Вы находитесь на странице: 1из 19

Molecular Physics

An International Journal at the Interface Between Chemistry and


Physics

ISSN: 0026-8976 (Print) 1362-3028 (Online) Journal homepage: https://www.tandfonline.com/loi/tmph20

Molecular simulation of shale gas adsorption onto


overmature type II model kerogen with control
microporosity

Lukáš Michalec & Martin Lísal

To cite this article: Lukáš Michalec & Martin Lísal (2017) Molecular simulation of shale gas
adsorption onto overmature type II model kerogen with control microporosity, Molecular Physics,
115:9-12, 1086-1103, DOI: 10.1080/00268976.2016.1243739

To link to this article: https://doi.org/10.1080/00268976.2016.1243739

Published online: 24 Oct 2016.

Submit your article to this journal

Article views: 420

View related articles

View Crossmark data

Citing articles: 25 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tmph20
MOLECULAR PHYSICS, 
VOL. , NOS. –, –
http://dx.doi.org/./..

SPECIAL ISSUE IN HONOUR OF JOHANN FISCHER

Molecular simulation of shale gas adsorption onto overmature type II model


kerogen with control microporosity
Lukáš Michaleca,b and Martin Lísala,b
a
Laboratory of Aerosols Chemistry and Physics, Institute of Chemical Process Fundamentals of the CAS, v. v. i., Prague, Czech Republic;
b
Department of Physics, Faculty of Science, J. E. Purkinje University, Ústí n. Lab., Czech Republic

ABSTRACT ARTICLE HISTORY


We use an all-atom molecular dynamics simulation to generate the dense porous structures of over- Received  July 
mature type II kerogen with control microporosity. The structures mimic the organic part of Barnett Accepted  September 
shale under a typical reservoir condition of 365 K and 275 bar. First, we build an atomistic model of
KEYWORDS
kerogen unit using the consistent valence force field and chemical structure proposed by Ungerer Barnett shale; clay; multiscale
and his colleagues. Second, we generate kerogen structures by gradual cooling and compression kerogen model; organic
of the initial low-density random configurations of kerogen units. During the structure generation, matter; pore space
we use a dummy particle of varying size to introduce microporosity into the kerogen structures. We accessibility
systematically characterise the microporous kerogen structures by calculating the geometric pore
size distribution, pore limiting diameter, maximum pore size, accessible surface area, and pore vol-
ume and by analysing the pore network accessibility. Third, we employ grand canonical Monte Carlo
(GCMC) to study the adsorption of two proxies of shale gas (pure methane and mixture of 82% of
methane, 12% of ethane and 6% of propane) in the kerogen structures. The shale gas adsorptions
are compared with GCMC simulation of CO2 adsorption in the kerogen structures. Hydrocarbons are
modelled using the all-atom optimized potential for liquid simulations (OPLS) force field while car-
bon dioxide is represented by the EPM2 model. We complement the adsorption studies by exploring
accessibility of pore space of the kerogen structures using molecular dynamics simulation. Finally,
we introduce a mesoscale pore void into a microporous kerogen structure and probe the adsorption
behaviour of the hydrocarbon mixture in such a multiscale kerogen model.

1. Introduction natural gas supply worldwide in recent years. The shale


gas is extracted by hydraulic fracturing which is the
Shale gas is an important unconventional energy resource
process where typically water with additives is pumped
that has had a potential game-changing effect on
into shale formations to fracture the rock and increase

CONTACT Martin Lísal lisal@icpf.cas.cz


©  Informa UK Limited, trading as Taylor & Francis Group
MOLECULAR PHYSICS 1087

the extraction rate of natural gas or crude oil [1,2]. It


should be noted that any liquid may be used to create
fractures and light hydrocarbons like propane have been
also considered for fracturing the rock. Shale is com-
prised of two distinct parts: organic material and clay
minerals. The organic material is primarily composed
of kerogen, a mixture of organic chemical compounds
insoluble in common polar solvents such as chloro-
form or dichloromethane [3]. Clays are comprised of
a large number of particles arranged in piles of sheets.
The most abundant types of clay minerals are smectites
(swelling) and illites (non-swelling). Besides clay miner-
als, the inorganic part of shales may also contain fine-
grained quartz, carbonates, pyrite and so on. Kerogen
is generally hydrophobic while clay is hydrophilic, and
water can easily be adsorbed onto clay surfaces [3,4].
In shale formations, shale gas (typically a mixture of
methane, ethane and propane with a very small amount
of butane and heavier hydrocarbons as well as carbon
dioxide and nitrogen) is stored as free gas, adsorbed gas
and dissolved gas. The free gas gathers in the fractures
and pores of a shale rock, the adsorbed gas occurs on
the surfaces of both the organic material and clay min-
erals, and the dissolved gas enriches the organic mat-
ter. The adsorbed and dissolved gas is in equilibrium
with a homogeneous free gas phase in an interconnected Figure . The van Krevelen diagram, a plot of hydrogen-to-carbon
shale pore structure. If some shale pores are not inter- (H/C) and oxygen-to-carbon (O/C) atomic ratios, was used to clas-
connected, there may be a departure from the equilib- sify different types of kerogen in terms of depositional origin and
rium between the adsorbed and dissolved gas, and the maturity [,]. R is the vitrinite reflectance, a geochemical indica-
tor that measures thermal maturity of type II kerogens by integrat-
free gas. The amount of free gas is relatively easy to esti- ing the effects of time and temperature during the maturation of
mate based on the temperature and formation pressure, sediments. The star indicates the overmature type II kerogen stud-
its porosity, and the fraction of porosity which is filled by ied in this work.
the gas. The contribution of adsorbed gas to the total gas
in place (GIP), estimated to reach up to 60%, is still poorly of kerogens due to exposure to high temperature and
understood, mainly due to complexity of the gas adsorp- pressure over geological time scales leads to a shift of the
tion on shale. Generally, most of the adsorption area is kerogen H/C and O/C atomic ratios in the van Krevelen
located in the organic material and the contribution of diagram from the top-right to bottom-left corners.
the adsorbed gas to the total GIP is less significant in the In recent years, there has been an increasing effort
inorganic matrix [3]. to use atomistic simulations and provide molecular level
Kerogen, a main part of the organic matter, is a insight into shale gas behaviour in different types of kero-
complex material composed of an amorphous porous gen. The effort starts with the construction of realis-
carbon skeleton and exhibits significant pore-shape and tic kerogen models which is followed by simulations of
pore-connectivity variations. Depending on its shale gas adsorption, diffusion and transport in the kero-
geographic origin, maturity and sedimentary history, gen models. The structure of real kerogens is unknown
kerogen displays a broad range of density and, chemical and experimental techniques such as XPS, 13 C NMR and
composition in terms of atomic contents and chemical S-XANES are utilised to characterise kerogen samples
functions, porosity and tortuosity. The van Krevelen in terms of their elemental composition (H, C, O, S,
diagram (Figure 1) [5,6], a plot of hydrogen-to-carbon N), functional groups containing carbon and hydrogen,
(H/C) and oxygen-to-carbon (O/C) atomic ratios, pro- oxygen, sulphur and nitrogen, percentage of aromatic
vides a tool to distinguish different types of kerogen in carbon with attachments, average number of aromatic
terms of depositional origin: type I (lacustrine), type II carbons per polyaromatic cluster and average aliphatic
(marine), type III (terrestrial) and type IV (originating carbon chain length [7]. The experimental data on the
from residues) and maturity. Increasing the maturity kerogen characterisation can then be employed first to
1088 L. MICHALEC AND M. LÍSAL

reconstruct a two-dimensional model of kerogens and the percentage of aromatic carbons linked with hydro-
subsequently based on the two-dimensional model, to gen atoms only. The process of building the kerogen unit
construct a three-dimensional model of kerogens using a avoids stacking the polyaromatic nuclei. Finally, start-
combination of ab initio and molecular mechanics calcu- ing from a low-density random configuration of kerogen
lations. Siskin’s model of Green River oil shale kerogen is units, a dense kerogen structure is typically obtained by
a typical example of such an approach [8–10]. However, gradual cooling and compression of the initial configura-
the size of such kerogen models is generally large since tion using a molecular dynamics (MD) simulation [18].
it is necessary to match the elemental analysis including Besides the effort to build realistic kerogen mod-
minor elements such as sulphur and nitrogen, and the els and determine their volumetric, thermodynamic
qualitative distribution of pyrolysis products. and mechanical properties [11,17], grand canonical
A somewhat different approach was taken by Coasne’s Monte Carlo (GCMC) and equilibrium as well as non-
group [11] that employed a Molecular Dynamics- equilibrium MD [18,19] have been used to study gas
Hybrid Reverse Monte Carlo (MD-HRMC) reconstruc- adsorption, diffusion and transport in model kerogen
tion method [12] and built various realistic kerogen structures under typical reservoir conditions. The studies
models that differ in geological origin, elemental com- include GCMC simulations of adsorption of pure alka-
position and maturity. By imposing the H/C and O/C nes and alkane mixtures, N2 and CO2 onto type II kero-
atomic ratios, distribution of chemical groups and carbon gen models as well as proxies of kerogen, the equilibrium
hybridisation, the MD-HRMC method, powered with MD of alkane and CH4 /CO2 diffusion in dense kerogen
reactive force fields [13,14], reconstructs a model of kero- structures, and non-equilibrium MD of alkane transport
gens by matching experimental and simulated C–C pair in microporous structures of kerogen [11,15,20–26], to
distribution functions. name just a few.
A rather general and efficient approach for the con- GCMC simulations provide the total gas content,
struction of realistic kerogen models of different types i.e. the entire quantity of gas that resides in the pore
and maturity has recently been proposed by Ungerer space at a given temperature and pressure. To bet-
and his colleagues [15–17]. The approach first consid- ter understand adsorption behaviour in the kerogens,
ers model units with 170–260 carbon atoms (300–650 it is useful to differentiate between the number of
atoms in total) that match as closely as possible the ele- gas molecules that would fill a given volume in the
mental analysis, the distribution of functional groups and absence of pore walls and the number of molecules that
the structural features of kerogen. The units then serve as would fill that same volume, but with the inclusion of
building blocks for the generation of dense porous kero- pore-wall effects at the same temperature and pressure
gen structures. [27–29]. This can be discussed in terms of the excess
Building a kerogen unit begins with the selection of adsorption defined as the additional amount of gas
a total number of carbon atoms and a number of aro- adsorbed per unit of pore volume compared with the
matic carbons to match the aromaticity that is known amount of gas in the same volume of a given pore in the
from XPS and NMR, the determination of the number of absence of pore walls [28]. To systematically examine
H, O, S and N atoms required to approximate the elemen- the effect of pressure, temperature and pore size on
tal analysis and distribution in functional groups, and methane adsorption in kerogens (which is carbon-based
determination of the number of aromatic units match- three-dimensional pore-networks with a significant
ing the number of aromatic carbons per aromatic unit. amount of surface area contained in micropores) Mosher
The second step includes building a preliminary struc- et al. [28] simulated methane adsorption in idealised
ture of the kerogen unit using an appropriate force field carbon slits with widths ranging from 4 to 90 Å at various
by matching equilibrium bond lengths, bond angles and temperatures and pressures characteristic of shale rock
dihedral angles as closely as possible and by selecting the conditions. They found that when pressure increased, the
appropriate level of cross-linking between aromatic units. amount of the excess increased up to a certain pressure,
The preliminary structure is then subject to energy min- and then decreased to zero at the pressure for which
imisation, i.e. atom coordinates of the structure are itera- there was no noticeable change in density as a result of
tively adjusted to minimise the potential energy given by adsorption. The increase with pressure followed by a sub-
the force field. In the last step, the minimised structure sequent decrease suggests that there exists a maximum
is refined to improve the agreement with elemental and adsorbed quantity. An increase of temperature reduced
functional group analyses. The refinement involves either adsorption for equivalent pressures, but the maximum
a change in the shape of polyaromatic units or a change number of adsorbed molecules remains constant. As
in the location of functional groups to better represent pressure increased at higher temperatures, the excess
the percentage of aromatic carbons with attachments and adsorption isotherm peaks and subsequently descends at
MOLECULAR PHYSICS 1089

higher pressures than for the same-sized pores at lower Table . The MD relaxation procedure for the generation of dense
temperatures. When the slit width was varied, larger porous kerogen structures. NVT denotes the canonical MD while
pores exhibit a lower excess density as compared to NPT denotes the constant pressure–constant temperature MD
with anisotropic box changes. T is the temperature, t is the length
smaller pores. At a pressure above 10 bar, the adsorp- of particular MD runs and P is the pressure. The values of t were
tion capacities of 6 Å pores drops below those of the set to be long enough to ensure the relaxation of the inter-
wider pores and decreases below that of the 12 Å pore at and intramolecular configurational energy, and system density to
180 bar. The density of the adsorbed methane changes their equilibrium values.
non-monotonically with increasing pore width, and Ensemble T (K) t (ns) P (bar)
drops to a minimum in 12 Å pores at 120 bar [28].
NVT  . –
In this work, we employ GCMC and study the adsorp- NPT  . 
tion of shale gas and carbon dioxide onto overma- NPT  →  . 
NPT  . 
ture type II kerogen structures of varying microporos- NPT  →  . 
ity which aim to mimic the organic part of Barnett shale. NPT  . 
Although kerogen is generally the most abundant compo- NPT  →  . 
NPT  . 
nent of the organic matter, a real organic matter also con- NPT  .  → 
tains asphaltenes, resins, hydrocarbons and other fluids NPT  . 
NPT  →  . 
such as carbon dioxide, water and nitrogen. Simulations NPT  . 
of organic matter containing both kerogen and minor
components will be the subject of our future work. Shale
gas and CO2 are modelled using the all-atom optimized kerogen is indicated by a star. The kerogen unit has a
potential for liquid simulations (OPLS) [30] and EPM2 chemical formula C175 H102 O9 S2 N4 , the molecular weight
[31,32] force fields, respectively. The kerogen structures is Mw = 2468.9 g/mol, and its molecular model is shown
are built using MD simulations from a kerogen unit in Figure 2(a). Figure 2 also provides the composition and
whose chemical structure was proposed by Ungerer et al. structural parameters of the kerogen unit.
[17]. The kerogen structures are modelled using the con- To generate a representative kerogen structure, we
sistent valence force field (CVFF) [33]. In addition, we placed 12 kerogen units (to avoid self-interactions of
explore the accessibility of the pore space of the kerogen kerogen units through the periodic boundary conditions)
structures and probe shale gas adsorption behaviour in to a large simulation box of 100 Å × 100 Å × 100 Å (initial
a multiscale kerogen model for the shale gas. The mul- kerogen density about 0.05 g/cm3 ) with periodic bound-
tiscale kerogen model is constructed by introducing a ary conditions in all directions and compressed the ini-
mesoscale pore void into our microporous kerogen struc- tial kerogen configuration from a high temperature to a
ture. The remainder of the paper is organised as fol- typical reservoir condition of 365 K and 275 bar, corre-
lows. The methodology for the generation of dense kero- sponding to the Barnett shale. The kerogen units were
gen structures with control microporosity is outlined in placed into the simulation box randomly without over-
Section 2. Section 3 presents molecular models of the lap and with the same orientation. We chose the same
kerogen structure, methane, ethane, propane and car- initial orientation of the kerogen units since they are flat
bon dioxide. Section 4 provides computational details and we anticipated they tend to stack parallel in a com-
about the MD and GCMC simulations. In Section 5, we pressed state. The compression and cooling of the initial
then present and discuss our results for the adsorption configuration was performed in stepwise fashion using
of methane, mixture of methane, ethane, and propane MD, and Table 1 summarises the details of MD relaxation
and carbon dioxide in the kerogen structures. Finally, we procedure. In contrast to Ungerer’s MD relaxation proce-
present our conclusions in Section 6. dure [15–17], we added a refinement step that involved
heating the final configuration at 365 K and 275 bar to
2000 K, followed by 1 ns gradual cooling back to 365 K
2. Methodology
under a constant pressure of 275 bar and a subsequent
1 ns relaxation run at 365 K and 275 bar. The tempera-
2.1. Microporous kerogen structures
ture of 2000 K guaranteed that the system did not expand
We are interested in shale gas behaviour in kerogens asso- when we switched on the gradual cooling. This leads to
ciated with unconventional gas reserves [34]. Therefore, refinement of the stacking of the kerogen units, and in
we used the chemical structure of kerogen unit proposed turn, to an increase in system density. Figure 2(b) displays
by Ungerer et al. [17] that represents overmature type II an example of the time variation of the system tempera-
kerogens found in shale formations of the Barnett play ture and kerogen density during MD generation of the
[35]. In the van Krevelen diagram (Figure 1), the studied kerogen structures. A final structure typically contained
1090 L. MICHALEC AND M. LÍSAL

Figure . (a) Molecular model of a kerogen unit representing overmature type II kerogen. The red, blue, yellow, grey and white spheres
represent oxygen, nitrogen, sulphur, carbon and hydrogen atoms, respectively. The chemical formula of the kerogen unit is C H O S N ,
its molecular weight Mw = . g/mol, and the atomic ratios H/C = ., O/C = ., S/C = . and N/C = .. Another composition
and structural parameters of the kerogen unit can be found in Ref. []. (b) Example of time variation of the system temperature and
kerogen density during the MD generation of the porous kerogen structure which was accomplished within . ns. (The colour ranges
refer to the online version of the paper).
MOLECULAR PHYSICS 1091

Figure . Geometric pore size distribution, PSD, as a function of the pore diameter for the porous kerogen structures with different micro-
porosity. The microporosity was introduced via the dummy particle of size σ LJ .

Table . The kerogen density, ρ, accessible surface area, SA , constant pressure–constant temperature (NPT) ensem-
pore volume, Vpore , pore limiting diameter, Dmin , and maxi- ble with anisotropic box changes. We employed the
mum pore size, Dmax , of the kerogen structures with different Nose–Hoover thermostat [37,38] and the Nose–Hoover
microporosities corresponding to the dummy particle of size
σ LJ . The simulation uncertainties are given in the last digits as barostat [39,40].
subscripts. The GCMC established an equilibrium between a
homogeneous free gas in an interconnected shale pore
σ LJ (Å) ρ (g/cm ) SA (m /g) Vpore (cm /g) Dmin (Å) Dmax (Å)
structure and adsorbed gas in the kerogen network, i.e.
 . . . . . equivalence of the chemical potentials of gas species. The
 . . . . .
 . . . . .
chemical potential of the free gas was obtained by config-
 . . . . . urational bias Widom’s insertion method [18,19] during
 . . . . . an NPT MC simulation at 365 K and 275 bar, performed
in a cubic simulation box with periodic boundary condi-
ultramicropores (pore width <1nm) that was a result of tions in all directions.
the void space left between kerogen units in the com- The GCMC employed two types of trial moves: parti-
pressed state. To introduce micropores (pore width of 1–2 cle displacement and particle insertion/deletion to/from
nm) into our kerogen models, we used a dummy particle the kerogen structure. The displacement moves involved
of varying size during the MD relaxation procedure. The transitional, rotational and configurational bias moves to
dummy particle was then removed from the compressed satisfy temperature equilibration. The insertion/deletion
kerogen structure, leaving a void where gas adsorption is moves included insertion of a gas molecule in the
likely to occur. kerogen structure or removal of a gas molecule from
the kerogen structure, both using the configurational
bias technique, to ensure chemical potential equili-
2.2. MD and GCMC bration [19]. The NPT MC and GCMC simulations
were carried out using the Towhee simulation package
The MD simulations were performed with the LAMMPS [41].
code [36] in a canonical (NVT) ensemble and
1092 L. MICHALEC AND M. LÍSAL

3. Molecular models 4. Computational details


We modelled kerogen structures using the CVFF [33] We used the particle–particle particle–mesh algorithm
which is a standard and reliable force field for simulations [44] in the MD and Ewald summation technique [18] in
of organic molecules. The functional form of the CVFF NPT MC and GCMC, both with a precision of 1 × 10−4
is [45], to treat long-range electrostatic interactions. We fur-
ther utilised the cut-off radius between 10 and 17 Å, and
U = Ubonded + Unonbonded time step of 1 fs in MD. In NPT MC, we typically used
  500 gas molecules, 5 × 106 equilibration moves, 10 × 106
= kb (b − b0 )2 + kθ (θ − θ0 )2
production moves, and frequency for conformation, rota-
bonds angles
    tion, translation and volume attempts equal to 0.33, 0.33,
+ kφ 1 + cos nφ φ (1) 0.33 and 0.01, respectively. Ten virtual insertions every
dihedrals
 500 production moves were performed to measure the
  
+ kψ 1 + cos nψ ψ chemical potential of the free gas by configurational bias
impropers Widom’s method. Similarly, in GCMC, we utilised 5 × 106
 
12
6 equilibration moves, 10 × 106 production moves, and fre-
 σi j σi j qi q j
+ 4εi j − + quency for conformation, rotation, translation and inser-
ri j ri j 4π0 ri j
i j>i tion/deletion attempts equal to 0.25, 0.25, 0.25 and 0.25,
respectively.
where b, θ, φ and ψ denote the bond length, bond angle, We systematically characterised the kerogen structures
dihedral angle and improper angle, respectively; b0 and θ 0 by calculating the geometric pore size distribution (PSD),
are the equilibrium bond length and bond angle, respec- pore limiting diameter, Dmin , maximum pore size, Dmax ,
tively; kb , kθ , kφ and kψ are the force constants; nφ and nψ accessible surface area, SA , and pore volume, Vpore , and
are the dihedral and improper angle parameters, respec- by analysing the pore network accessibility [46]. The PSD
tively; εij and σ ij correspond to the Lennard–Jones (LJ) was determined by MC procedure that probed the kero-
energy and size parameters, respectively; qk is the partial gen pore space to find the largest spheres that contained
charge on atom k; ϵ0 is the permittivity in vacuum and rij test points and did not overlap with the atoms of the kero-
is the distance between atom i and atom j. The dummy gen structure. The test points were randomly placed in the
particle, employed to control the structure microporos- simulation box with no overlaps with the kerogen atoms.
ity, was represented by a repulsive LJ particle with Each sphere then defined the volume of a void space that
Mw = 220 g/mol, εLJ /k = 100 K and σLJ = can be covered by a sphere of radius r or smaller, Vp (r).
{0, 9, 11, 13, 15}Å; k is the Boltzmann constant. The derivative −dVp /dr is the PSD and was obtained by
We employed the all-atom OPLS models for methane, numerical differentiation of Vp (r) [47]. The value of Dmax
ethane and propane [30], and the flexible EPM2 model is a by-product of the PSD calculation.
for carbon dioxide [31,32]. Both OPLS and EPM2 force The accessible surface area corresponds to the posi-
fields have been shown to reproduce rather well the ther- tions of the centre of a probe particle rolling over the
modynamic and vapour–liquid equilibrium properties kerogen atoms. It is thus defined as a locus of the points
[42,43], and in addition, the force fields are compati- that represents the location of the probe particle at a dis-
ble with CVFF [33]. The intramolecular interactions in tance of the collision diameter σ from a kerogen atom
methane, ethane and CO2 included the vibrational and and at least a distance σ from all other atoms of the kero-
bending contributions, while those in propane involved gen structure. Calculation of SA did not involve the actual
the vibrational, bending and torsion contributions. The rolling of the probe particle but employed an MC proto-
functional forms of the intramolecular contributions col, described in detail elsewhere [46]. As a probe parti-
are given by Equation (1). The intermolecular interac- cle, we used nitrogen atom with σN = 3.314 Å to directly
tions were modelled by the LJ and Coulombic potentials relate the calculated SA to that measured in the BET
(see Equation (1)). The cross-LJ parameters were given by adsorption experiments. An analogous MC approach was
the Lorentz–Berthelot combining rules [18]: used for calculation of the pore volume where a He-
size probe of σHe = 2.58 Å was used to provide values of
√ σii + σ j j Vpore consistent with that determined experimentally by
εi j = εii ε j j , σi j = (2) helium pycnometry.
2
The analysis of the pore space accessibility aimed to
For the values of LJ parameters and partial charges, we detect if the pore network in the kerogen structures was
refer the reader to the original papers [30–33]. fully accessible to a spherical probe of the size chosen to
MOLECULAR PHYSICS 1093

be equal to σ N . The kerogen pore network was accessible is in subcritical state. The model critical temperature, Tc ,
to the spherical probe if, for this probe, it was possible and pressure, Pc , have the following values: (Tc , Pc ) ࣃ
to construct a continuous trajectory from one face of the (191K, 46bar) for methane, (Tc , Pc ) ࣃ (285K, 47bar) for
simulation box to the opposite face of the simulation box ethane, (Tc , Pc ) ࣃ (372K, 46bar) for propane, and (Tc , Pc )
without overlapping with the kerogen atoms. The value ࣃ (306K, 77bar) for CO2 [42,43]. They compare rather
of Dmin is a by-product of the percolation analysis. The well with experimental Tc and Pc : (Tc , Pc ) ࣃ (191K, 46bar)
pore limiting diameter, Dmin , corresponds to a maximum for methane, (Tc , Pc ) ࣃ (305K, 49bar) for ethane, (Tc ,
probe size for which a pore network still percolates. The Pc ) ࣃ (370K, 43bar) for propane, and (Tc , Pc ) ࣃ (304K,
accessibility of the kerogen pore space was explored in all 74bar) for CO2 [48]. In addition, the model vapour
three directions and details about its determination are pressure of propane is Psat ࣃ 42 bar [42] while the exper-
given elsewhere [46]. imental Psat ࣃ 39 bar [48].

5. Results and discussion


5.1. Kerogen structure characterisation
All of the simulations were performed at a temperature
T = 365 K and pressure P = 275 bar, corresponding to a Starting from different initial configurations of kerogen
reservoir condition of the Barnett shale formation. The units, we typically generated 20 compressed configura-
mixture of 82% of methane, 12% of ethane and 6% of tions for each size of the LJ dummy particle. The density
propane used in this work as a proxy of the shale gas rep- equilibration was slightly influenced by the initial con-
resents a typical composition of the shale gas extracted figurations. For adsorption simulations, we then chose
in the Barnett play [35]. At this T, methane, ethane and a high-density structure out of the 20 generated con-
carbon dioxide are in supercritical state, while propane figurations that had the lowest configurational energy.

Figure . Examples of the kerogen structures with low and high microporosities. The left column corresponds to the structure generated
without the dummy particle while the right column is for the structure generated with the dummy particle of size σLJ = 13 Å. We displayed
the periodic images for the sake of clarity. In portions (a) and (b) of the figure, the red, blue, yellow, grey and white spheres represent
oxygen, nitrogen, sulphur, carbon and hydrogen atoms, respectively. The portions (c) and (d) of the figure show the front and top views
of the kerogen structures with the cutting plane in the middle of simulation box to demonstrate percolation of the pore network probed
by nitrogen sphere. (The colour ranges refer to the online version of the paper).
1094 L. MICHALEC AND M. LÍSAL

In addition, we ran a 1 ns NPT MD and collected 10 struc-


tures separated by 0.1 ns for this structure. The dummy
particle was then removed from the collected structures.
The structures were statistically equivalent and repre-
sented the overmature type II kerogen. The size of corre-
sponding simulation boxes was (40–50) Å × (30–35) Å ×
(25–30) Å.
Table 2 summarises the various properties used to
characterise the kerogen structures. Figure 3 displays the
PSD of the kerogen structures with different microp-
orosities corresponding to the dummy particle with σLJ =
{0, 9, 11, 13, 15}Å. Results in Table 2 and Figure 3 were
averaged over the collected structures. Figure 4 provides
examples of the kerogen structures with low and high
microporosities. First, we see that density of the kerogen
structures was typically between 1.2 and 1.3 g/cm3 which
compared quite well with a range of the experimental
density 1.2–1.4g/cm3 found for mature and overmature
kerogens [49]. The kerogen density also agrees with a
range of simulation density 1.21–1.28 g/cm3 reported by
Ungerer et al. [17] on the same kerogen model but sim-
ulated with a different force field. Second, the PSD cor-
responding to no dummy particle suggested the pres-
ence of ultramicropores of size from 3–5 Å (see also
Figure 4(a)). The dummy particle created a host microp-
ore (where gas adsorption is likely to occur) proportional
to σ LJ as indicated by the pronounced peak in PSDs (see
also Figure 4(b)). Besides the host micropore, the kero-
gen structures corresponding to σLJ = {9, 11, 13, 15}Å
also contained smaller micropores and ultramicropores.
Experimental PSDs for isolated kerogens cannot be reli-
ably measured. However, experimental PSDs for shale
(i.e. including organic and inorganic matter) suggest the
existence of micropores larger than 1 nm [50,51], cf. PSDs
corresponding to σLJ = {13, 15}Å. Finally, the values of
SA , Vpore , and Dmax gradually increased with increasing
σ LJ while values of Dmin did not vary significantly with
σ LJ . In addition, the values of Dmin were small in com-
parison with the typical size of gas molecules, indicat-
ing a poor connectivity of pores in the kerogen structures
(see Figure 4(c,d)).

5.2. Adsorption
In GCMC simulations, the kerogen structures were
considered rigid for the sake of computational effi- Figure . The adsorbed amount of (a) methane, (b) shale gas and
ciency, neglecting swelling of adsorbent by the adsorbed (c) carbon dioxide, n, as a function of microporosity introduced
molecules. Tables 3–5 list the simulated adsorbed amount via the dummy particle of size σ LJ at a temperature of  K and
of methane, mixture of methane, ethane and propane pressure of  bar. The open symbols represent the GCMC results
as a proxy of the shale gas, and carbon dioxide, respec- while the filled symbols denote the MD results. For the MD sim-
ulations, we plot the maximal values of n only. The dotted lines
tively. Figure 5 then plots the simulated adsorbed amount,
through the GCMC results serve as a guide for the eye.
n, as a function of microporosity, i.e. in terms of
σ LJ .
MOLECULAR PHYSICS 1095

Table . The adsorbed amount of methane, nCH4 , in the Table . The adsorbed amount of the shale gas (i.e. mixture of
microporous structures of overmature type II kerogen at a % of methane, % of ethane and % of propane), nCH4 , nC2 H6 ,
temperature of  K and pressure of  bar. The microp- and nC3 H8 , in the microporous structures of overmature type II
orosity was introduced via the dummy particle of size σ LJ . kerogen at a temperature of  K and pressure of  bar. The
The results written in black correspond to GCMC simula- microporosity was introduced via the dummy particle of size
tions while those written in red and blue represent MD sim- σ LJ . The results written in black correspond to GCMC simulations
ulations for rigid and flexible kerogen structures, respec- while those written in red and blue represent MD simulations
tively, in the x-, y- and z-directions. The GCMC simulation for rigid and flexible kerogen structures, respectively, in the x-,
uncertainties are given in the last digits as subscripts. The y- and z-directions. The GCMC simulation uncertainties are given
MD simulations were typically run for  ns. The correspond- in the last digits as subscripts. The MD simulations were typically
ing value of the chemical potential μCH4 /k = 76668 K was run for  ns. The corresponding values of the chemical poten-
determined by the configurational bias Widom’s method tial μCH4 /k = 75767 K, μC2 H6 /k = 1709812 K, and μC3 H8 /k =
during NPT MC simulations; k is the Boltzmann constant. 2669215 K were determined by the configurational bias Widom’s
See Ref. [] for definition of the reference state for μCH4 . method during NPT MC simulations; k is the Boltzmann constant.
(The colour ranges refer to the online version of the paper). See Ref. [] for definition of the reference state for μCH4 , μC2 H6
and μC3 H8 . (The colour ranges refer to the online version of the
σ LJ (Å) nCH (mmol/g) paper).
4

 . σ LJ (Å) nCH (mmol/g) nC (mmol/g) nC (mmol/g)


H H
0.21/0.03/0.19 4 2 6 3 8

0.45/0.26/0.44
 . . .
 .
0.13/0.02/0.11 0.03/0.01/0.03 0.06/0.01/0.05
0.24/0.23/0.21
0.23/0.07/0.19 0.08/0.02/0.07 0.06/0.01/0.05
0.64/0.66/0.37
 . . .
 .
0.14/0.13/0.15 0.03/0.03/0.02 0.08/0.06/0.04
0.06/0.43/0.15
0.18/0.18/0.13 0.04/0.09/0.03 0.16/0.20/0.13
0.23/0.56/0.34
 . . .
 .
0.04/0.25/0.10 0.01/0.07/0.03 0.02/0.10/0.04
0.29/0.27/0.47
0.14/0.38/0.15 0.02/0.07/0.03 0.07/0.10/0.12
1.13/0.49/1.33
 . . .
 .
0.23/0.17/0.37 0.07/0.04/0.06 0.02/0.06/0.10
0.17/0.37/0.39
0.56/0.28/0.66 0.11/0.06/0.14 0.26/0.14/0.35
0.39/1.33/1.57
 . . .
0.11/0.24/0.24 0.02/0.05/0.07 0.05/0.07/0.09
0.14/0.34/0.70 0.01/0.12/0.18 0.14/0.28/0.29

We first see that n’s increased with rising σ LJ although


that increase was small for the kerogen structure corre-
sponding to σLJ = 9 Å. The values of nCH4 in the kero- the modified model was still higher than that for CH4 ,
gen structures corresponding to σLJ = {13, 15}Å com- demonstrating stronger van der Waals interactions for
pared rather well with the simulation results of Ho et al. CO2 –kerogen systems compared to CH4 –kerogen sys-
[26] in similar kerogen structures when Ho at al.’s sim- tems. Third, our results for the proxy of Barnett’s shale
ulation results were extrapolated to our temperature and gas, i.e. a mixture of 82% of methane, 12% of ethane
pressure. Second, gas adsorption onto the kerogen struc- and 6% of propane indicated preferential adsorption of
tures was stronger for CO2 than for CH4 and the ratio propane and to a lesser extent ethane. An example is
nCO2 /nCH4 was about 1.3. The same finding about CO2 that for the kerogen structure without the host microp-
and CH4 adsorption onto similar type II kerogen struc- ore (σLJ = 0 Å), the adsorbed mixture contained about
tures was reported by Sui and Yao [25]. A stronger CO2 74% of CH4 , 20% of C2 H6 and 6% of C3 H8 . With increas-
adsorption in comparison with CH4 adsorption on Bar- ing microporosity (i.e. with increasing σ LJ ), the molar
nett’s shale samples was also experimentally measured percentage of methane decreased, that of ethane was
by Heller and Zoback [52]. This can be attributed to roughly constant (i.e. around 20%) and that of propane
stronger van der Waals and electrostatic gas–gas and gas– gradually increased. For instance, for the kerogen struc-
adsorbent interactions in the case of CO2 adsorption. ture corresponding to σLJ = 15 Å, the adsorbed mixture
To better understand the stronger CO2 adsorption, we consisted of about 59% of CH4 , 18% of C2 H6 and 23%
performed an additional GCMC simulation for the kero- of C3 H8 . The strong preferential adsorption of propane
gen structure corresponding to σLJ = 15 Å and switched molecules can be associated with the condensation of the
off the partial charges in the CO2 model. We found propane molecules in larger micropores since, in contrast
that the adsorbed amount decreased by about 10% with to methane and ethane (which are in supercritical state),
respect to the original CO2 model, demonstrating an propane is in subcritical state above the vapour pressure.
enhancement of CO2 adsorption due to electrostatic A similar observation was reported by Falk et al. [22] for
interactions. However, the adsorbed amount of CO2 of subcritical dodecane in model kerogens.
1096 L. MICHALEC AND M. LÍSAL

Table . The adsorbed amount of car- [48]. From Equation (3), we obtained nexcess CH4  0.77 and
bon dioxide, nCO2 , in the microporous 0.95 mmol/g for the kerogen structures corresponding to
structures of overmature type II kerogen σ LJ = 13 and 15 Å, respectively. Extrapolation of experi-
at a temperature of  K and pressure
of  bar. The microporosity was intro- mental data to our T and P yielded a nexcessCH4  1.7 mmol/g
duced via the dummy particle of size σ LJ . for data from Gasparik et al.’s group [53] and nexcess CH4 
The results written in black correspond 0.7 mmol/g for data from Zhang et al.’s group [54].
to GCMC simulations while those writ- The excess adsorption data for the Barnett shale were
ten in red and blue represent MD sim- measured for the whole shale (i.e. including organic and
ulations for rigid and flexible kerogen
inorganic matter) and normalised by the total organic
structures, respectively, in the x-, y- and
z-directions. The GCMC simulation uncer- carbon. The fact that two different samples from Barnett
tainties are given in the last digits as sub- shales exhibited different excess adsorption data demon-
scripts. The MD simulations were typically strated the diversity and heterogeneity of shales. Hence,
run for  ns. The corresponding value of the simulation nexcessCH4 for the model Barnett kerogen was
the chemical potential μCO2 /k = 18095 within a range of experimental nexcess
K was determined by the configurational CH4 for the Barnett
bias Widom’s method during NPT MC sim- shale [53,54]. In addition, the total and excess adsorp-
ulations; k is the Boltzmann constant. See tion amount together with the pore volume enables us to
Ref. [] for definition of the reference evaluate the average total and excess density of methane,
state for μCO2 . (The colour ranges refer to excess
ρ av and ρav , respectively, in the kerogen pore networks
the online version of the paper). and compare them with the density of bulk methane,
σ LJ (Å) nCO (mmol/g) ρ PT = 891750 mol/m3 . For the kerogen structures
corresponding to σLJ = {0, 9, 11, 13, 15}Å, we obtained
2

 .
0.19/0.22/0.17
ρ av = {15068, 16097, 16474, 17235, 16876} mol/m3 and
excess
0.90/0.45/0.62 ρav = {9041, 8828, 8840, 8259, 8009} mol/m3 . Values
excess
 . of ρ av are about twice of ρ PT while values of ρav are
0.22/0.20/0.19
0.80/0.98/0.72 comparable or slightly below ρ PT .
 .
0.06/0.37/0.14
0.48/1.84/1.08
 .
5.3. Pore space accessibility
0.25/0.25/0.48
1.64/1.08/2.08 The particle insertion/deletion moves in GCMC can take
 . place anywhere in the kerogen structure and GCMC thus
0.16/0.37/0.40
0.48/1.97/2.31
allowed adsorption in pores that were not connected to
the external surface of the structure. In contrast, these
inaccessible pores would stay empty in a real adsorption
experiment. To understand the effect of the pore network
Our simulation data on the methane adsorption in the accessibility on the gas adsorption in the kerogen struc-
kerogen structures with the host micropore of size above tures we set up MD simulations that mimicked the real
10 Å, i.e. in those corresponding to σLJ = {13, 15}Å can adsorption experiment. In the following, we assume the
be compared with the experimental data for the Bar- kerogen is protected from the total stress supported by
nett shale reported by Gasparik et al. [53] and Zhang the mineral part of the shale.
et al. [54]. Comparison of the methane adsorption from An empty rigid kerogen matrix made up two copies
the simulations and experiments was made through the of the kerogen structure without the periodic boundary
excess adsorbed amount, nexcess . Excess adsorption is conditions in the α-direction (α ࣕ x, y, z) was put in con-
the difference between the (total) adsorbed amount and tact with a gas reservoir. The two copies of the kerogen
the amount of gas in the free volume in kerogen pores and structure guaranteed that the original kerogen structure
was estimated as was recovered in the middle of the kerogen matrix. For
practical reasons, we used two reservoirs on the opposite
nexcess = n − Vpore ρPT (3) sides of the kerogen matrix which enabled us to employ
the periodic boundary conditions in all directions and
where ρ PT is the density of bulk methane at T = 365 K speed up the adsorption process (see Figure 6(a)). We first
and P = 275 bar. The value of ρ PT = 891750 mol/m3 was equilibrated the system by NVT MD and then used NPT
obtained by NPT MC simulation during measurement of MD with box changes in the α-direction only (NPαα T
the chemical potential. The simulated value agreed rather MD) to allow the gas molecules to imbibe the kero-
well with the experimental value of ρ PT = 9263 mol/m3 gen matrix (see Figure 6(b–d)). We used Pαα = 275 bar,
MOLECULAR PHYSICS 1097

Figure . Molecular dynamics simulations mimicking a real adsorption experiment. (a) An empty rigid kerogen matrix in contact with
shale gas (i.e. a mixture of % of methane, % of ethane and % of propane). (b) Shale gas molecules imbibed the rigid kerogen matrix
from one side only. (c) Shale gas molecules imbibed a small percolating pore in the rigid kerogen matrix. (d) Methane molecules imbibed
the rigid kerogen matrix with poor space accessibility from the external surface. The red, blue and green spheres represent methane,
ethane and propane molecules, respectively. The light blue spheres denote the host micropores. (The colour ranges refer to the online
version of the paper).

which corresponds to a reservoir condition of the Bar- gas reservoirs as the location where the kerogen structure
nett shale formation, and Pαα represents the fluid pres- reached its bulk density reported in Table 2. We then com-
sure. We evaluated the adsorbed amount from density puted the values of n by integrating the gas number den-
profiles of gas molecules and kerogen atoms as depicted in sity profiles and kerogen mass density profile, and taking
Figure 7. Figure 7 shows an example of methane, ethane their ratio.
and propane number density profiles and kerogen mass The MD simulation results are summarised in
density profile in the direction of applied pressure. We Tables 3–5, and are also shown in Figure 5. In Figure 5,
first determined boundaries between bulk kerogen and we plotted the maximal values of n only. Figure 6(b–d)
1098 L. MICHALEC AND M. LÍSAL

Figure . The number density profiles of methane, ethane and propane and the mass density profile of kerogen atoms in the direction
of applied pressure. The dash–dotted lines denote the boundaries between the kerogen and shale gas (i.e. a mixture of % of methane,
% of ethane and % of propane).

further demonstrates various scenarios of the gas flexible kerogen matrix should be taken with care since
imbibition of the kerogen structures. We can see that for the moderate molecular weight of the model kerogen unit
methane and carbon dioxide, the MD results were below along with absence of the cross-linking between the kero-
the GCMC results by about 60 to 80%. gen units allows too much flexibility when compared with
The above MD results for the rigid kerogen matrix real kerogens.
elucidated the pore space accessibility of the kerogen
structures with respect to GCMC. We further explored
the effect of the flexibility of the kerogen matrix on 5.4. Kerogen model combining different porosity
gas imbibition. In contrast to MD simulations on the scales
rigid kerogen matrix, we allowed the kerogen atoms to
move. The kerogen matrix flexibility led to enhanced Kerogens usually have pore networks that combine more
gas imbibition of the kerogen structures since the than one length scale. Hence, we considered a multiscale
imbibed molecules might more easily diffuse through the model of kerogen with micropores (pore width <2 nm)
kerogen structures or might cause an opening and inter- and mesopores (pore width of 2–50 nm). Similarly, as
connection of the originally inaccessible pores. The MD was done in the previous Section 5.3, we used a rigid
simulation results for the flexible kerogen matrix are kerogen matrix formed by two copies of the kerogen
listed in Tables 3–5, and are displayed in Figure 5 where structure with micropores corresponding to the dummy
only maximal values of n were plotted. We can see that the particle with σLJ = 11 Å (cf. PSD in Figure 3) and with-
MD results for the flexible kerogen matrix were still below out the periodic boundary conditions in the y-direction.
the GCMC results by about 10%–50%. An exception was The mesopores were introduced through a 14-Å-wide
the mixture adsorption onto the kerogen structures cor- slit cavity placed in contact with the kerogen matrix (see
responding to σ LJ = 9 and 13 Å where nC3 H8 from MD Figure 9(a)). The size of the corresponding simulation
simulations was above that from GCMC simulations as box was 49.62 Å × 85.32 Å× 28.03 Å. The PSD of
a result of the propane preferential adsorption. Finally, the multiscale kerogen model is shown in Figure 9(b).
Figure 8 demonstrates the effect of the kerogen-matrix Micropores with the width <1 nm came from the kero-
flexibility by showing examples of methane imbibition for gen matrix while mesopores with the width around
the rigid and flexible kerogen matrix. The results for the 2.5–2.7 nm resulted from the addition of the cavity. Pores
MOLECULAR PHYSICS 1099

Figure . Methane molecules imbibed the rigid and flexible kerogen matrix during molecular dynamics simulations that mimic a real
adsorption experiment. In each portion of the figure, the upper snapshot shows a simulation corresponding to the rigid kerogen matrix
while the bottom snapshot displays a simulation corresponding to the flexible kerogen matrix. (a) The pronounced effect of kerogen-
matrix flexibility on the methane imbibition of the kerogen matrix. (b) The small effect of kerogen-matrix flexibility on the methane
imbibition of the kerogen matrix.

with the width ranging from 1–2.5 nm were due to the shale gas in the multiscale kerogen model. From the
interface between the kerogen matrix and cavity. GCMC
simulation,we obtained the adsorbed amounts:
We then used GCMC and studied the adsorp- nCH4 , nC2 H6 , nC3 H8 = {4.3913 , 1.168 , 1.278 } mmol/g.
tion of the shale gas, represented by the mixture of Similarly, as was done in the previous Section 5.3,
82% of methane, 12% of ethane and 6% of propane we can compute the contribution to ni (i ࣕ CH4 , C2 H6 ,
[35]. Figure 10(a) displays the simulation density C3 H8 ) from the kerogen matrix using the density profiles
profiles of methane, ethane and propane, and kero- from Figure 10(a). Integration of the density profiles
gen atoms along the y-direction and Figure 10(b) in a y-range corresponding to the (bulk) kerogen den-
shows a snapshot demonstrating the adsorption of the sity of ∼1.25 g/cm3 (see Table 2), i.e. from 19 to 64 Å,
1100 L. MICHALEC AND M. LÍSAL

Figure . (a) The multiscale kerogen model that consisted of the rigid kerogen matrix with micropores corresponding to the dummy
particle with σLJ = 11 Å and a .-nm-wide slit cavity; density ρ = . g/cm , accessible surface area SA =  m /g, pore volume Vpore =
. cm /g, pore limiting diameter Dmin = 22.5 Å, and maximum pore size Dmax = 27.3 Å. The red, blue, yellow, grey and white spheres
represent oxygen, nitrogen, sulphur, carbon and hydrogen atoms, respectively. (b) The geometric pore size distribution, PSD, as a function
of the pore diameter for the multiscale kerogen model. (The colour ranges refer to the online version of the paper).

  
CH4 , nC2 H6 , nC3 H8
resulted in nmatrix matrix matrix
= {0.81, 0.29, 0.27} cavity cavity cavity
ρCH4 , ρC2 H6 , ρC3 H8 = {3.98, 1.00, 1.12}1/nm3 and
mmol/g. The value of nmatrix CH4 agrees within statisti- cavity
cal uncertainties with the previous GCMC result (see compared ρi with the bulk densities obtained 
Table 4) while the values of nmatrix matrix from NPT MC: ρPT,CH4 , ρPT,C2 H6 , ρPT,C3 H8 =
C2 H6 and nC3 H8 are only
slightly higher {4.593 , 0.671 , 0.341 }1/nm3 . From this comparison, we
than the GCMC results  in Table 4. By sub- cavity cavity
can see that ρCH4 < ρPT,CH4 while ρC2 H6 > ρPT,C2 H6 and
tracting nmatrix
CH4 , nmatrix
C2 H6 , nmatrix
C3 H8 = {0.81, 0.29, 0.27} cavity
mmol/g from ni , we evaluated the average densi- ρC3 H8 > ρPT,C3 H8 , indicating, similarly as was the case
ties of methane, ethane and propane in the cavity: in the microporous kerogen structures, the preferential
MOLECULAR PHYSICS 1101

Figure . (a) The number density profiles of methane, ethane and propane and mass density profile of kerogen atoms for the multiscale
kerogen model. The dash–dotted lines denote boundaries between the kerogen and cavity. (b) A snapshot demonstrating the adsorption
of the shale gas (i.e. a mixture of % of methane, % of ethane and % of propane) in the multiscale kerogen model. The red, blue and
green spheres represent methane, ethane and propane molecules, respectively. (The colour ranges refer to the online version of the
paper).

adsorption of propane and to a lesser extent ethane in width ranging from 4 to 13 Å, in agreement with the
the mesopores (cavity). experimental PSDs for shale [50,51] that indicate the exis-
tence of micropores larger than 10 Å. The densities of our
model kerogens were between 1.2 and 1.3 g/cm3 , in agree-
6. Conclusions
ment with the experimental data [49] on such kerogen at
In this work, we combined the existing atomistic model the same maturity level.
of a kerogen unit [17] and experimental data, and gen- We then used GCMC simulations and studied the
erated generic structures of overmature type II kerogen adsorption of pure methane, mixture of methane, ethane
corresponding to the organic part of Barnett shale. The and propane as a proxy of shale gas, and pure carbon
kerogen structures contained micropores with a pore dioxide in the model rigid kerogen structures. The
1102 L. MICHALEC AND M. LÍSAL

adsorbed amount of CO2 was about 30% higher than [2] A. Yethiraj and A. Striolo, J. Phys. Chem. Lett. 4, 687
that of CH4 (in agreement with the experimental data (2013).
[52]) due to stronger van der Waals and electrostatic gas– [3] B.P. Tissot and D.H. Welte, Petroleum Formation and
Occurrence (Springer: Berlin, 1984).
gas and gas–adsorbent interactions in the case of the [4] M.F. Brigatti, E. Galan, and B.K.G. Theng, Developments
CO2 adsorption. The structures further showed preferen- in Clay Science (Elsevier: Amsterdam, 2006).
tial adsorption for propane and to a lesser extent ethane [5] D.W. van Krevelen, Coal: Typology, Chemistry, Physics,
which, in the case of propane, can be associated with con- Constitution (Elsevier: Amsterdam, 1961).
densation of propane molecules in larger kerogen micro- [6] J.S. Seewald, Nature 426, 327 (2003).
[7] S.R. Kelemen, M. Afeworki, M.L. Gorbaty, M. Sansone,
pores. In the case of CH4 adsorption, we were able to
P.J. Kwiatek, C.C. Walters, H. Freund, M. Siskin, A.E.
compare the simulated excess adsorbed amount with the Bence, D.J. Curry, M. Solum, R.J. Pugmire, M. Vanden-
experimental data for the Barnett shale [53,54] and found broucke, M. Leblond, and F. Behar, Energy Fuels 21, 1548
that our simulation values were within the range of the (2007).
experimental values. [8] M. Siskin, C.G. Scouten, K.D. Rose, D. Aczel, S.G. Col-
We further employed MD simulations with a set- grove, and R.E. Pabst, Composition, Geochemistry and
Conversion of Oil Shales (Kluwer Academic: Boston, MA,
up that mimicked the real adsorption experiment and 1995).
explored connectivity of pore network in the model [9] S.R. Kelemen and M. Siskin, Prepr. Am. Chem. Soc. Div.
rigid kerogen structures to the external surface. The MD Pet. Chem. 49, 73 (2004).
simulations revealed limiting pore space accessibility of [10] A.M. Orendt, I.S.O. Pimienta, S.R. Badu, M.S. Solum,
certain kerogen pores which is related to extremely low R.J. Pugmire, J.C. Facelli, D.R. Locke, K.W. Chapman,
P.J. Chupas, and R.E. Winans, Energy Fuels 27, 702
permeability of real shales [1,2]. The kerogen pore net-
(2013).
work accessibility was enhanced when we considered the [11] C. Bousige, C.M. Ghimbeu, C. Vix-Guterl, A.E. Pomer-
flexibility of the kerogen structures in MD simulations. antz, A. Suleimenova, G. Vaughan, G. Garbarino, M. Fey-
Finally, we considered a multiscale kerogen model by genson, C. Wildgruber, F.-J. Ulm, R.J.-M. Pellenq, and B.
combining our rigid kerogen matrix with a 14-Å-wide Coasne, Nat. Mater. 15, 576 (2016).
cavity. The structure contained micropores as well as [12] C. Bousige, A. Boţan, F.-J. Ulm, R.J.-M. Pellenq, and B.
Coasne, J. Chem. Phys. 142, 114112 (2015).
mesopores with the pore width ranging from 4 to 27 Å. [13] A.C. van Duin, S. Dasgupta, F. Lorant, and W.A. Goddard,
We studied adsorption of the shale gas in the multiscale J. Phys. Chem. A 105, 9396 (2001).
kerogen model by GCMC simulations and found prefer- [14] B. Ni, K.-H. Lee, and S.B. Sinnott, J. Phys. Cond. Mater.
ential adsorption of propane and to a lesser extent ethane 16, 7261 (2004).
in both the micropores and mesopores of the structure. [15] J. Collell, G. Galliero, F. Gouth, F. Montel, M. Pujol,
P. Ungerer, and M. Yiannourakou, Micropor. Mesopor.
Mater. 197, 194 (2014).
[16] J. Collell, P. Ungerer, G. Galliero, M. Yiannourakou, F.
Acknowledgment Montel, and M. Pujol, Energy Fuels 28, 7457 (2014).
The computer time was provided by the METACentrum com- [17] P. Ungerer, J. Collell, and M. Yiannourakou, Energy Fuels
puting facility. 29, 91 (2015).
[18] M.P. Allen and D.J. Tildesley, Computer Simulation of Liq-
uids (Clarendon Press: Oxford, 1987).
[19] D. Frenkel and B. Smit, Understanding Molecular Simula-
Disclosure statement tion: From Algorithms to Applications (Elsevier: Amster-
No potential conflict of interest was reported by the authors. dam, 2002).
[20] M. Yiannourakou, P. Ungerer, B. Leblanc, X. Rozanska, P.
Saxe, S. Vidal-Gilbert, F. Gouth, and F. Montel, Oil Gas
Funding Sci. Technol. Rev. IFP 68, 977 (2013).
[21] A. Boţan, R. Vermorel, F.-J. Ulm, and R.J.-M. Pellenq,
This work has received funding from the European Union’s
Langmuir 29, 9985 (2013).
Horizon 2020 research and innovation programme [grant num-
[22] K. Falk, R. Pellenq, F.-J. Ulm, and B. Coasne, Energy Fuels
ber 640979], and was also supported by the Czech Science
29, 7889 (2015).
Foundation (Project No. 16-12291S) and the Internal Grant
[23] J. Collell, G. Galliero, R. Vermorel, P. Ungerer, M. Yian-
Agency of J. E. Purkinje University (Project No. 5322 16
nourakou, F. Montel, and M. Pujol, J. Phys. Chem. C 119,
0003 01).
22587 (2015).
[24] K. Falk, B. Coasne, R. Pellenq, F.-J. Ulm, and L. Bocquet,
Nat. Commun. 6, 6949 (2015).
References [25] H. Sui and J. Yao, J. Nat. Gas Sci. Eng. 31, 738 (2016).
[26] T.A. Ho, L.J. Criscenti, and Y. Wang, Sci. Rep. 6, 28053
[1] A. Striolo, F. Klaessig, D. R. Cole, J. Wilcox, G. G. Chase,
(2016).
C. H. Sondergeld, and M. Pasquali, Workshop Report,
[27] D.J.K. Ross and R.M. Bustin, Fuel 86, 2696 (2007).
NSF Grant Number CBET-1229931, 2012.
MOLECULAR PHYSICS 1103

[28] K. Mosher, J. He, Y. Liu, E. Rupp, and J. Wilcox, Int. J. Coal [42] B. Chen and J.I. Siepmann, J. Phys. Chem. B 103, 5370
Geol. Geology 109–110, 36 (2013). (1999).
[29] T. Li and C. Wu, Energy Fuels 29, 634 (2015). [43] J.J. Potoff and J.I. Siepmann, AIChE J. 47, 1676 (2001).
[30] W.L. Jorgensen, D.S. Maxwell, and J. Tirado-Rives, J. Am. [44] R.W. Hockney and J.W. Eastwood, Computer Simulation
Chem. Soc. 118, 11225 (1996). Using Particles (Adam Hilger, Bristol, 1989).
[31] J.G. Harris and K.H. Yung, J. Phys. Chem. 99, 12021 [45] J. Kolafa and J.W. Perram, Molec. Simul. 9, 351 (1992).
(1995). [46] L. Sarkisov and A. Harrison, Molec. Simul. 37, 1248
[32] R.T. Cygan, V.N. Romanov, and E.M. Myshakin, J. Phys. (2011).
Chem. C 116, 13079 (2012). [47] L.D. Gelb and K.E. Gubbins, Langmuir 15, 305 (1999).
[33] P. Dauber-Osguthorpe, V.A. Roberts, D.J. Osguthorpe, J. [48] NIST webbook: Thermophysical Properties of Fluid Sys-
Wolff, M. Genest, and A.T. Hagler, Proteins 4, 31 (1988). tems. <http://webbook.nist.gov/chemistry/fluid>.
[34] Maximizing the EU Shale Gas Potential by Minimiz- [49] K.S. Okiongbo, A.C. Aplin, and S.R. Larter, Energy Fuels
ing its Environmental Footprint (ShaleXenvironmenT). 19, 2495 (2005).
<https://shalexenvironment.org> [50] M.E. Curtis, R.J. Ambrose, D. Energy, C.H. Sondergeld,
[35] S.L. Montgomery, D.M. Jarvie, K.A. Bowker, and R.M. and C.S. Rai, Soc. Petr. Eng. 1, 137693 (2010).
Pollastro, AAPG Bull. 89, 155 (2005). [51] D.J.K. Ross and R.M. Bustin, Mar. Petr. Geol. 26, 916
[36] S.J. Plimpton, J. Comput. Phys. 117, 1 (1995). (2009).
[37] S. Nosé, J. Chem. Phys. 81, 511 (1984). [52] R. Heller and M. Zoback, J. Unconven. Oil Gas Resour. 8,
[38] W.G. Hoover, Phys. Rev. A 31, 1695 (1985). 14 (2014).
[39] S. Nosé and M.L. Klein, J. Chem. Phys. 78, 6928 (1983). [53] M. Gasparik, P. Bertier, Y. Gensterblum, A. Ghanizadeh,
[40] G.J. Martyna, D.J. Tobias, and M.L. Klein, J. Chem. Phys. B.M. Krooss, and R. Littke, Int. J. Coal Geol. Geol. 123,
101, 4177 (1994). 34 (2013).
[41] MCCCS Towhee version 7.1.0. <http: // towhee.source [54] T.W. Zhang, G.S. Ellis, S.C. Ruppel, K. Milliken, and R.S.
forge.net> (last accessed September 22, 2015). Yang, Org. Geochem. 47, 120 (2012).

Вам также может понравиться