822 Kob Walter Andersen Hans C 1995 05 01 Testing mode coupling theory for a supercooled binary Lennard Jones mixture I The van Hove correlation function Physical Review E 51 5 4626 4641 arXiv cond mat 9501102 Bibcode 1995PhRvE 51 4626K doi 10 1103 PhysRevE 51 4626 PMID 9963176 S2CID 17662741 Gruneisen Edward 1911 Das Verhaltnis der thermischen Ausdehnung zur spezifischen Warme fester Elemente Zeitschrift fur Elektrochemie und angewandte physikalische Chemie 17 17 737 739 doi 10 1002 bbpc 191100004 S2CID 178760389 Gruneisen E 1912 Theorie des festen Zustandes einatomiger Elemente Annalen der Physik in German 344 12 257 306 Bibcode 1912AnP 344 257G doi 10 1002 andp 19123441202 Stockmayer W H 1941 05 01 Second Virial Coefficients of Polar Gases The Journal of Chemical Physics 9 5 398 402 Bibcode 1941JChPh 9 398S doi 10 1063 1 1750922 ISSN 0021 9606 External links editLennard Jones model on SklogWiki Retrieved from https en wikipedia org w index php title Lennard Jones potential amp oldid 1216631372, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.