Type: | Package |
Title: | Estimation of Gas Properties from the Lennard-Jones Potential |
Version: | 0.4 |
Date: | 2025-07-27 |
Depends: | R (≥ 2.12.0), methods, Bessel |
Description: | Estimation of gas transport properties (viscosity, diffusion, thermal conductivity) using Chapman-Enskok theory (Chapman and Larmor 1918, <doi:10.1098/rsta.1918.0005>) and of the second virial coefficient (Vargas et al. 2001, <doi:10.1016/s0378-4371(00)00362-9>) using the Lennard-Jones (12-6) potential. Up to the third order correction is taken into account for viscosity and thermal conductivity. It is also possible to calculate the binary diffusion coefficients of polar and non-polar gases in non-polar bath gases (Brown et al. 2011, <doi:10.1016/j.pecs.2010.12.001>). 16 collision integrals are calculated with four digit accuracy over the reduced temperature range [0.3, 400] using an interpolation function of Kim and Monroe (2014, <doi:10.1016/j.jcp.2014.05.018>). |
License: | GPL-3 |
URL: | https://github.com/langenbergstefan/chapensk |
BugReports: | https://github.com/langenbergstefan/chapensk/issues |
NeedsCompilation: | no |
LazyData: | true |
Packaged: | 2025-07-27 07:31:01 UTC; stefan |
Author: | Stefan Langenberg |
Maintainer: | Stefan Langenberg <langenberg@uni-bonn.de> |
Repository: | CRAN |
Date/Publication: | 2025-07-29 13:50:08 UTC |
Estimation of Gas Properties from the Lennard-Jones Potential
Description
Estimation of gas transport properties (viscosity, diffusion, thermal conductivity) using Chapman-Enskok theory (Chapman and Larmor 1918, <doi:10.1098/rsta.1918.0005>) and of the second virial coefficient (Vargas et al. 2001, <doi:10.1016/s0378-4371(00)00362-9>) using the Lennard-Jones (12-6) potential. Up to the third order correction is taken into account for viscosity and thermal conductivity. It is also possible to calculate the binary diffusion coefficients of polar and non-polar gases in non-polar bath gases (Brown et al. 2011, <doi:10.1016/j.pecs.2010.12.001>). 16 collision integrals are calculated with four digit accuracy over the reduced temperature range [0.3, 400] using an interpolation function of Kim and Monroe (2014, <doi:10.1016/j.jcp.2014.05.018>).
Introduction
Transport properties, such as viscosity, diffusion and thermal conductivity, play a crucial role in the modeling of combustion processes and chemical reactions. They depend on the intermolecular potential. In practice it is not necessary to have a detailed calculation of the intermolecular potential for the calculation of transport properties.
The interaction between spherical gas particles without a dipole moment can be described by the Lennard-Jones potential. It is given by the following equation:
U(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, ε is the depth of the potential well and σ is the distance at which the
particle-particle potential energy U
is zero. The theory of van der Waals interaction gives the exponent 6 for the attractive term (London, 1937).
The 12 exponent in the repulsive term is largely a matter of computational practicality, though it does represent the general nature of Pauli repulsion.
The Lennard-Jones potential only can be used for non-polar molecules, but sometimes is also used for polar molecules. However, for the latter the Stockmayer (12-6-3) potential is more appropriate (Mourits and Rummens, 1977).
For monoatomic gases σ and ε are independent of temperature. However, for non-monoatomic gases, averaging over different orientations and vibrational states gives temperature dependent parameters σ and ε (Zarkova and Hohm, 2002). The temperature dependency is simplified in this package as follows by a linear temperature coefficient ζ:
\sigma(T) = \sigma + \zeta T
\varepsilon(T) = \varepsilon\left(\frac{\sigma}{\sigma + \zeta T}\right)^6
The Chapman-Enskog theory is a theoretical framework used to describe the transport properties of gases, such as viscosity, thermal conductivity, and diffusion coefficients. The theory is based on the idea that the properties of a gas can be related to the collisional interactions between individual gas molecules (Chapman and Larmor 1918).
Collision integrals are mathematical expressions that arise in the Chapman-Enskog theory. They quantify the effects of molecular collisions on the transport properties of a gas.
Methods
An object-oriented framework has been developed to calculate transport properties from potential parameters and vice versa. A class Gas has been defined for the calculation of the properties of gas. To facilitate calculations, a gas data set is provided for the properties of some common gases. The CollisionIntegral class is used to calculate collision integrals using an interpolation function and fit parameters from dataset coefficients_collisionintegral.
Results
Lennard-Jones parameters of non-polar molecules can be estimated using high quality of viscosity and second virial coefficients. This is demonstrated for ethane, see data set ethane_data. For polar molecules Lennard-Jones parameters for the Van der Walls interaction part can be estimated from measurements of binary diffusion coefficients, see data set binary_diffusion.
Nomenclature
Symbol | Description | Unit | Global variable |
b | temperature coefficient of diffusion | - |
|
B | second virial coefficient | m3 |
|
D | diffusion coefficient | m2/2 |
|
k | Boltzmann constant | J/K | pkg.env$k |
m | molecular mass | kg |
|
M | relative molecular mass | - |
|
N_a | Avogadro constant | 1/mol | pkg.env$Na |
n | mole | mol |
|
p | pressure | Pa |
|
p_0 | standard pressure 101325 Pa | Pa | pkg.env$p0 |
p_c | critical pressure | Pa |
|
R | gas constant | J/(K.mol) | pkh.env$R |
T | temperature | K |
|
T_0 | standard temperature 273.15 K | K | pkg.env$T0 |
T_c | critical temperature | K |
|
V | gas volume | m3 |
|
\alpha | polarizability | Ao3 |
|
\bar{\alpha} | reduced polarizability | - |
|
\eta | dynamic viscosity | Pa.s |
|
\epsilon | permittivity of vacuum | F/m | pkg.env$eps0 |
\varepsilon | depth of potential well | J |
|
\kappa | thermal conductivity | W/(m.K) |
|
\mu | dipole moment | D |
|
\bar{\mu} | reduced dipole moment | - |
|
\Omega | reduced collision integral | - |
|
\rho | gas density | kg/m3 |
|
\rho_c | critical density | mol/l |
|
\Theta | reduced temperature | - |
|
\sigma | distance at which the potential energy is zero | Ao |
|
\xi | scaling parameter | - |
|
\zeta | temperature coefficient of \sigma | Ao/K
|
Physical units are displayed in UCUM notation.
Author(s)
Stefan Langenberg [aut, cre] (ORCID: <https://orcid.org/0000-0001-5817-5469>)
Maintainer: Stefan Langenberg <langenberg@uni-bonn.de>
References
Brown NJ, Bastien LAJ, Price PN. Transport properties for combustion modeling. Progress in Energy and Combustion Science 2011;37:565-82. doi:10.1016/j.pecs.2010.12.001.
Chapman S, Larmor J. V. On the kinetic theory of a gas. Part II. A composite monatomic gas: diffusion, viscosity, and thermal conduction. Philosophical Transactions of the Royal Society of London. Series A 1918;217:11597. doi:10.1098/rsta.1918.0005.
London F. The general theory of molecular forces. Transactions of the Faraday Society 1937;33:8b. doi:10.1039/tf937330008b.
Mourits FM, Rummens FHA. A critical evaluation of Lennard-Jones and Stockmayer potential parameters and of some correlation methods. Can. J. Chem. 1977;55:300720. doi:10.1139/v77-418.
Zarkova L, Hohm U. pVT-Second Virial Coefficients B(T)
, Viscosity η and Self-Diffusion \rho D(T)
of the Gases: BF3, CF4,
SiF4, CCl4, SiCl4, SF6, MoF6, WF6, UF6, C(CH3)4, and Si(CH3)4 Determined by Means of an Isotropic Temperature-Dependent Potential.
Journal of Physical and Chemical Reference Data 2002;31:183216. doi:10.1063/1.1433462.
Class "CollisionIntegral"
Description
Class for the empirical interpolation functions for 16 reduced Lennard-Jones (12-6) collision integrals
in the range 0.3 <= Θ <= 400.
Collision integrals as function of reduced temperature Θ
Fields
l
:order of collision integral
l
s
:order of collision integral
s
coeff
:coefficients for collision integrals for order
l
ands
, seecoefficients_collisionintegral
A
:model parameter A
B
:model parameter B1 ... B6
C
:model parameter C1 ... C6
Methods
initialize(s, l)
:Initialize reduced collision integral of order
l
ands
Omega(Theta)
:Calculates reduced collision integral Ω(l,s) at reduced temperature
\Theta
using the empirical interpolation function
References
Kim SU, Monroe CW. High-accuracy calculations of sixteen collision integrals for Lennard-Jones (12-6) gases and their interpolation to parameterize neon, argon, and krypton. Journal of Computational Physics 2014;273:35873. doi:10.1016/j.jcp.2014.05.018.
Examples
O11 <- CollisionIntegral(l=1,s=1)
print(O11$Omega(0.3))
Class "Gas"
Description
Reference class for gas with methods for the estimation of transport properties. The equations for the higher order corrections are taken from Kim and Monroe 2014. The second virial coefficient is calculated using the equation given by Vargas et al. 2001.
Fields
name
:Chemical name
M
:relative molecular mass
M
m
:Mass
m
of one gas particle inkg
sigma
:Lennard-Jones Parameter σ at
T = 0
zeta
:Change of σ with temperature. For the noble gases
\zeta = 0
. Due to the vibrational excitation of molecules, σ is not independent of temperature but increasing withT
. If\sigma_0
is σ atT = 0
, σ atT
is given by\sigma = \sigma_0 + \zeta T
.epsk
:Well depth of the Lennard-Jones potential
\varepsilon/k
atT = 0
inK
.dipole_moment
:dipole moment
\mu
inDebye
.polarizability
:polarizability
\alpha
inAo3
.
Methods
check()
:Checks for presence of required data.
B(T)
:Second virial coefficient
B(T)
in unitsm3/mol
. The second virial coefficient provides systematic corrections to the ideal gas law. The second virial coefficientB
depends only on the pair interaction between the particles (Vargas et. al. 2001). It is calculated from the modified Bessel functionI
(BesselI):B(\Theta) = \frac{\sqrt{2}\pi^2\sigma^3}{3\Theta} \left[ I_{-3/4}\left(\frac{1}{2\Theta}\right) + I_{3/4}\left(\frac{1}{2\Theta}\right) - I_{1/4}\left(\frac{1}{2\Theta}\right) - I_{-1/4}\left(\frac{1}{2\Theta}\right) \right]
density(p=p0,T=T0)
:Gas density in units
kg/m3
incorporating the second virial correction.\rho = \frac{M}{2B(\Theta)} \left(\sqrt{1 + \frac{4pB(\Theta)}{RT}} - 1\right)
diffusion(p=p0, T=T0, second_order_correction=TRUE)
:-
Calculates the self diffusion coefficient in
m2/s
.D = \frac{3 \sqrt{\pi m kT}}{8\pi\sigma^2 \rho(p,T)\Omega^{(1,1)}(\Theta)} f_{D}^{(n)}
f_{D}^{(n)}
is the second order correction term. Ifthird_order_correction=FALSE
this term is set to 1. thermal_conductivity(T, third_order_correction=TRUE)
:-
Calculates the thermal conductivity for a monoatomic gas in
W/(m.K)
.\kappa = \frac{75k}{64 \sigma^2 \Omega^{(2,2)}(\Theta)} \sqrt{\frac{kT}{\pi m}} f_{\kappa}^{(n)}
f_{\kappa}^{(n)}
is the third order correction term. Ifthird_order_correction=FALSE
this term is set to 1. For a polyatomic gas the thermal conductivity can be calculated from the expression for a monoatomic gas using the classical Eucken correction (Bechtel et al. 2020)\kappa_p = \kappa \left(\frac{4 c_v}{15 R} + \frac{3}{5} \right)
where
c_v
is the molar heat capacity of the polyatomic gas at constant volume. viscosity(T,third_order_correction=TRUE)
:-
The viscosity
\eta
is given gas as\eta = \frac{5\sqrt{\pi m T}}{16 \pi \sigma^2 \Omega^{(2,2)}(\Theta)} f_{\eta}^{(n)}.
f_{\eta}^{(n)}
is the third order correction term. Ifthird_order_correction=FALSE
this term is set to 1. binary_diffusion(p=p0, T=T0, bathGas)
:-
Binary Diffusion coefficient of gas 1 in a bath gas 2
bathGas
. The gas 1 may be polar or nonpolar. The bath gas 2 must be nonpolar (Brown et al. 2010, Langenberg et al. 2020). The binary diffusion coefficientD_{12}
is given asD_{12} = \frac{3}{16} \sqrt{\frac{2 \pi k T (m_1 + m_2)}{m_1 m_2}} \left(\frac{kT}{\pi \sigma_{12}^2 \Omega^{(1,1)}(\Theta)p} \right)
The influence of the dipole moment of the gas is treated by the scaling parameter (Brown et al. 2011)
\xi = 1 + \frac{\bar{\alpha}\bar{\mu}^2}{16 \pi \epsilon_0} \sqrt{\frac{\varepsilon_1}{\varepsilon_2}},
where the reduced dipole moment is given by
\bar{\mu}^2=\frac{\mu^2}{\varepsilon_1 \sigma_1^3}
and the reduced polarizability of the bath gas is given by
\bar{\alpha} = \frac{\alpha}{\sigma_2^3}.
The scaling parameter
\xi
is used in the following combination rules to calculate the well depth of the effective interaction potential\varepsilon_{12} = \xi^2 \sqrt{\varepsilon_1 \varepsilon_2}
and the collision diameter
\sigma_{12} = \xi^{-1/6} \frac{\sigma_1 + \sigma_2}{2}.
The second order correction (Marrero and Mason 1972) is not considered yet. Therefore, the diffusion coefficient of a polar gas in a non-polar bath gas is equal to the diffusion coefficient of a non-polar gas in a polar bath gas. For diffusion of a polar gas in a polar bath gas, this formula cannot be used.
fit_B_data(B_df)
:-
Determination of Lennard-Jones parameters σ and ε by nonlinear regression from a data frame of second virial coefficient data. The data frame must contain the columns
T
for the temperature andvalue
for the viscosity in unitscm3/mol
. fit_viscosity_data(viscosity_df)
:-
Determination of Lennard-Jones parameters σ and ε by nonlinear regression from a data frame of viscosity data. The data frame must contain the columns
T
for the temperature andvalue
for the viscosity in unitsuPa.s
. fit_B_viscosity_data(B_df, viscosity_df, log=FALSE)
:-
Determination of Lennard-Jones parameters σ and ε by simultaneous nonlinear regression from a data frame of viscosity data and a data frame of second virial coefficient data. If
log=FALSE
the function to be minimized is given by (Bechtel et. al. 2020)\chi^2(\sigma, \varepsilon) = \sum{\left(\frac{\eta(\sigma, \varepsilon, T) - \eta_{\mathrm{exp}}}{\Delta\eta_{\mathrm{exp}}}\right)^2} + \sum{\left(\frac{B(\sigma, \varepsilon, T) - B_{\mathrm{exp}}}{\Delta B_{\mathrm{exp}}} \right)^2}
where
\Delta\eta_{\mathrm{exp}} = | \mathrm{max}(\eta_{\mathrm{exp}}) - \mathrm{min}(\eta_{\mathrm{exp}}) |
and
\Delta B_{\mathrm{exp}} = | \mathrm{max}( B_{\mathrm{exp}}) - \mathrm{min}(B_{\mathrm{exp}}) |
If
log=TRUE
the function to be minimized is given by (Zarkova and Hohm 2009)\chi^2(\sigma, \varepsilon) = \sum\left[\ln\left(\frac{\eta_{\mathrm{exp}}}{\eta(\sigma, \varepsilon, T)} \right) \right]^2 + \sum\left[\ln\left(\frac{|B_{\mathrm{exp}}|}{|B(\sigma, \varepsilon, T)|} \right) \right]^2
References
Bechtel S, Bayer B, Vidakovic-Koch T, Wiser A, Vogel H, Sundmacher K. Precise determination of LJ parameters and Eucken correction factors for a more accurate modeling of transport properties in gases. Heat and Mass Transfer 2020;56:2515-27. doi:10.1007/s00231-020-02871-4.
Brown NJ, Bastien LAJ, Price PN. Transport properties for combustion modeling. Progress in Energy and Combustion Science 2011;37:56582. doi:10.1016/j.pecs.2010.12.001.
Kim SU, Monroe CW. High-accuracy calculations of sixteen collision integrals for Lennard-Jones (12-6) gases and their interpolation to parameterize neon, argon, and krypton. Journal of Computational Physics 2014 273:358-73, doi:10.1016/j.jcp.2014.05.018.
Langenberg S, Carstens T, Hupperich D, Schweighoefer S, Schurath U. Technical note: Determination of binary gas-phase diffusion coefficients of unstable and adsorbing atmospheric trace gases at low temperature arrested flow and twin tube method. Atmospheric Chemistry and Physics 2020;20:366982. doi:10.5194/acp-20-3669-2020.
Marrero TR, Mason EA. Gaseous Diffusion Coefficients. J. Phys. Chem. Ref. Data 1972;1:3-118. doi:10.1063/1.3253094.
Vargas P, Munoz E, Rodriguez L. Second virial coefficient for the Lennard-Jones potential. Physica A: Statistical Mechanics and Its Applications 2001;290:92-100. doi:10.1016/s0378-4371(00)00362-9.
Zarkova L, Hohm U. Effective (n-6) Lennard-Jones Potentials with Temperature-Dependent Parameters Introduced for Accurate Calculation of Equilibrium and Transport Properties of Ethene, Propene, Butene, and Cyclopropane. Journal of Chemical & Engineering Data 2009;54:164855. doi:10.1021/je800733b.
Examples
# Second virial coefficient of methane at 300 K and standard pressure
CH4 <- Gas("methane")
print(CH4$B(T=300))
# Self-diffusion coefficient at 300 K
print(CH4$diffusion(T=300))
# create an instance of Gas for a molecule not listed in data_frame gas
Hg <- Gas("mercury")
# relative molecular mass
Hg$M <- 200.59
# mass of 1 molecule in kg
Hg$m <- Hg$M / pkg.env$Na / 1000
Hg$sigma <- 2.969
Hg$epsk <- 750
print(Hg$thermal_conductivity(T=700))
Binary gas phase diffusion coefficients of methane, ethane, propane and butane in helium and nitrogen; fluoromethane, difluoromethane and trifluoromethane in nitrogen measured by reverse-flow gas chromatography
Description
Reference data of binary diffusion coefficients for comparison with calculated diffusion coefficients. Diffusion coefficients were determined using a reversed-flow gas chromatography system.
Usage
data("binary_diffusion")
Format
A data frame with 91 observations on the following 6 variables.
doi
DOI of data source
bath_gas
Bath gas
helium
ornitrogen
gas
Diffusing species
C2H6
,C3H8
,C4H10
,CH2F2
,CH3F
,CH4
,CHF3
T
Temperature in
K
D
Diffusion coefficient in
cm2/s
U_D
Uncertainty of diffusion coefficient in
cm2/s
Details
Plot of experimental diffusion coefficient vs. temperature. (a) diffusion of nitrogen and
argon in helium. (b) diffusion of methane, ethane, propane and butane in helium. (c) diffusion of methane, ethane, propane and butane in nitrogen. (d) diffusion of fluoromethane,
difluoromethane and trifluoromethane in nitrogen. The solid lines are calculated using the Lennard-Jones model. For figures (a)-(c) the Lennard-Jones parameters are taken from
gas, for figure (d) the Lennard-Jones parameters are estimated by nonlinear regression using optim from experimental data.
The diffusion coefficient D
as function of pressure in a narrow temperature range close to the reference temperature T_0
is usually expressed as
(Langenberg et al. 2020)
D = D_0 \left(\frac{p_0}{p}\right)\left(\frac{T}{T_0}\right)^b
For the experimental data, the temperature coefficient b
is obtained from the fit. For the calculated diffusion coefficients, the temperature coefficient is calculated by
b = \left(\frac{\partial D}{\partial T}\right)_{T_0} \left(\frac{T_0}{D_0}\right).
The diffusion coefficients D_\mathrm{calc}
are calculated using Gas-class. The deviation is calculated by
\frac{D_\mathrm{exp} - D_\mathrm{calc}}{D_\mathrm{exp}}.
Gas | Bath gas | Experimental | Calculated | Deviation | ||
D_0 / [cm2/s] | b | D_0 / [cm2/s] | b |
|||
nitrogen | helium | 0.605(3) | 1.664(8) | 0.620 | 1.68 | -3% |
argon | helium | 0.630(2) | 1.665(6) | 0.640 | 1.68 | -2% |
methane | helium | 0.575(3) | 1.675(7) | 0.597 | 1.68 | -4% |
ethane | helium | 0.421(5) | 1.68(2) | 0.446 | 1.70 | -6% |
propane | helium | 0.341(7) | 1.67(2) | 0.361 | 1.70 | -7% |
butane | helium | 0.294(6) | 1.65(2) | 0.368 | 1.74 | -32% |
methane | nitrogen | 0.201(2) | 1.74(2) | 0.186 | 1.83 | 7% |
ethane | nitrogen | 0.136(2) | 1.70(2) | 0.123 | 1.87 | 7% |
propane | nitrogen | 0.106(2) | 1.72(3) | 0.094 | 1.88 | 7% |
butane | nitrogen | 0.090(1) | 1.72(2) | 0.084 | 1.97 | -8% |
The values in brackets indicate the uncertainties (0.95 confidence level) of the fit parameters. With the exception of the diffusion of butane in helium, the calculated diffusion coefficients well resemble the measured diffusion coefficients within an error limit of < 10%. For larger non spherical molecules like butane in helium more advanced combining rules need to be applied (Li et al. 2023).
The experimental data for the diffusion coefficients of fluoromethanes can in turn be used to estimate the Lennard-Jones parameters for the Van der Waals
interaction. The values for \sigma
obtained are smaller than \sigma
obtained from data of viscosity measurements (Shibasaki-Kitakawa et. al. 1995,
Clifford et al. 1979).
Gas | D_0 / [cm2/s] | b | Viscosity | Diffusion | ||
\sigma / [Ao] | \varepsilon/k / [K] | \sigma / [Ao] | \varepsilon/k / [K] |
|||
fluoromethane | 0.1576(7) | 1.784(8) | -- | -- | 3.5 | 174 |
difluoromethane | 0.133(2) | 1.76(2) | 4.9 | 204 | 3.9 | 153 |
trifluoromethane | 0.123(2) | 1.73(2) | 4.4 | 182 | 4.5 | 63 |
This is due to the fact that both molecules have a dipole moment. This is why the intermolecular interaction of polar molecules cannot be described in terms of the Lennard-Jones potential.
Source
McGivern WS, Manion JA. Extending reversed-flow chromatographic methods for the measurement of diffusion coefficients to higher temperatures. J. Chromatogr. A 2011;1218:8432-42. doi:10.1016/j.chroma.2011.09.035.
McGivern WS, Manion JA. Hydrocarbon binary diffusion coefficient measurements for use in combustion modeling. Combustion and Flame 2012;159:3021-6. doi:10.1016/j.combustflame.2012.04.015.
McGivern WS, Manion J. Binary Diffusion Coefficients for Methane and Fluoromethanes in Nitrogen. Journal of Chemical & Engineering Data 2021. doi:10.1021/acs.jced.1c00161.
References
Clifford AA, Gray P, Scott AC. Viscosities of CFCl3, CF3Cl, CHFCl2, CHF2Cl and CHF3 from 373 to 570 K. J. Chem. Soc., Faraday Trans. 1, 1979;75:1752. doi:10.1039/F19797501752
Langenberg S, Carstens T, Hupperich D, Schweighoefer S, Schurath U. Technical note: Determination of binary gas-phase diffusion coefficients of unstable and adsorbing atmospheric trace gases at low temperature arrested flow and twin tube method. Atmospheric Chemistry and Physics 2020;20:366982. doi:10.5194/acp-20-3669-2020.
Li Y, Gui Y, You X. On the binary diffusion coefficients of n-alkanes in He/N2. Combustion and Flame 2023;257:112795. doi:10.1016/j.combustflame.2023.112795.
Shibasaki-Kitakawa N, Takahashi M, Yokoyama C, Takahashi S. Gas Viscosity of Difluoromethane from 298.15 to 423.15 K and up to 10 MPa J. Chem. Eng. Data 1995; 40:900-902 doi:10.1021/je00020a036
Examples
# binary diffusion data of nitrogen in bath gas helium
nitrogen_in_helium <- subset(binary_diffusion,(gas=="nitrogen" & bath_gas=="helium"))
print(nitrogen_in_helium)
Coefficients to calculate collision integrals for the Lennard Jones (12-6) potential
Description
Coefficients for "CollisionIntegral"
, Table 1 from Kim and Monroe (2014).
Usage
data("coefficients_collisionintegral")
Format
l
order of collision integral
s
s
order of collision integral
l
A
model parameter
A
B1
model parameter
B_1
B2
model parameter
B_2
B3
model parameter
B_3 \times 10
B4
model parameter
B_4 \times 10
B5
model parameter
B_5 \times 100
B6
model parameter
B_6 \times 1000
C1
model parameter
C_1
C2
model parameter
C_2 \times 10
C3
model parameter
C_3 \times 10
C4
model parameter
C_4 \times 100
C5
model parameter
C_5 \times 1000
C6
model parameter
C_6 \times 10000
Source
Kim SU, Monroe CW. High-accuracy calculations of sixteen collision integrals for Lennard-Jones (12-6) gases and their interpolation to parameterize neon, argon, and krypton. Journal of Computational Physics 2014;273:35873. doi:10.1016/j.jcp.2014.05.018.
Viscosity and second virial coefficient of ethane
Description
Reference values for the second virial coefficient and viscosity of ethane from a intermolecular potential energy surface. The second virial coefficient was calculated semiclassically by means of the Mayer-sampling Monte Carlo technique, while the transport properties were obtained using the classical kinetic theory of polyatomic gases. The computed thermophysical property values are in excellent agreement with the best available experimental data and are recommended as reference values.
Usage
data("ethane_data")
Format
A data frame with 107 observations on the following 4 variables.
T
Temperature in
K
property
Type of property.
B
: classically calculated second virial coefficient,BQFH
: second virial coefficient calculated by a modification of the pair potential known as the quadratic Feynman-Hibbs (QFH) effective pair potential.viscosity
: gas phase viscosityunit
a factor with levels
cm3/mol
,uPa.s
value
numerical value of property. For the second virial coefficient the calculated data are supported by experimental data in the temperature range 220 - 623 K. For the viscosity the calculated data are supported by experimental data in the temperature range 90 - 675 K.
Details
(a) Viscosity of ethane. (b) second virial coefficient of ethane. The black solid curves are fits using viscosity data and second virial coefficient data
respectively. The red dotted curves are simultaneous fits against the viscosity and second virial coefficient data.
Type of fit | σ | ε/k |
vs. viscosity data | 4.38(2) | 235(6) |
vs. second virial coefficient data | 4.95(6) | 202(3) |
vs. viscosity and second virial coefficient data | 4.35 | 244 |
Source
Hellmann R. Reference Values for the Second Virial Coefficient and Three Dilute Gas Transport Properties of Ethane from a State-of-the-Art Intermolecular Potential Energy Surface. Journal of Chemical & Engineering Data 2018;63:470-81. doi:10.1021/acs.jced.7b01069.
Examples
c2h6 <- Gas("ethane")
# estimate LJ-coefficients from viscosity data
ethane_viscosity <- subset(ethane_data,
(property=="viscosity") & (T>=90) & (T<=675), select=c(T, value)
)
c2h6$zeta <- 0
ethane_viscosity$value <- 1E-6*ethane_viscosity$value
c2h6$fit_viscosity_data(ethane_viscosity)
print(c2h6$sigma)
print(c2h6$epsk)
# estimate LJ-coefficients from second virial coefficient
ethane_B <- subset(ethane_data, (property=="BQFH") & (T>= 220) & (T<=623))
ethane_B$value <- 1E-6*ethane_B$value
c2h6$fit_B_data(ethane_B)
print(c2h6$sigma)
print(c2h6$epsk)
Gas data
Description
Physical properties of some gas species.
Usage
data("gas")
Format
A data frame with properties of 43 gases on the following 12 variables.
formula
chemical formula
name
chemical name
CAS
Chemical abstracts registry number
group
M
dipole_moment
electric dipole moment in
Debye
polarizability
electric polarizability in
Ao
IE
ionization energy in
eV
Tc
critical temperature in
K
pc
critical pressure in
bar
rhoc
critical density in
mol/l
sigma
distance at which the intermolecular potential between the two particles is zero in
Ao
epsk
Well depth
\varepsilon/k
of the Lennard-Jones potential inK
DOI
Data source of Lennard-Jones parameters. If not specified otherwise, they are taken from Poling et al. (2004)
Details
Properties of simple gases for the calculation of transport properties. Not all properties are given for all molecules. However, correlations exists between the Lennard-Jones parameters and critical data as derived from numerous numerical simulations of the Lennard-Jones fluid (Stephan et al. 2019): ε and σ can be determined from critical temperature
T_c = (1.321 \pm 0.007) (\varepsilon/k)
critical density
\rho_c = (0.316 \pm 0.005) / \sigma^3
and critical pressure
p_c = (0.129 \pm 0.005) (\varepsilon/\sigma^3)
where k
is the Boltzmann constant.
Correlation of Lennard-Jones parameters of nonpolar gases with critical data: The solid lines are the
expectations from numerical simulations of the Lennard-Jones Fluid.
(a) critical temperature vs. well depth \varepsilon
. (b) critical density vs. 1/\sigma^3.
References
NIST Chemistry WebBook. NIST Standard Reference Database, vol. 69, 2023. doi:10.18434/T4D303.
NIST. Experimental Polarizabilites. in: III RDJ, editor. NIST Computational Chemistry Comparison and Benchmark Database, NIST; 2020. doi:10.18434/T47C7Z.
Poling BE, Prausnitz JM, OConnell JP. The Properties of Gases and Liquids. 5 ed. New York: McGraw-Hill; 2004.
Stephan S, Thol M, Vrabec J, Hasse H. Thermophysical Properties of the Lennard-Jones Fluid: Database and Data Assessment. Journal of Chemical Information and Modeling 2019;59:424865. doi:10.1021/acs.jcim.9b00620.