+ All documents
Home > Documents > A simple van’t Hoff law for calculating Langmuir constants in clathrate hydrates

A simple van’t Hoff law for calculating Langmuir constants in clathrate hydrates

Date post: 21-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
8
A simple van’t Hoff law for calculating Langmuir constants in clathrate hydrates Azzedine Lakhlifi a,, Pierre Richard Dahoo b , Sylvain Picaud a , Olivier Mousis a,c a Institut UTINAM-UMR 6213 CNRS, Université de Franche-Comté, Observatoire de Besançon, 41 bis avenue de l’Observatoire, BP 1615, 25010 Besançon Cedex, France b Université de Versailles-Saint-Quentin-en-Yvelines, Sorbonne Universités, Laboratoire Atmosphères Milieux Observations Spatiales, CNRS, UMR 8190, Observatoire de Versailles Saint-Quentin-en-Yvelines, 11 Bd d’Alembert, F-78820 Guyancourt, France c Aix Marseille Université, CNRS, LAM (Laboratoire d’Astrophysique de Marseille) UMR 7326, 13388 Marseille, France article info Article history: Received 10 October 2014 In final form 5 January 2015 Available online 13 January 2015 Keywords: Clathrate hydrates Langmuir constants Atom–atom interaction potential energy abstract This work gives a van’t Hoff law expression of Langmuir constants of different species for determining their occupancy in clathrate hydrates. First, a pairwise site–site interaction potential energy model is used to calculate the Langmuir constants in an otherwise anisotropic potential environment, as a function of temperature. The results are then fitted to a van’t Hoff law expression to give a set of parameters that can be used for calculating clathrates compositions. The van’t Hoff law’s parameters are given for eigh- teen gas species trapped in the small and large cavities of structure types I and II. The accuracy of this approach is based on a detailed comparison with available experimental and/or previously calculated data for ethane, cyclo-propane, methane and carbon dioxide clathrate hydrates. A comparison with the analytical cell method is also carried out to better understand the importance of asymmetry and possible limitations of the van’t Hoff temperature dependence. Ó 2015 Elsevier B.V. All rights reserved. 1. Introduction A clathrate is an ice-like crystalline solid consisting of water molecules forming a cage structure around smaller guest mole- cules under suitable conditions of low temperature and high pres- sure. On Earth, it is considered that clathrate hydrates are the most important reservoirs of fossil energy [1,2], and that favorable con- ditions for gas hydrate formation exist in about 25% of the earth’s land mass. Moreover, the thermodynamics conditions of pressure and temperature prevailing in the oceans are such that hydrates should easily be formed in about 90% of the ocean or sediments. The most common guest molecules in terrestrial clathrates are of organic aliphatic nature like methane, ethane, propane or butane, but other small inorganic molecules like nitrogen, carbon dioxide, and hydrogen sulfide can also be trapped in the cages of clathrates [3–9]. In the advent of global warming, these clathrates can enhance the temperature rise when the trapped species are released. Clathrate hydrates are also suspected to be extensively present on several planets, satellites and comets of the Solar Sys- tem. Planetologists are thus concerned with the possible clathrate impact on the distribution of the planet’s volatiles and on the mod- ification of their atmosphere’s compositions [10]. Hence, it is of great interest to correctly determine the amount of species poten- tially trapped in the cages of clathrates, i.e. the fractional occu- pancy of guest species under the thermodynamic conditions (pressure and temperature) prevailing in the regions where clath- rates might form. From a theoretical point of view, the thermodynamics of the formation or dissociation of clathrates is most often based on the model developed by van der Waals and Platteeuw [11] following the same hypotheses under which was developed the adsorption theory of Langmuir [12]. The Langmuir isotherms of adsorbed mol- ecules on a surface are determined from the calculation of the Langmuir constant, which is also the main parameter to be consid- ered in the determination of the amount of species trapped in the clathrate cages as a function of pressure and temperature. To calculate these Langmuir constants, most of the models are based on a molecular description of the guest-water interactions using a Lennard-Jones or Kihara potential form. The parameters of these potentials are usually empirically obtained from experi- mental data of phase equilibrium. Such models most often neglect interactions of the guest molecules with water beyond a few cages only and are therefore questionable [13–18]. Moreover, it is generally assumed that the environment of the cage in which a gas molecule is trapped in clathrates is of spherical symmetry. Whereas this assumption may be justified for molecules such as CH 4 or NH 3 , it is certainly not well-suited for http://dx.doi.org/10.1016/j.chemphys.2015.01.004 0301-0104/Ó 2015 Elsevier B.V. All rights reserved. Corresponding author. E-mail address: azzedine.lakhlifi@obs-besancon.fr (A. Lakhlifi). Chemical Physics 448 (2015) 53–60 Contents lists available at ScienceDirect Chemical Physics journal homepage: www.elsevier.com/locate/chemphys
Transcript

Chemical Physics 448 (2015) 53–60

Contents lists available at ScienceDirect

Chemical Physics

journal homepage: www.elsevier .com/locate /chemphys

A simple van’t Hoff law for calculating Langmuir constants in clathratehydrates

http://dx.doi.org/10.1016/j.chemphys.2015.01.0040301-0104/� 2015 Elsevier B.V. All rights reserved.

⇑ Corresponding author.E-mail address: [email protected] (A. Lakhlifi).

Azzedine Lakhlifi a,⇑, Pierre Richard Dahoo b, Sylvain Picaud a, Olivier Mousis a,c

a Institut UTINAM-UMR 6213 CNRS, Université de Franche-Comté, Observatoire de Besançon, 41 bis avenue de l’Observatoire, BP 1615, 25010 Besançon Cedex, Franceb Université de Versailles-Saint-Quentin-en-Yvelines, Sorbonne Universités, Laboratoire Atmosphères Milieux Observations Spatiales, CNRS, UMR 8190, Observatoire deVersailles Saint-Quentin-en-Yvelines, 11 Bd d’Alembert, F-78820 Guyancourt, Francec Aix Marseille Université, CNRS, LAM (Laboratoire d’Astrophysique de Marseille) UMR 7326, 13388 Marseille, France

a r t i c l e i n f o a b s t r a c t

Article history:Received 10 October 2014In final form 5 January 2015Available online 13 January 2015

Keywords:Clathrate hydratesLangmuir constantsAtom–atom interaction potential energy

This work gives a van’t Hoff law expression of Langmuir constants of different species for determiningtheir occupancy in clathrate hydrates. First, a pairwise site–site interaction potential energy model isused to calculate the Langmuir constants in an otherwise anisotropic potential environment, as a functionof temperature. The results are then fitted to a van’t Hoff law expression to give a set of parameters thatcan be used for calculating clathrates compositions. The van’t Hoff law’s parameters are given for eigh-teen gas species trapped in the small and large cavities of structure types I and II. The accuracy of thisapproach is based on a detailed comparison with available experimental and/or previously calculateddata for ethane, cyclo-propane, methane and carbon dioxide clathrate hydrates. A comparison with theanalytical cell method is also carried out to better understand the importance of asymmetry and possiblelimitations of the van’t Hoff temperature dependence.

� 2015 Elsevier B.V. All rights reserved.

1. Introduction

A clathrate is an ice-like crystalline solid consisting of watermolecules forming a cage structure around smaller guest mole-cules under suitable conditions of low temperature and high pres-sure. On Earth, it is considered that clathrate hydrates are the mostimportant reservoirs of fossil energy [1,2], and that favorable con-ditions for gas hydrate formation exist in about 25% of the earth’sland mass. Moreover, the thermodynamics conditions of pressureand temperature prevailing in the oceans are such that hydratesshould easily be formed in about 90% of the ocean or sediments.The most common guest molecules in terrestrial clathrates are oforganic aliphatic nature like methane, ethane, propane or butane,but other small inorganic molecules like nitrogen, carbon dioxide,and hydrogen sulfide can also be trapped in the cages of clathrates[3–9]. In the advent of global warming, these clathrates canenhance the temperature rise when the trapped species arereleased. Clathrate hydrates are also suspected to be extensivelypresent on several planets, satellites and comets of the Solar Sys-tem. Planetologists are thus concerned with the possible clathrateimpact on the distribution of the planet’s volatiles and on the mod-ification of their atmosphere’s compositions [10]. Hence, it is of

great interest to correctly determine the amount of species poten-tially trapped in the cages of clathrates, i.e. the fractional occu-pancy of guest species under the thermodynamic conditions(pressure and temperature) prevailing in the regions where clath-rates might form.

From a theoretical point of view, the thermodynamics of theformation or dissociation of clathrates is most often based on themodel developed by van der Waals and Platteeuw [11] followingthe same hypotheses under which was developed the adsorptiontheory of Langmuir [12]. The Langmuir isotherms of adsorbed mol-ecules on a surface are determined from the calculation of theLangmuir constant, which is also the main parameter to be consid-ered in the determination of the amount of species trapped in theclathrate cages as a function of pressure and temperature.

To calculate these Langmuir constants, most of the models arebased on a molecular description of the guest-water interactionsusing a Lennard-Jones or Kihara potential form. The parametersof these potentials are usually empirically obtained from experi-mental data of phase equilibrium. Such models most often neglectinteractions of the guest molecules with water beyond a few cagesonly and are therefore questionable [13–18].

Moreover, it is generally assumed that the environment of thecage in which a gas molecule is trapped in clathrates is of sphericalsymmetry. Whereas this assumption may be justified formolecules such as CH4 or NH3, it is certainly not well-suited for

54 A. Lakhlifi et al. / Chemical Physics 448 (2015) 53–60

molecules such as CO2 and N2O or SO2 which are of cylindrical oroblate symmetry and for which the hypothesis of a free rotationat the center of a spherical cage is no longer valid.

In the case of ethane and cyclo-propane clathrate hydrates, ananalytical method based on the spherical cell model has been usedto extract spherically averaged intermolecular potentials fromexperimental data on the temperature dependence of the Langmuirconstant by Bazant and Trout [19]. The cell potential method hasthen been extended to other molecules like methane, propane, iso-butane and small chlorofluorocarbon molecules, showing a goodaccuracy in reproducing both dissociation pressures and phase dia-grams of hydrocarbon mixtures [7,8]. This method thus appears as avery interesting and quite simple approach for predictions of clath-rate hydrate equilibrium properties. However, it should be men-tioned that, for some cases, the central-well potential does notsatisfactorily reproduce the shape of the guest- water potential atthe center of the clathrate cages. Although Anderson et al. [8]claimed that this has no influence on the accuracy of the Langmuirconstants calculated for Argon, the influence of this failure shouldbe carefully analyzed each time it is evidenced. For other guests inclathrate hydrates, in particular, argon, hydrogen, nitrogen, meth-ane, ethane, propane, cyclo-propane, and carbon dioxide, otherworkers [3–6,9] have explicitly taken into account the angle-dependence of the guest-water intermolecular potential in anatom–atom or site–site description to calculate the correspondingLangmuir constants. However, in most of these studies the waterand guest molecules were simply described as one, two or threeinteraction sites for the Lennard-Jones or Kihara potential contribu-tions although more than three sites are involved.

This work aims at providing a van’t Hoff law expression [19] ofthe Langmuir constant for single guest molecules incorporated inclathrate hydrates as a function of the temperature by improvingthe potential model used in the determination of the Langmuirconstants as discussed in reference [10], that is by using anatom–atom and site–site potential and by considering explicitlythe effect of water molecules beyond the trapping cage and theresulting anisotropic environment. In the present work, the Lang-muir constant is determined by taking into account all the externaldegrees of freedom of the guest molecules, i:e. the center of mass(c.m.) translational motion and the orientational motion in a truecrystallographic clathrate lattice, not necessarily of spherical sym-metry as it is often assumed when using the van der Waals andPlatteeuw model [11]. In the following, we recall in Section 2 themodel used for the calculations of the Langmuir constants thatare necessary to determine the fractional occupancy of guest spe-cies in clathrates. Then, in Section 3, the geometry and interactionpotential considered here are described. Finally, in Section 4, thecoefficients for a simple van’t Hoff expression of the Langmuir con-stants are given for a large set of guest molecules. The Langmuirconstants calculated using this model are compared with availableexperimental and/or calculated data, i.e., for ethane, cyclo-propane,methane and carbon dioxide guest molecules.

2. Statistical thermodynamic approach

In contrast to natural ice which solidifies in the hexagonalstructure, clathrate hydrates form, as water crystallizes, in thecubic system in several different structures which are character-ized by specific cages of different sizes. The two most commontypes are ‘‘structure I’’ and ‘‘structure II’’. In structure I, the unit cellis made of 46 water molecules forming 2 small (12 pentagonalfaces 512) and 6 large (12 pentagonal and 2 hexagonal faces51262) cages, while in structure II, the unit cell is made up of 136water molecules forming 16 small (12 pentagonal faces 512) and8 large (12 pentagonal and 4 hexagonal faces 51264) cages [2].

Calculations of the relative abundances of guest species incor-porated in a clathrate lattice structure of type I (sI) or II (sII) atgiven temperature–pressure conditions can be performed usingclassical statistical mechanics which allows the macroscopic ther-modynamic properties of the clathrates to be determined from theinteraction energies between the guest species and the clathratewater molecules.

In 1959 van der Waals and Platteeuw [11] developed a model ofclathrate’s formation in which the trapping of guest molecules innano-cages was considered to be a generalized case of the three-dimensional ideal localized adsorption.

Their model is based on the following hypotheses:

1. The contribution of the host molecules to the free energy isindependent of the occupational mode in the cages. This impliesin particular that the guest species do not distort the trappingcage.

2. The encaged molecules are localized in the cages, each of whichcan never hold more than one guest.

3. The mutual interaction of the guest molecules is neglected, i.e.,the partition function for the motion of a guest molecule in itscage is independent of the other guests.

4. Classical statistics is valid, i.e., quantum effects are negligible.

From the configuration partition function and the thermody-namic equilibrium condition on the chemical potentials of theguest and host molecules in coexisting phases in clathrate [20],the fractional occupancy of a guest molecule K in a given ‘‘structuretype-cage size’’ t (t = structure-type I or II-small or large cage) canbe written as:

yK;t ¼CK;tf K

1þP

JCJ;t f J; ð1Þ

where the sum in the denominator includes all the species presentin the initial gas phase, CK;t is the Langmuir constant of species K inthe structure type-cage size t, and f K is the fugacity of the species Kwhich depends on the total pressure P of the initial gas phase andon the temperature T.

The Langmuir constant in Eq. (1) depends on the temperature Tand on the strength of the interaction energy between the guestspecies K and the water molecules in the cage. It is expressed as:

CK;t ¼1

kBT

Zexp �VK;tðr;XÞ

kBT

� �drdX: ð2Þ

In this equation VK;tðr;XÞ is the interaction potential energy experi-enced by the guest molecule for a given position vector r of its cen-ter of mass with respect to the cage center and its orientationalvector X, and kB is the Boltzmann constant. The integral value mustbe calculated for all external degrees of freedom of the guest mole-cule inside the structure type-cage size t.

To evaluate the Langmuir constant CK;t , two additional assump-tions are often made, namely: (i) the symmetry of the guest mole-cule’s environment is considered to be spherical and (ii) the guestmolecule can freely rotate in the corresponding spherical cage(spherical cell potential approximation), in accordance with theLennard-Jones and Devonshire [21,22] theory applied to liquids.

Then, the Langmuir constant can be cast as:

CK;t ¼4pkBT

Z Rc

0exp �VK;tðrÞ

kBT

� �r2dr; ð3Þ

where Rc is the radius of the spherical cage and VK;tðrÞ is the spher-ically averaged potential energy between the guest molecule andthe clathrate water molecules.

Note that Eq. (3) is commonly used by planetologists and as aresult, it may introduce significant inaccuracies in the evaluation

A. Lakhlifi et al. / Chemical Physics 448 (2015) 53–60 55

of the relative abundances of gas species in clathrate hydrateswhen the assumed spherical symmetry of the interaction potentialis questionable.

In contrast, the Langmuir constant for enclathrated single guestmolecules is determined here by taking into account all the exter-nal degrees of freedom of the enclathrated and water molecules,i.e. the center of mass (c.m.) translational motion and the orienta-tional motion in a real crystallographic clathrate lattice, not neces-sarily of spherical symmetry. As consequence, Eq. (3) is no longervalid in our approach and should be replaced by a more generalform, as explained below.

3. Potential energy model

Let us consider a non-vibrating single gas species (atom or mol-ecule) trapped in a small or large cage of a clathrate structure oftype sI or sII. The positions of the hydrogen and oxygen atomsare given by their coordinates in a reference frame fixed to theclathrate lattice, while those of the guest molecule are usuallygiven in a frame tied to the molecule with the z axis coincidingwith the rotational axis of highest symmetry as shown for the trap-ping of a guest molecule (CH4 for instance) in the small cage of theclathrate structure type I, in Fig. 1.

The interaction potential energy VK;t between the guest mole-cule K and the surrounding water molecules of the structuretype-cage size t, considered as a rigid clathrate crystal, is modeledas a sum of a 12–6 Lennard-Jones (LJ) pairwise atom–atom poten-tial characterizing the repulsion-dispersion contributions and anelectrostatic part due to charge-charge interactions between the

Fig. 1. Trapping geometry of a guest molecule (CH4 for instance) in a small cage ofthe structure type I. (O,X,Y,Z) represents the absolute frame tied to the clathratelattice.

Table 1Pure Lennard-Jones potential parametersa used in our calculations.

H–H O–O C–C N–N

�i (K) 8.59 57.41 42.88 37.29ri (Å) 2.810 3.030 3.210 3.310

a Obtained from Refs. [23–25].

charges in the guest molecule and those in the water moleculesof the clathrate system. It is expressed as:

VK;t ¼XNC

k¼1

XNW

j¼1

XNK

i¼1

4�ijrij

rijk

�� �� !12

� rij

rijk

�� �� !6

8<:

9=;þ 1

4p�0

qiqj

rijk

�� ��24

35;ð4Þ

where rijk is the distance vector between the ith site, with electriccharge qi, of the guest molecule (NK sites) and the jth site, with elec-tric charge qj, of the kth water molecule (NW sites) of the clathratematrix (containing NC water molecules); �ij and rij are the mixed LJpotential parameters, using the Lorentz–Berthelot combinationrules as:

�ij ¼ffiffiffiffiffiffiffiffi�i�j

pand rij ¼

ri þ rj

2; ð5Þ

where �i and ri are the Lennard-Jones parameters for the i� i inter-acting atomic pair, which are taken from the literature [23–25] (seeTable 1). The effective electric charges for the water moleculesqH ¼ þ0:4238 e; qO ¼ �0:8476 e are from reference [26] and thosefor the guest species are from references [26,25,27–29].

Calculations of the Langmuir constant CK;t from Eq. (2) requirean explicit determination of the external degrees of freedom ofthe guest molecule, that is, its center of mass (c.m.) position vectorr and its orientation vector X = (u; h;v). Therefore, we define anabsolute frame (O,X,Y,Z) connected to the clathrate matrix.

Fig. 2 gives the geometrical characteristics of the guest-waterinteracting molecules, and a description of the internal positionsof the sites in the guest molecule with respect to its frame(G,x,y,z) and its external orientational and translational degreesof freedom with respect to the absolute frame (O,X,Y,Z).

Then, the distance vector rijk in Eq. (4) can be expressed in termsof the position vector r of the c.m. of the guest molecule and rjk ofthe jth site of the kth water molecule with respect to the absoluteframe (O,X,Y,Z), and of the position vector gi of the ith site of theguest molecule with respect to its associated frame (G,x,y,z) (seeFig. 2), as:

rijk ¼ rjk � r� gi: ð6Þ

It should be noted that the explicit dependence upon the angu-lar degrees of freedom u; h, and v of the guest molecule requiresthe determination of the site position vectors {g} with respect tothe absolute frame.

Then, by assuming (i) that the guest species and water mole-cules are rigid, and (ii) that the clathrate lattice is undistorted (sta-tic lattice), the Langmuir constant is calculated from a six-dimensional configurational integral written as:

CK;t ¼1

kBT

Zexp �VK;tðx; y; z;u; h;vÞ

kBT

� �dxdydzdu sin hdhdv: ð7Þ

Of course, it should be noted that (i) for atomic species, thereare no orientational variables, and (ii) for linear molecules thereis no spinning variable v.

S–S Ne–Ne Ar–Ar Kr–Kr Xe–Xe

73.79 43.16 126.90 179.85 226.033.390 2.764 3.405 3.650 3.970

Fig. 2. Geometrical characteristics of a guest molecule (CH4 for instance) interact-ing with a water molecule of the clathrate matrix. (O,X,Y,Z) and (G,x,y,z) representthe absolute frame tied to the clathrate matrix and the molecular frame,respectively.

56 A. Lakhlifi et al. / Chemical Physics 448 (2015) 53–60

4. Results and discussion

A realistic interaction potential energy between the guest spe-cies and its surrounding is the prerequisite for an accurate deter-mination of both cage occupancy and phase equilibrium of gashydrates from the calculation of the corresponding Langmuir con-stants. In the present approach, we take into account the internalgeometry of the water and guest molecules, on the one hand,and the geometry of the cages and the size of the clathrate latticewhich depend on the guest occupying the cages, on the other hand.It is thus obvious from Eq. (7) that, in contrast to the simplifiedcentral-well potential approach [8], any analytical integrationwould be intractable here due to the asymmetries of the host lat-tice cages and of the guest molecule itself.

As already described above, we herein use a pairwise atom–atom Lennard-Jones and a site–site electrostatic potentials for cal-culating the interaction of single guest molecules with clathratematrices containing 4 � 4 � 4 unit cells (up to 2944 water mole-cules) for the structure sI, and 3 � 3 � 3 unit cells (up to 3672water molecules) for the structure sII. Indeed, it has been previ-ously shown that these number of water molecules are largeenough to ensure a good convergence of the corresponding calcu-lations [30]. Moreover, in order to determine the Langmuir con-stant and its temperature dependence for a guest molecule K in astructure type-cage size t of clathrate, the potential energy surfaceVK;tðx; y; z;u; h;vÞ in Eq. (7) is first numerically computed for all thecenter of mass positions and orientations of the molecule withrespect to the absolute frame as defined, and then the numerical

Table 2Parameters AK ;t (Pa�1) and BK;t (K) for the van’t Hoff expression of the Langmuir constant

Structure type-cage size sI-small cage sI-larg

Guest species K AK;t AK;t

BK;t BK;t

Ne 1.0308 �10�9 2.71941187.948 1015.8

Ar 1.5210 �10�10 7.78292961.545 2521.7

Kr 0.5985 �10�10 3.87993885.383 3454.0

Xe 1.9389 �10�11 13.8504547.654 4572.7

integration is performed. Note that the minimum of the potentialenergy Vmin

K;t and the associated molecular position and orientationcan also be determined in this way.

4.1. Calculation and fit of the Langmuir constants

A clathrate is characterized by its degree of occupancy i:e. thefraction of cage sites that is occupied by a guest molecule, whichcan be calculated from the knowledge of the Langmuir constantCK;t . Because such constants are not necessarily extracted fromexperiments in a large range of temperatures, we propose here toexpress them as a function of temperature through a simple van’tHoff law as already proposed by Bazant and Trout [19] and Ander-son et al. [8]:

CK;tðTÞ ¼ AK;t exp ðBK;t=TÞ; ð8Þ

in which A (Pa�1) and B (K) are constant values fitted to valuesderived from an atomic description of the clathrate systems. Thissimple van’t Hoff expression could thus be readily used, forinstance, by planetologist in the large range of temperaturesencountered in the Solar System.

In this work Langmuir constants have been calculated for eigh-teen species, in the temperature range between 50 K and 300 K, asgiven in Tables 2–4 for rare gas atoms, linear molecules and non-linear molecules, respectively.

For comparison, it must be noted that the Langmuir constantvalues depend not only on the minimum potential energy Vmin

K;t ,but also on the whole potential energy surface experienced bythe guest species.

In the case of argon, methane, ethane, cyclo-propane, propaneand iso-butane, the values of Vmin

K;t calculated here are consistentwith the well depth of the cell potentials used previously by Ander-son et al. [7,8], as shown in Table 5. However, it should be notedthat the different potential energy surfaces show that, except forAr, the five other molecules cannot freely rotate inside the clath-rate cage. Even CH4 which was expected to rotate freely, couldbe constrained in its rotational motion at low temperatures by abarrier height of about 60 meV. Moreover, as far as translationalmotions are concerned, the position of the minimum energy forthe guest molecules is not necessarily at the center of the cage,in contrast to usual assumptions. As a consequence, the valuesreported in Table 5 shows that our potential energy minimum val-ues are always equal or lower than those given in reference [8],except for cyclo-C3H6 in sII which is known to be stable only insI [2].

Let us finally note that Klauda and Sandler [5,31], and Sun andDuan [9] used ab initio approaches to fit the parameters of site–siteintermolecular potentials to calculate Langmuir constants for dif-ferent hydrates. The former authors modeled CH4, C2H6, C3H8, N2,H2, and CO2 and their mixtures as guests but only the oxygen atom

for simple guest clathrate hydrates.

e cage sII-small cage sII-large cage

AK;t AK;t

BK;t BK;t

�10�9 0.8154 �10�9 5.6151 �10�9

62 1233.898 898.062

�10�10 1.0456 �10�10 2.4531 �10�10

58 2977.025 2195.964

�10�10 0.4004 �10�10 16.2508 �10�10

28 3789.957 3021.690

9 �10�11 1.2913 �10�11 83.2357 �10�11

32 4085.506 4103.672

Table 3Parameters AK;t (Pa�1) and BK;t (K) for the van’t Hoff expression of the Langmuir constant for simple guest clathrate hydrates.

Structure type-cage size sI-small cage sI-large cage sII-small cage sII-large cage

Guest species K AK;t AK;t AK;t AK;t

BK;t BK;t BK;t BK;t

H2 4.7301 �10�9 16.0695 �10�9 5.5295 �10�9 64.0074 �10�9

1265.757 1515.721 1203.620 873.259

O2 1.3153 �10�9 7.4606 �10�9 0.8004 �10�9 25.8421 �10�9

2917.693 2558.746 3044.536 2238.052

N2 3.9496 �10�10 25.6897 �10�10 4.8836 �10�10 201.3238 �10�10

2869.400 2680.372 2679.423 2226.480

CO 2.5937 �10�10 16.5331 �10�10 4.5198 �10�10 229.8631 �10�10

3518.021 3075.059 3088.930 2275.803

CO2 7.7765 �10�12 520.5579 �10�12 7.9970 �10�12 6907.0012 �10�12

2976.629 4674.690 2277.757 3370.363

HCN 7.6653 �10�12 131.1607 �10�12 13.9141 �10�12 8224.5133 �10�12

4085.369 4328.556 2593.031 2640.868

C2H2 0.9702 �10�12 221.9622 �10�12 0.2439 �10�12 3909.8421 �10�12

735.205 3076.356 321.114 2837.467

Table 4Parameters AK;t (Pa�1) and BK;t (K) for the van’t Hoff expression of the Langmuir constant for simple guest clathrate hydrates.

Structure type-cage size sI-small cage sI-large cage sII-small cage sII-large cage

Guest species K AK;t AK;t AK;t AK;t

BK;t BK;t BK;t BK;t

H2S 2.3444 �10�10 7.2080 �10�10 3.6415 �10�10 758.3575 �10�10

4463.910 4073.045 3073.324 2495.937

SO2 1.2311 �10�12 75.0641 �10�12 6.1515 �10�12 17926.7530 �10�12

4374.084 6272.810 1548.504 4139.948

CH4 8.3453 �10�10 116.6313 �10�10 5.4792 �10�10 829.8039 �10�10

2901.747 2959.901 2546.660 2629.194

C2H6a – 3.5164 �10�11 – 727.2717 �10�11

– 4226.997 – 4440.484

cyc-C3H6a – 1.4881 �10�11 – 402.4295 �10�11

– 4781.938 – 5161.620

C3H8a – 5.5707 �10�13 – 597.9850 �10�13

– 3537.025 – 7118.782

iso-C4H10a – 2.7970 �10�14 – 208.3210 �10�14

– 1598.004 – 7103.169

a When the potential energy surfaces are positive, the Langmuir constants are negligible and the parameters AK;t and BK;t are not given.

Table 5Minimum potential energy Vmin

K;t (meV). For ethane, cycle-propane, propane and iso-butane, there are no trapping in the small cage because of the repulsive interactionpotential energy.

Cage size Anderson [8] This work

A. Lakhlifi et al. / Chemical Physics 448 (2015) 53–60 57

was considered as the interacting site for water. In a similar way,for the alkanes, only carbon atoms was considered as interactingsites. Electrostatic contributions were taken into account for CO2

and C3H8, only. In the present calculations, each atom of waterand the guest molecules have been considered as interaction sites.

Structure sIArgon small �214.5 �256.7

large �217.6 �221.8Methane small �244.8 �266.1

large �245.7 �262.9Ethane large �353.5 �384.5Cyclo-propane large �419.6 �432.9

Structure sIIArgon small �214.4 �258.7

large �170.3 �184.9Methane small �239.1 �238.1

large �215.2 �229.3Ethane large �377.9 �392.4Cyclo-propane large �510.2 �458.4Propane large �507.1 �635.5iso-butane large �553.7 �628.1

4.2. Comparison with available Langmuir constants

4.2.1. Ethane and cyclo-propane simple hydratesIn Tables 6 and 7, the Langmuir constants calculated here are

compared to experimental values obtained from the dissociationpressure data by Sparks and Tester [3] and to those calculated fromthe work of Anderson et al. [8] for temperatures ranging between200 K and 290 K and between 240 K and 290 K, for ethane andcyclo-propane trapped in large cages of structure sI, respectively.

It can be seen that the values calculated here are in good agree-ment with those extracted from experimental data. The ratiosCcal=Cexp vary between 0.5 and 1.2 for ethane, and between 0.7and 1.1 for cyclo-propane, depending on temperature. Note that

Table 6Calculated and experimental Langmuir constants (Pa�1) for ethane trapped in large cage of structure sI. Experimental values are obtained from reference [3]. The values in thethird column are from [32].

Temperature (K) Ccal: this work Cexp: Ccal: Chen Ccal: Anderson Ccal: spherical model

200 5.3 �10�2 9.9 �10�2 3.9 �10�2 3.1 �10�2 14.7 �10�2

210 2.0 �10�2 3.2 �10�2 1.5 �10�2 1.2 �10�2 5.2 �10�2

220 7.8 �10�3 12.0 �10�3 6.2 �10�3 4.8 �10�3 20.2 �10�3

230 3.4 �10�3 4.6 �10�3 2.8 �10�3 2.1 �10�3 8.5 �10�3

240 1.6 �10�3 1.9 �10�3 1.4 �10�3 1.0 �10�3 3.9 �10�3

250 7.7 �10�4 8.4 �10�4 6.9 �10�4 5.1 �10�4 18.6 �10�4

260 4.0 �10�4 4.1 �10�4 3.7 �10�4 2.7 �10�4 9.5 �10�4

270 2.2 �10�4 2.0 �10�4 2.1 �10�4 1.5 �10�4 5.1 �10�4

280 1.3 �10�4 1.1 �10�4 1.2 �10�4 0.9 �10�4 2.9 �10�4

290 7.5 �10�5 6.1 �10�5 7.3 �10�5 5.3 �10�5 16.7 �10�5

Table 7Calculated and experimental Langmuir constants (Pa�1) for cyclo-propane trapped inlarge cage of the sI structure. Experimental values are obtained from reference [3].

Temperature(K)

Ccal: thiswork

Cexp: Ccal:

AndersonCcal: Sphericalmodel

240 6.7 �10�3 9.5 �10�3 9.5 �10�3 0.12 �10�3

250 3.0 �10�3 3.8 �10�3 4.2 �10�3 0.06 �10�3

260 1.5 �10�3 1.7 �10�3 2.0 �10�3 0.04 �10�3

270 7.3 �10�4 7.6 �10�4 9.9 �10�4 0.23 �10�4

280 3.9 �10�4 3.8 �10�4 5.2 �10�4 0.15 �10�4

290 2.2 �10�4 2.0 �10�4 2.9 �10�4 0.10 �10�4

Fig. 3. Langmuir constants calculated for (a) ethane and (b) cyclo-propaneclathrates from the van’t Hoff parameters (blue curve: this work, green curve:Anderson et al. [8]) compared to experimental values (black squares: Sparks andTester [3]) as a function of temperature. (For interpretation of the references tocolour in this figure caption, the reader is referred to the web version of this article.)

58 A. Lakhlifi et al. / Chemical Physics 448 (2015) 53–60

the same ratios calculated within the same temperature range forthe large cages of structure sII lead to values ranging from 320 to540 for ethane, and from 920 to 1100 for cyclo-propane. Theseresults are consistent with the fact that ethane and cyclo-propaneform sI-large cage clathrate hydrates only. Moreover, Tables 6 and7 show that the best agreement between the values calculated hereand those extracted from experiments is found when temperatureis higher than 260 K. Below this temperature, the calculated valuesare systematically lower than the experimental ones.

In Table 6 we also give the Langmuir constants in the same tem-perature range for ethane obtained by Chen and coworkers [32]from a fit of experimental data [2] to the empirical model of Parrishand Prausnitz [33].

To extract a quantitative value for our comparison, we use thefollowing least square deviation parameter r defined as:

r ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiPNi¼1ðCcal � CYÞ2

N � 1

s; ð9Þ

where Ccal are the values calculated at different temperatures in thiswork and CY are the values obtained experimentally [3] or from cal-culations [32]. N is the number of temperatures considered (N ¼ 10in Table 6). We find values of r equal to 1:6� 10�2 Pa�1 and5:0� 10�3 Pa�1, respectively. Such calculations with calculated val-ues of reference [32] lead to r ¼ 2:1� 10�2 Pa�1, that is slightlyhigher than the r value calculated with the values issued fromthe present work. Moreover, note that in our calculations, the min-imum of the potential energy corresponds to ethane’s center ofmass located at 0.12 Å from the center of the large cage of structuresI, i.e., a value that is consistent with observations from single crys-tal X-ray diffraction studies that place the center of mass at 0.17 Åfrom the center of the large cage [34].

As for cyclo-propane (values given in Table 7 (N ¼ 6)), wecalculate a value of r equal to 1:3� 10�3 Pa�1.

To summarize, when compared to experimental results, thevalues of the Langmuir constants calculated here show a better

agreement with experimental data than those given by Chen[32], although they are of the same order of magnitude. In addition,Tables 6 and 7 show that values obtained from a simple sphericalmodel as often used in the literature are certainly not correct.

Therefore, the comparison with values issued from experimen-tal work provides firm and sufficient grounds to validate theresults obtained here for ethane and cycle-propane.

Table 8Parameters AK;t (Pa�1) and BK;t (K) for the van’t Hoff expression of the Langmuir constant for simple guest clathrate hydrates.

Structure type-cage size sI-small cage sI-large cage sII-small cage sII-large cage

Guest species K AK;t AK;t AK;t AK;t

BK;t BK;t BK;t BK;t

Ar (this work) 1.5210 �10�10 7.7829 �10�10 1.0456 �10�10 2.4531 �10�10

2961.545 2521.758 2977.025 2195.964

Ar (Ref. [8]) 1.7032 �10�10 6.3837 �10�10 1.6498 �10�10 21.4355 �10�10

2489.155 2245.461 2487.994 1976.238

CH4 (this work) 8.3453 �10�10 116.6313 �10�10 5.4792 �10�10 829.8039 �10�10

2901.747 2959.901 2546.660 2629.194

CH4 (Ref. [8]) 0.8262 �10�10 3.5985 �10�10 0.8267 �10�10 16.5647 �10�10

2840.769 2851.213 2774.624 2497.278

C2H6 (this work) – 3.5164 �10�11 – 727.2717 �10�11

– 4226.997 – 4440.484

C2H6 (Ref. [8]) – 3.9294 �10�11 – 22.1561 �10�11

– 4102.173 – 4385.322

cyc-C3H6 (this work) – 1.4881 �10�11 – 402.4295 �10�11

– 4781.938 – 5161.620

cyc-C3H6 (Ref. [8]) – 1.4635 �10�11 – 1.9609 �10�11

– 4869.227 – 5920.590

C3H8 (this work) – 5.5707 �10�13 – 597.9850 �10�13

– 3537.025 – 7118.782

C3H8 (Ref. [8]) – – – 86.7165 �10�13

– – – 5884.617

iso-C4H10 (this work) – 2.7970 �10�14 – 208.3210 �10�14

– 1598.004 – 7103.169

iso-C4H10 (Ref. [8]) – – – 182.328 �10�14

– – – 6425.384

A. Lakhlifi et al. / Chemical Physics 448 (2015) 53–60 59

It is also interesting to compare the results obtained here for theLangmuir Constants to those calculated from the parameters givenby Anderson et al. [8] that are based on the cell potential method.As an example, Fig. 3 shows the Langmuir constants calculated forethane and cyclo-propane (values given in Tables 6 and 7) mole-cules in clathrates of sI structure, in the temperature range forwhich the parameters of the cell potential method have been fit-ted. In addition to the curves corresponding to the calculations ofAnderson et al. and those issued from the present approach, datafrom experimental measurements are also given for comparison.For ethane, the values of the Langmuir constants calculated hereshow a better agreement with experimental data than those issuedfrom the cell potential method, especially at lowest temperatures(Fig. 3(a)). The reverse situation is however obtained for cyclo-propane (Fig. 3(b)), for which the cell potential method appearsin better agreement with experimental data than the values com-ing from the present approach. This could indicate that, for largeprolate molecules such as ethane, taking into account the atomisticgeometry of the guest molecule gives better results than thesimplified cell potential approach, whereas in the case of cyclo-propane which is rather oblate, analytical cell model could beapplied. Of course, new experimental work would be muchwelcome to validate this conclusion. Moreover, it should beemphasized that the parameters of the cell potential have beenfitted to experimental data in a given range of temperatures andthus, the corresponding Langmuir constants are supposed to beused in the same range, only. In contrast, Langmuir constants cal-culated here are based on potential energy calculations that do notdepend on temperature. As a consequence, they might be usedirrespective of the temperature. On the other hand, the cell poten-tial method takes implicitly into account the internal dynamics ofthe clathrates at given temperature through the fit to experimental

results, which is not accounted for in our approach. Honestly, it isthus quite difficult to say that one method is better than the otherone, because both are based on different assumptions.

The cell potential method has also been used to calculate theLangmuir constants for methane and argon, on the basis of site–site ab initio potentials, aiming at removing any fitting parametersin the calculations [8]. Indeed such approach seems a priori betterthan a fitting to experimental data. Nevertheless, it also suffersfrom some weakness due to (i) the ab initio potential do not dependon temperature (the clathrate is considered as a rigid body whencalculating the guest-water interactions) and (ii) it can dependon the level of accuracy chosen for the quantum calculations.

Anyway, the values of van’t Hoff parameters given by Andersonet al. [8] for a series of molecules are compared to our own valuesin Table 8. These two sets of parameters do not show significantdifferences, thus indicating that both cell potential and atom–atomapproaches can certainly be used with the same level of confidenceand accuracy.

4.2.2. Methane and carbon dioxideWhen considering CH4, Table 5 shows that the minimum of the

potential energies calculated in this work compare qualitativelywell with Anderson’s values [8]. In the [150–300] K temperaturerange, the ratio between the Langmuir constants calculated in thiswork from the potential given in Eq. (4) and that determined byAnderson from a fit to experimental data is between 13 and 16,and 1 and 4 for CH4 located in small cages of structures sI andsII, respectively.

In the case of CO2, Uchida [35] determined the cage occupancyand the hydration number by different methods: Phase equilibria,Raman spectra . . .and used the expression given by Parrish and

60 A. Lakhlifi et al. / Chemical Physics 448 (2015) 53–60

Prausnitz [33] based on a Lennard-Jones and Devonshire model tocalculate the Langmuir constant:

CK;tðTÞ ¼AK;t

TexpðBK;t=TÞ; ð10Þ

where A (K.Pa�1) and B (K) are adjustable parameters.In the [260–300] K temperature range, the ratio between the

Langmuir constants calculated in the present work and that deter-mined by Uchida [35] is between 0.27 and 0.29, and 0.05 and 0.07when considering that the carbon dioxide molecules are trapped inthe small cages of structures sI and sII, respectively. In the largecages, the corresponding values are between 350 and 200, and80 and 60, respectively. The comparison between our results andthose obtained by Uchida [35] is thus quite disappointing butcan be explained by the different methods used to calculate theinteractions between carbon dioxide molecules and the cages ofthe clathrates. Indeed, the model used by Uchida [35] is based onthe Lennard-Jones and Devonshire formalism whereas we usedhere an all-atom approach.

5. Conclusions

In the present work calculations of the Langmuir constants andtheir temperature dependence for simple-guest clathrate hydrateshave been performed for eighteen gas species using the van derWaals and Platteeuw model and an all-atom approach for calculat-ing the interactions between guest and water molecules in theclathrate cages. This approach accounts for the atomic characterof the guest species and for the non-spherical water environment.

Then, the temperature dependence of the Langmuir constants inthe range 50 K–300 K is given in the form of a van’t Hoff expressionwith parameters obtained from a fit to calculated values.

This simple expression could thus be easily applied in varioussituations, especially when experimental data are not available.For example, it can help planetologists in the determination ofthe fractional occupancies of gas species trapped in the clathratesthat are suspected to exist in various solid bodies of the Solar Sys-tem. Indeed, this information is mandatory to analyze the atmo-spheric compositions of planets and satellites like Mars andTitan, for example, and thus to better understand their way of for-mation. Unfortunately, the temperatures usually considered inclathrate experiments are often far from the temperature rangeof interest for Planetology. As a consequence calculations of clath-rate occupancies is a preliminary step for any formation scenariothat would consider clathrate influence. In that purpose, usingthe van’t Hoff expression given here would be much more simplethan the usual approaches.

Conflict of interest

There is no conflict of interest.

Acknowledgements

This work has been carried out thanks to the support of theINSU EPOV interdisciplinary program.

Simulations have been executed on computers from the UtinamInstitute of the Université de Franche-Comté, supported by theRégion de Franche-Comté and Institut des Sciences de l’Univers(INSU).

A. Lakhlifi also thanks Kevin van Keulen and Sékou Diakité forfruitful discussions.

O. Mousis thanks the A⁄MIDEX project (no ANR-11-IDEX-0001–02) founded by the ‘‘Investissements d’Avenir’’ French Governmentprogram, managed by the French National Research Agency (ANR).

References

[1] E.D. Sloan, Clathrate Hydrates of Natural Gases, Marcel Dekker, New York,1998.

[2] E.D. Sloan, C.A. Koh, Clathrate Hydrates of Natural Gases, third ed., CRC/Taylor& Francis, Boca Raton, 2008.

[3] K.A. Sparks, J. Tester, J. Phys. Chem. 96 (1992) 11022.[4] Z. Cao, J.W. Tester, K.A. Sparks, B.L. Trout, J. Phys. Chem. B 105 (2001) 10950.[5] J.B. Klauda, S.I. Sandler, J. Phys. Chem. B 106 (2002) 5722.[6] B.J. Anderson, J.W. Tester, B.L. Trout, J. Phys. Chem. B 108 (2004) 18705.[7] B.J. Anderson, Molecular modeling of hydrate-clathrates via ab initio, cell

potential, and dynamic methods (Ph.D. thesis), Cambridge, 2005.[8] B.J. Anderson, M.Z. Bazant, J.W. Tester, B.L. Trout, J. Phys. Chem. B 109 (2005)

8153.[9] R. Sun, Z. Duan, Geochim. Cosmochim. Acta 69 (2005) 4411.

[10] O. Mousis, E. Chassefière, J. Lassue, V. Chevrier, M.E.E. Madden, A. Lakhlifi, J.I.Lunine, F. Montmessin, S. Picaud, F. Schmidt, T.D. Swindle, Space Sci. Rev. 174(2013) 213.

[11] J.H. van der Waals, J.C. Platteeuw, Adv. Chem. Phys. 2 (1959) 1.[12] I. Langmuir, Phys. Rev. 8 (1916) 149.[13] C. Thomas, Etude de la séquestration d’éléments volatils par des clathrates

hydrates: Application aux atmosphères planétaires (Ph.D. thesis), BesançonCedex, France, 2009.

[14] C. Thomas, O. Mousis, V. Ballenegger, S. Picaud, Astron. Astrophys. 474 (2007)L17.

[15] C. Thomas, S. Picaud, O. Mousis, V. Ballenegger, Planet. Space Sci. 56 (2008)1607.

[16] O. Mousis, J.I. Lunine, C. Thomas, M. Pasek, U. Marboeuf, Y. Alibert, V.Ballenegger, D. Cordie, Y. Ellinger, F. Pauzat, S. Picaus, Astrophys. J. 691 (2009)1786.

[17] C. Thomas, O. Mousis, S. Picaud, V. Ballenegger, Planet. Space Sci. 57 (2009) 42.[18] T.D. Swindle, C. Thomas, O. Mousis, J.I. Lunine, S. Picaud, Icarus 203 (2009) 66.[19] M.Z. Bazant, B.L. Trout, Physica A 300 (2001) 139.[20] J.I. Lunine, D.J. Stevenson, Astrophys. J. Suppl. Ser. 58 (1985) 493.[21] J.E. Lennard-Jones, A.F. Devonshire, Proc. Roy. Soc. London A 163 (1937) 53.[22] J.E. Lennard-Jones, A.F. Devonshire, Proc. Roy. Soc. London A 165 (1938) 1.[23] A. Lakhlifi, C. Girardet, J. Mol. Struct. 110 (1984) 73.[24] D.J. Wales, P.L.A. Popelier, A.J. Stone, J. Chem. Phys. 102 (1995) 5551.[25] M.B.H. Ketko, G. Kamath, J.J. Potoff, J. Phys. Chem. B 115 (2011) 4949.[26] S. Alavi, R. Susilo, J.A. Ripmeester, J. Chem. Phys. 130 (2009) 174501.[27] M.A. Spackman, J. Chem. Phys. 85 (1986) 6587.[28] E.A. Pidko, V.B. Kazanski, Kinetics Catalysis 46 (2005) 407.[29] J.S. Muenter, J. Chem. Phys. 94 (1991) 2781.[30] C. Thomas, S. Picaud, V. Ballenegger, O. Mousis, J. Chem. Phys. 132 (2010)

104510.[31] J.B. Klauda, S.I. Sandler, Chem. Eng. Sci. 58 (2003) 27.[32] L.J. Chen, C.H. Lin, Y.T. Yeh, M.K. Hsieh, S.T. Lin, Y.P. Chen, in: Proceedings of the

Seventh International Conference on Gas Hydrates: July 17–21, Edinburgh,Scotland, United Kingdom, 2011.

[33] W.R. Parrish, J. Prausnitz, Indus. Eng. Chem. Process Des. Dev. 11 (1972) 26.[34] K.A. Udachin, C.I. Ratcliffe, J.A. Ripmeester, in: Proceedings of the Fourth

International Conference on Gas Hydrates: May 19–23, Yokohama Symposia,Yokohama, Japan, 2002.

[35] T. Uchida, Waste Manag. 17 (1997) 343.


Recommended