Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Chemical Engineering Journal 411 (2021) 128423

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Dynamic fluid states in organic-inorganic nanocomposite: Implications for


shale gas recovery and CO2 sequestration
Liang Huang a, b, c, *, Wen Zhou a, b, Hao Xu a, Lu Wang b, Jie Zou b, Qiumei Zhou b
a
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Chengdu University of Technology, Chengdu, Sichuan 610059, PR China
b
College of Energy, Chengdu University of Technology, Chengdu, Sichuan 610059, PR China
c
Department of Chemical and Biomolecular Engineering, University of California, Berkeley, Berkeley, CA 94720-1462, USA

A R T I C L E I N F O A B S T R A C T

Keywords: Despite wide research on organic-inorganic composites in various fields, the molecular-scale partitioning of
Nanocomposite fluids in nanoporous organic-inorganic composites remains poorly understood. The microscopic characteristics of
Fluid state fluid states in nanopores of shale composites during gas production are key to the enhancement of shale gas
Shale gas recovery
recovery, which remain unexplored. Here we conduct a microscopic modeling for shale nanocomposite, which
CO2 sequestration
Molecular dynamics
realizes separate force field parameterization and controllable pore network optimization. A shale clay-kerogen
nanocomposite with heterogeneous pore structure is thereafter built to investigate the dynamic characteristics of
fluid states during pressure depletion and CO2 sequestration using molecular simulations for the first time. In this
study, the different gas recovery mechanisms for pressure depletion and CO2 sequestration are revealed: pressure
depletion mainly produces CH4 free state, while CO2 injection predominately extracts CH4 dissolved state. We
propose CH4 adsorbed state as the potential target for enhanced shale gas recovery in the late production stage
due to its minimal recovery. Three distribution forms of water molecules and two developing trends of water
clusters are distinguished in the nanocomposite. Increasing water content can reduce the CH4 recovery during
pressure depletion, but improve the CH4 recovery during CO2 injection. We suggest low water content may
optimize the shale gas recovery efficiency. The deformation of the kerogen structure is observed to enhance the
gas solubility in the kerogen matrix. In the CO2 injection stage, multi-components can promote the recovery of
the CH4 adsorbed state. In a broader perspective, the microscopic modeling is widely applicable to other natural
and synthetic organic-inorganic nanocomposites interacting with fluids.

1. Introduction states dominate the fluid states in shale nanopores [9–13]. Much
research has been conducted to characterize the two fluid states in shale
The sharp expansion of shale gas production, with the advancement nanopores [14–17]. A recent atomistic simulation study [18] discussed
of techniques such as horizontal well and hydraulic fracturing, has the change of the adsorbed and free states of CH4/CO2 in a kerogen
provoked the shale gas revolution [1–3]. However, the current recovery structure during a CO2 huff-n-puff process. However, the fluid states
factor of most shale gas reservoirs is less than 20% due to depletion were qualitatively analyzed, and the simulations were performed based
production [4]. CO2 sequestration in shale gas reservoir has tremendous on a fixed kerogen structure. By contrast, only limited research on the
potential for enhancing the shale gas recovery and mitigating the global dissolved state of fluid has been reported. Some authors have suggested
warming simultaneously, which has raised wide attention [5,6]. that the contribution of dissolved gas in kerogen can be up to 22% by
Of significant importance to shale gas recovery and CO2 sequestra­ experiment [7] and 30% by thermodynamic theory [8]. An extended
tion is the dynamic fluid state in nanoporous media of shale, which thermodynamic model has been developed to quantify the adsorption
determines the late development strategies in the field [7]. There are and dissolution of various hydrocarbons in swelling kerogen nanopores,
mainly three fluid states in shale nanoporous structure: free state in but it provided no discussion on the evolution of various gas states
various pores, adsorbed state on pore surfaces, and dissolved state in the during gas production [19]. The quantification of adsorbed, dissolved,
kerogen matrix [8]. It is commonly believed that free and adsorbed and free states of CH4/CO2 in shale nanopores during shale gas recovery

* Corresponding author.
E-mail address: huangliang@cdut.edu.cn (L. Huang).

https://doi.org/10.1016/j.cej.2021.128423
Received 19 October 2020; Received in revised form 24 December 2020; Accepted 31 December 2020
Available online 7 January 2021
1385-8947/© 2021 Elsevier B.V. All rights reserved.
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

and CO2 injection remains to be unraveled. conducted, which realizes separate force field parameterization and
Shale is characterized as a tight sedimentary rock, which is mainly controllable pore network optimization. We construct a shale organic-
composed of organic matters, clays, and other mineral components [20]. inorganic nanocomposite with heterogeneous pore structure, in which
A large number of fluid molecules are stored in shale organic matters the dynamic contributions of dissolved, adsorbed, and free state of CH4/
and clay minerals due to their abundant nanopores and large specific CO2 during shale gas recovery and CO2 injection are quantitatively
surface areas [21,22]. Extensive experiments have observed the pres­ distinguished for the first time. Furthermore, we explore the effect of
ence of intercalated clay-organic nanocomposites in immature shale water content, shale gas composition, and kerogen structure flexibility
source rocks [23–25]. In the nanocomposites, the amorphous organic on various fluid states in the nanocomposite. These results can help
matter is incorporated in the interlayer space of a sandwich-like clay understand the microscopic recovery process and adjust the production
structure. The organic matter and the clay possess some distinguished strategies in shale gas reservoirs. In a broader perspective, similar to
properties such as wettability, porous structure, and surface heteroge­ shale nanocomposite [20,48], there are also a great variety of other
neity, which aggravates the complexity of fluid states. Laboratory test on organic-inorganic nanocomposites, including the natural bio-
shale rocks is one of the most direct methods to integrate the combined nanocomposites such as bones [49] and seashells [50] and the syn­
effects of clay minerals and organic matters on fluid states, but it may thetic reinforced nanocomposites like clay-polymer materials [51].
not distinguish their contributions and gain insights into the underlying These nanocomposites possess promising applications in various fields
mechanisms. By contrast, molecular simulation is a powerful computa­ and have close interaction with fluids under changing circumstances.
tional tool to shed light on the microscopic conformation, interaction, Thus, results of this work are expected to provide inspiring implications
and fluid fractionation in clay-organic complex. These microscopic for these hybrid materials.
properties determine the macroscopic behaviors of fluids in shale
nanocomposite during production. Although a few attempts have been 2. Methodology
performed to reproduce the behavior of shale organic-inorganic nano­
composite at the molecular scale, these reported studies were mainly Nanocomposite model. In this work, a microscopic modeling process
focused on reinforced mechanical properties [26], non-bonded in­ is conducted to prepare the shale organic-inorganic nanocomposite. The
teractions [27,28], and kerogen extraction [29] in the nanocomposite. inorganic and organic phases are initially prepared separately with in­
Also, these reported shale nanocomposites had simplified porous net­ dividual force field parameterization, and thereafter integrated as the
works, and an integrated force field was used for parameterizing both shale nanocomposite. In the shale nanocomposite, the desired slot size is
the organic and inorganic parts. The dynamic fluid states in shale obtained by controlling the compressed strain of the inorganic layers,
nanocomposite with heterogeneous pores during gas production have while the porous network of the organic matrix is optimized by inserting
not been investigated at the molecular scale yet. designed dummy particles. The representative shale nanocomposite
An important feature of shale is the presence of residual water in used in this work is shown in Fig. 1. The nanocomposite is constituted by
heterogeneous nanopores, wherein clay minerals like montmorillonite a dense kerogen matrix sandwiched between two Na-montmorillonite
are deemed as hydrophilic [30], while organic matter, consisting of walls. The Na-montmorillonite is selected due to its abundance in
hydrophobic carbon skeleton and polar functional groups, is considered shale formations [27]. The clay sheet has a tetrahedral-octahedral-
as mix-wet [31]. Extensive experiments have validated the presence of tetrahedral (TOT) structure with isomorphic substitutions in both the
water in shale and concluded its effect on gas adsorption [32,33]. In our octahedral (O) and the tetrahedral (T) layers. The substitutions induce
prior work, molecular simulations have been performed to gain insights negative charges in the clay layers which are compensated with potas­
into the microscopic distribution of water and its effect mechanisms on sium ions in the interlayer spacing. The unit cell of Na-montmorillonite
CH4/CO2 adsorption in the organic matter [34–37] and clay mineral (chemical formula: Na0.75(Si7.75Al0.25)(Al3.5Mg0.5)O20(OH)4)) is char­
[38]. To our best knowledge, no molecular simulation study on water acterized as a three-dimension monoclinic-B lattice type with the typical
distribution mechanisms and the corresponding effect on CH4/CO2 cell lengths (a = 5.23 Å, b = 9.06 Å, and c = 9.60 Å) and cell angles (α =
states in shale nanocomposite has been reported. Another key issue γ = 90◦ , β = 99◦ ) [52]. The bulk clay model was constructed with 8 × 4
associated with CH4/CO2 states in shale nanopores is the effect of shale × 1 unit cells, resulting in a dimension of 41.84 Å×36.24 Å×9.6 Å along
gas compositions. In addition to the major component CH4, shale gas the x, y, and z directions, respectively. The atomic positions and
also contains various light hydrocarbons such as C2H6, C3H8, and C4H10 isomorphic substitutions were described in detail in our previous work
and some nonhydrocarbon constituents such as N2, CO2, and H2O. In our [38]. In the nanocomposite, the Na-montmorillonite wall was generated
previous work, we have characterized the microscopic distribution of by cleaving the most probable surface (0 0 1) with a thickness of 19.2 Å
various shale gas components in organic matter by molecular simula­ from the bulk clay model.
tions [39,40]. The effect mechanism of shale gas compositions on CH4/ For the organic phase, it was condensed based on three kerogen
CO2 states in shale nanocomposites remains elusive. Moreover, the groups, each consisting of three type IA kerogen units
ductile organic matter in shale possesses strong structural flexibility like (C251H385O13N7S3). Type IA kerogen, representative of immature
a soft nanoporous matter [41]. The interaction of organic matter with organic matter in Green River shale, was developed based on analytical
fluids can lead to the change of internal porous structure, which feeds data [53]. It possesses a complex structural skeleton, made up of
back on fluid distribution [42]. Plenty of work has studied the gas dominant aliphatic chains and a small amount of aromatic rings, and has
adsorption behaviors in flexible kerogen structure by poromechanical strong surface chemical heterogeneity due to various functional groups.
theory [43] and atomistic simulation [44]. In this context, we have In this study, we merge three kerogen units into one group with artificial
recently reported the sorption-swelling coupling in kerogen-fluid bonds, termed as joints in Fig. 1. The joints are designed to extend the
interaction by poromechanical theory [45,46], thermodynamic model, topology of kerogen macromolecule and keep the flexibility of kerogen
and experiment [47]. However, the effect of kerogen structure flexibility structure simultaneously. The chemical formula of the kerogen group is
on dynamic fluid states with the restriction of clay mineral remains to be C753H1151O39N21S9, which is close to that of reported Green River type I
clarified. kerogen [27]. Note the location of the joints may affect the topology and
Despite previous efforts in the literature, the dynamic characteristic chemistry of the kerogen structure, thus changing the structural prop­
of fluids states in shale organic-inorganic nanocomposite during gas erties and fluid behaviors. In our previous work [34,35], we have shown
production has not been quantified yet, and the effect mechanisms of the strong effects of structural topology and surface chemistry hetero­
water content, shale gas composition, and kerogen structure flexibility geneity of kerogen on gas sorption behaviors. To mitigate the effect of
on fluid states remain poorly understood. the joints, there are three criteria for determining the location of the
In this work, a microscopic modeling for shale nanocomposite is joints, namely topology keeping, chemistry keeping, and easy

2
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

Fig. 1. Proposed clay-organic nanocomposite representative of shale. The nanocomposite is composed of Na-montmorillonite sheets and kerogen groups. The model
size is 41.84 × 36.24 × 75 Å along the x , y, and z direction, respectively. The sizes of the clay walls and kerogen phase are reduced along the z direction during the
compression process.

connecting. Specifically, the joints may not be placed on a benzene ring, the dummy particles (30 Å for 2 particle and 20 Å for 2 particles) are
on a functional group, or in the middle of a aliphatic chain. The dense optimized to approach a reported porous structure that can reasonably
amorphous kerogen matrix is combined with the Na-montmorillonite represent the sorption behavior of the studied adsorbate species [54].
walls to represent the shale nanocomposite. This kind of combination The shale nanocomposite at equilibrium state has a dimension of 41.84
is common in shale formations [23–25] and has been adopted in pre­ Å × 36.24 Å × 75 Å along the x, y, and z directions, respectively.
vious simulation studies [27,28]. To expand the pore sizes in the shale Simulation procedure. To construct the nanocomposite model, the
nanocomposite, a series of dummy particles with various LJ diameters clay and organic phases are initially prepared separately. The Na-
are inserted to generate heterogeneous pores. The sizes and amount of montmorillonite walls are obtained by cutting the bulk clay model

3
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

along the cleavage surface with a certain thickness. The bulk clay model nanocomposite with multiple fluid components at the equilibrium state.
is modified in thickness from our previous work [38]. For the organic In this process, the clay sheet and the kerogen matrix are kept fixed,
phase, three kerogen groups after geometry optimization are randomly while the fluid components are flexible.
incorporated into a large cuboid box (41.84 Å × 36.24 Å × 250 Å along Simulation details. Molecular simulations are implemented using the
the x, y, and z direction). Four dummy particles are then put into the box Large-Scale Atomic/Molecular Massively Parallel Simulator (LAMMPS)
by randomly. Layer building is subsequently performed to generate the [61]. The atomic types and point charges of kerogen are parameterized
initial nanocomposite structure by sandwiching the organic box be­ with the CVFF force field [62], while these of Na-montmorillonite are
tween the two clay walls. The nanocomposite system is thereafter described by the CLAYFF force field [63]. The bonding interactions in
annealed by molecular dynamics (MD) simulations in the NVT ensemble the CVFF force field include bond, angle, dihedral, and improper terms.
(constant number of molecules, box volume, and temperature), in which By contrast, only the H-O bond term is considered for the bonding
temperature is reduced from 900 K, 700 K, 500 K to 300 K in a stepwise interaction in the CLAYFF force field. CH4 molecule is represented with
manner (100 ps at each temperature). The rigorous annealing procedure the TraPPE United-Atom model [64], CO2 molecule is treated by the
is aimed to bring the structure of the nanocomposite to its equilibrated TraPPE-EH model [65], H2O molecule is described using the SPC/E
state. The similar annealing procedures have been widely utilized to model [66], and other shale gas compositions are parameterized with
relax the kerogen structures and have been validated to efficiently the CVFF force field. Nonbonded interactions between atomic pairs are
overcome the energy barriers of kerogen structures calculated by the Lennard-Jones (LJ) 12-6 function and the unscreened
[34,35,40,41,53,54]. During the annealing procedure, the clay walls are Coulomb potential. Lorentz-Berthelot combining rules [67] are adopted
kept rigid while the kerogen structures as well as potassium ions are for unlike atoms. The cutoff distance for short-range pairwise interaction
flexible. To obtain the nanocomposite with the desired size, the is 14 Å, and tail corrections [68] are employed. The long-range elec­
annealed system is compressed every 100 steps along the z direction trostatic interaction is described by the Ewald summation [68,69]. Pe­
within the NVT ensemble [55]. The box dimension in the z direction riodic boundary conditions are implemented along the x and y direction,
decreases from 250 Å to 75 Å after the compression treatment in 200 ps; while the nanocomposite along the z direction is non-periodic. In the
the box dimension is fixed along the x and y-direction. The similar simulation, the temperature at 338 K is controlled by the Nosé-Hoover
strategy has been conducted to compress the kerogen and muscovite thermostat [70], and the timestep is set as 0.1 fs. Gas adsorption is
system [55]. During the compression, the size of clay walls and kerogen performed in the grand canonical ensemble by the hybrid GCMC-MD
phase were allowed to change along the z-direction, which is aimed to method, wherein gas in the nanocomposite model at equilibrium state
simulate the deformation of inorganic and organic matters under a has the same chemical potential and temperature with an external bulk
certain compression stress [56,57]. Next, the compressed nano­ reservoir. Chemical potential is required as input in the adsorption
composite model is further relaxed in the NVT ensemble for 200 ps. simulation. In this study, we perform the hybrid GCMC-MD simulation
The wet nanocomposite models are built by randomly preloading for fluids in the bulk phase to obtain the relationship between pressure
various water molecules in the dry nanocomposite model. The number and chemical potential. The hybrid GCMC-MD simulation has a total of
of water molecules is determined based on our previous work on clay 4000 cycles, each including 2500 GCMC exchanges after every 10,000
[38] and kerogen [34]. During the simulation, water molecules are MD steps. During the simulation, the adsorption amount, temperature,
allowed to move, while the number of them is fixed, as observed in and energy terms are monitored to guide the equilibrium.
previous simulation work [58]. MD simulations in the NVT ensemble are In this work, only physical interactions were taken into account in
performed for 1000 ps to bring the water molecules to the equilibrium the nanocomposite and fluids system since our major objective is to
state. Gas adsorption is conducted with the hybrid GCMC-MD method, in construct a representative intercalated clay-organic composite in shale,
which the grand canonical Monte Carlo (GCMC) algorithm for gas ex­ and thereafter to investigate the distribution of water and the mecha­
change and MD algorithm for structure relaxation are integrated. More nisms of shale gas recovery and CO2 sequestration. The adopted force
details about the hybrid method can be found in Ref. [59]. In this study, fields and fluid models have been widely reported to well describe the
we refer to the reported recovery process in Ref. [18], which includes structural characteristics and fluid behaviors. The initially assigned
initial pressure depletion, subsequent CO2 injection, and late pressure force field parameters remained unchanged in the subsequent course of
depletion. Details of the recovery process are presented in Table 1. In MD simulations, since the presence of fluid molecules has little effect on
this process, the first two stages correspond to single CH4 adsorption; the the assigned force field parameters of nanocomposite with respect to
late two stages correspond to CH4/CO2 competitive adsorption. physical interactions.
The shale nanocomposite with multiple fluid components is con­ Computational analysis. Fluid uptakes, pore structures, energy
structed by integrating water molecules and shale gas compositions. The terms, and fluid states are computed based on post-processing of the
shale gas compositions refer to the realistic shale gas compositions in nanocomposite trajectories. Fluid uptakes are directly obtained by
Eagle Ford [60]. Initially, 192 water molecules, 330 CH4 molecules, and counting fluid molecules in the nanocomposite model. Pore structures
corresponding shale gas compositions are randomly loaded into the including porosity and pore size distribution (PSD) are characterized
kerogen matrix of the nanocomposite. Then MD simulations in the NVT using the Metropolis Monte Carlo integration method based on geome­
ensemble are performed for 1000 ps to bring the fluid components to the try determination [71]. The intra-kerogen energy terms are calculated
equilibrium state. Subsequently, the hybrid GCMC-MD method is by first removing the clay phase. The interaction energy is obtained by
adopted to perform the recovery process in Table 1 based on the shale subtracting the energy of isolated components from the total energy of
the mixtures. Fluid states are divided into free, adsorbed, and dissolved
Table 1
states. The free state is obtained by performing separate simulations
Details of fluids in an external bulk reservoir during the recovery process at 338 under bulk fluid conditions. The dissolved state is defined as the number
K. of fluid molecules in the confining kerogen matrix [19]. The sizes of
pores in the kerogen matrix are smaller than 10 Å. The adsorbed state
Stage Stage Pressure Bulk fluid
symbol (atm) composition refers to the excess adsorption, which is further determined by sub­
tracting the free and dissolved states from the total uptake. To eliminate
Initial gas reservoir I 200 Pure CH4
After first pressure II 96 Pure CH4
the uncertainty, fluid uptakes are averaged among 500 structure con­
depletion figurations; pore structures, energy terms, and fluid states are deter­
After CO2 injection III 113 80% CH4, 20% CO2 mined from 10 structure configurations.
After second pressure IV 96 80% CH4, 20% CO2
depletion

4
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

3. Results and discussion between 8 and 10 Å disappears on the nanocomposite PSD. This implies
that the induced matrix pores join the slit pores and are extended to be
3.1. Nanocomposite characterization larger pores. The maximum pore size is 25 Å for the nanocomposite, and
extensive pores have sizes varying between 18 and 25 Å. The simulated
Fig. 2a shows the pore structure of shale nanocomposite. The porous results are compared with an experimental measurement of kerogen PSD
network is composed of slit pore, induced matrix pores, and natural by CO2 and N2 adsorption experiments [72]. The measured kerogen PSD
matrix pores. The kerogen matrix is brought in the proximity of the clay is approximately close to the simulated PSD, containing substantial
walls, forming a slit pore in the nanocomposite. The slit surfaces, pores with sizes in the range of 18–25 Å and 4–10 Å.
constituted by kerogen units, possess strong physical and chemical The aspect ratio between the pore size and the adsorbate size has
heterogeneity. In the inner kerogen matrix, the irregular arrangement of significant effects on the sorption behavior of adsorbate molecules [47].
kerogen structures results in abundant natural matrix pores; the inser­ In our previous work [40], we have studied the representation of a
tion and removal of dummy particles produce induced matrix pores. The confining kerogen matrix (pore size <6 Å) for describing the sorption
induced matrix pores act as bridges that connect the natural matrix behavior of CO2 and various light hydrocarbons by molecular simula­
pores and the slit pore. tions. We found the confining kerogen matrix can capture the sorption
Fig. 2b presents the pore size distribution (PSD) of kerogen matrix behavior of CO2 and CH4, with the simulated sorption isotherms similar
and shale nanocomposite. In addition to natural matrix pores, the to the measured results. However, the simulated isotherms for C2H6,
kerogen matrix also contains some induced matrix pores. In our previous C3H8, and C4H10 are in poor agreement with the experimental data,
work [34,35], the maximum pore size of the kerogen matrix having indicating the pores in the kerogen matrix may need to be expanded to
natural pores alone is smaller than 6 Å. In Fig. 2b, the induced matrix obtain reasonable results for these species. Subsequently, Tesson and
pores by dummy particles extend the maximum matrix pore to approach Firoozabadi [54] investigated the sorption of CO2, CH4, C2H6, C3H8,
10 Å. The PSD segment of the kerogen matrix is roughly coincident with C4H10, and C5H12 in deformable kerogen structures by molecular sim­
that of the nanocomposite with the pore size smaller than 8 Å. However, ulations. The maximum pore size of their kerogen structures was
the PSD segment of the kerogen matrix with the pore size ranging expanded to be around 24 Å. They found the range of pore size was

Fig. 2. Porous structure and energy origin of shale nanocomposite. (a) Pore distribution (shown in maroon). (b) Pore size distribution (PSD). The pore is probed by
helium (diameter = 2.6 Å) in the simulations. The experimental PSD was obtained from CO2 and N2 adsorption measurements [72]. (c) Interaction energy between
kerogen and clay components. (d) Intra-kerogen energy in shale nanocomposite.

5
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

adequate to represent the sorption behavior of CO2 and these light hy­ interaction is ignorable. The negative nonbonded energy implies that
drocarbons by comparing the simulated and measured results. In our the mutual attraction between kerogen molecules leads to the conden­
work, the maximum pore size of the nanocomposite is about 25 Å, and sation of the kerogen matrix. The attractive nonbonded interaction is
substantial pores possess sizes varying between 18 and 25 Å. For the concordant with the reported results for interaction within different
main adsorbate species, the dynamic diameter for H2O, CO2, CH4, C2H6, kerogen fragments [27]. By contrast, the intramolecular energy is pos­
C3H8 and C4H10 is 2.65, 3.3, 3.8, 4.0, 4.2, 4.3 Å, respectively. Given the itive, indicating there is mutual repulsion between kerogen segments in
reported investigations, we believe the nanocomposite can reasonably the same molecular skeleton. The positive intramolecular energy can
represent the sorption behavior of these adsorbate species. extend and swell the kerogen matrix. The total energy of intramolecular
Fig. 2c shows the nonbonded interaction energy between different interaction is four times more than that of the nonbonded interaction,
components in the nanocomposite. The negative energy indicates which suggests that the intramolecular interaction is the major inter­
mutual attraction, while the positive energy indicates mutual repulsion. action form in the kerogen counterparts. Various energy items in the
There is a mutual attraction between kerogen groups and clay sheets intramolecular interaction of kerogen matrix are in the order of bond-
since their interaction energy is negative, which is consistent with the stretching energy > angle-bending energy > dihedral-distortion en­
observation in Fig. 1 that the kerogen groups are adsorbed on the clay ergy > improper-vibration energy.
surfaces. In addition to the electrostatic interaction, the VDW interaction
also has a significant contribution to the kerogen-clay interaction, which
3.2. Distribution characteristics of water molecules
is accordant with the previous report for clay interaction with various
kerogen fragments [27]. The kerogen groups and the sodium ions repel
Fig. 3 shows the snapshots of water distribution under increasing
each other due to positive interaction energy with dominant electro­
water content. In the nanocomposite with 32 water molecules, water
static interaction and ignorable VDW interaction. The clay sheets are
molecules exist in three forms: some water molecules are adsorbed on
negatively charged, while the sodium ions are positively charged. Thus
the clay surface in a cluster form; some water molecules occupy the slit
the sodium ions are apt to occupy the clay surfaces. On the other hand,
pore in a cluster form; a small part of water molecules are adsorbed
kerogen groups also tend to be attracted by the clay sheets, which leads
around the polar functional groups of kerogen in a dispersed form. With
to the competitive adsorption of kerogen groups and sodium ions on the
the water content increasing to 64 molecules per unit cell, the water
clay surfaces. Compared with the kerogen groups, the sodium ions are
cluster in the slit pore merges with the water cluster on the clay surface.
much closer to the clay surfaces indicating stronger attractive interac­
Also, there is a small water cluster adsorbed around the polar functional
tion (see Fig. 1). Fig. 2c also shows that the sodium ions reduce the
groups of kerogen. With the further increase of water content, the water
attractive energy between the kerogen groups and the clay sheets,
cluster on the clay surface evolves in two trends: on one hand, the water
especially the electrostatic energy. This observation suggests that so­
cluster extends along the clay surface, gradually peeling off the kerogen
dium ions have the potential to remove kerogens adsorbed on clay
molecules adsorbed on the clay surface; on the other hand, the water
surfaces, providing theoretical implications for kerogen extraction from
cluster develops perpendicular to the clay surface, extending to the
oil shale.
kerogen matrix and blocking part of the seepage channels. Also, there
Fig. 2d presents the intra-kerogen energy components of the nano­
are still a small number of water molecules scatteringly adsorbed around
composite, which are composed of nonbonded energy and intra­
the polar functional groups of kerogen.
molecular energy. Given the equilibrium of kerogen counterpart and the
The distribution characteristics of the water molecule are closely
mitigated effect of joints, the intra-kerogen energy components are
correlated with the interplay between water and clay/kerogen and the
determined by the innate structures of kerogen units including the to­
interaction between water molecules. The clay surfaces are negatively
pology and the surface chemistry, which are mainly correlated with the
charged due to atomic substitutions, which facilitates the preferential
organic types and maturities of kerogen [34,35]. The VDW interaction
adsorption of water molecules on the clay surfaces without being
energy dominates in the nonbonded energy, while the electrostatic
covered by kerogen matrix and sodium ions. There is a strong attractive

Fig. 3. Water distribution under different water contents. Water molecules are shown in red and gray balls. (a) With 32 water molecules. (b) With 64 water
molecules. (c) With 128 water molecules. (d) With 192 water molecules. (For interpretation of the references to color in this figure legend, the reader is referred to
the web version of this article.)

6
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

interaction between water molecules because of hydrogen bonding, (Fig. 4b). With the further increase of simulation time (Fig. 4c), the
which promotes the aggregation of water molecules into clusters. The adjacent small water clusters mix together. The upper three small
kerogen matrix in the nanocomposite possesses mixed wettability: the clusters merge into one big cluster adsorbed on the clay surface, while
carbon skeleton is hydrophobic, while the polar functional groups are the lower three small clusters aggregate into one big cluster in the
hydrophilic. Some water molecules are apt to scatter around these hy­ kerogen slit. With the simulation time increasing to 80 ps (Fig. 4d), the
drophilic functional groups. In our previous work [34,35], we have big water cluster in the kerogen slit joins the one on the clay surface,
observed the aggregated water clusters in nanopores and the scattering finally forming an exclusive water cluster in the nanocomposite. Fig. A1
water molecules around polar functional groups in the kerogen matrix. shows the snapshots of water distribution on the clay surface with
In this work, we propose the third form that water clusters are prefer­ increasing simulation time. As the simulation time evolves, the water
entially adsorbed on the uncovered clay surfaces with negative charges, cluster extends along the clay surface and gradually peels off the
which is the predominant distribution form of water molecules in shale adsorbed kerogen molecules.
clay-organic nanocomposite.
The dynamic water distribution characteristics reveal the implica­
3.3. Dynamic fluid states during CH4 recovery and CO2 sequestration
tions for the two-fold role of water in CO2 sequestration. The water
distribution is governed by the energetic effect from the interactions of
Fig. 5a shows the CH4 recovery during the gas recovery process (see
water/sorption sites at low water content and the entropy effect from the
Table 1) with different water contents. Pressure depletion (I–II) is per­
interactions of water/water molecules at high water content. The en­
formed in the initial production stage, the CH4 recovery in the dry
ergetic effect of water mainly affects the CO2 stability, while the entropy
nanocomposite is 28.6% as the system pressure decreases from 200 to
effect of water mainly affect the CO2 capacity. As the water content
96 atm. To improve CH4 recovery, CO2 sequestration is conducted in the
increases (from Fig. 3a to b), some water molecules initially adsorbed on
shale nanocomposite. CO2 sequestration has two main mechanisms for
the kerogen matrix transport and merge with the water cluster on the
improving CH4 recovery, namely the competitive adsorption effect and
clay surface due to strong hydrogen bonding. The transposition of water
the pressure changing effect. On one hand, CO2, having low surface
molecules releases some occupied polar functional groups of kerogen,
tension and high adsorption affinity, can replace CH4 from nanopores to
which can potentially increase the CO2 stability. The observation sug­
seepage channels by occupying adsorption sites. Note the occupation of
gests that although water can reduce the CO2 capacity, it may improve
these adsorption sites by CO2 can make the nanocomposite more hy­
the CO2 stability, thus there is an optimized water content that balancing
drophobic. On the other hand, the injection of CO2 can increase the
CO2 capacity and CO2 stability. The finding has significant implications
system pressure, providing additional energy for the second pressure
for CO2 sequestration in shale gas reservoirs. It indicates that the CO2
depletion, in which the partial pressure of CH4 is reduced. As CO2 is
stability can be optimized by operating the water content in the nano­
injected (II–III), CH4 recovery increases by 11.9% and the system pres­
pores of shale rocks.
sure rises to 113 atm. Then the second pressure depletion production is
Fig. 4 presents the evolution of water distribution in the nano­
carried out to bring the pressure back to 96 atm (III–IV), which further
composite with increasing simulation time. In the initial stage (Fig. 4a),
increases CH4 recovery by 4.1%. Overall, CO2 sequestration can pro­
water molecules randomly scatter in the nanocomposite: most water
mote CH4 recovery in the dry nanocomposite from 28.6% to 44.7%,
molecules exist in the slit pore, while a small part of water molecules
among which the CH4 recovery improved by the competitive adsorption
dissolve in the kerogen matrix. With the increase of simulation time,
effect is around three times more than that by the pressure changing
water molecules aggregate into six small clusters, one of them is
effect. Thus CO2 replacement induced by the competitive adsorption is
adsorbed on the clay surface, while the other five clusters are distributed
the main mechanism for improving CH4 recovery during CO2 seques­
around the polar functional groups on the slit surface of kerogen
tration. With the increase of water content in Fig. 5a, CH4 recovery

Fig. 4. Dynamic evolution of water molecules (192 molecules) at 1 ps (a), 7 ps (b), 12 ps (c), and 80 ps (d).

7
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

Fig. 5. CH4 recovery (a), gas amount (b), composition of CH4 states (c), and CO2 states (d) in shale nanocomposite during the gas recovery process with different
water contents.

decreases in pressure depletion stages (I–II and III–IV), and increases in contributions of various gas states to gas uptake are shown in Fig. A2. At
CO2 injection stage (II–III). Compared with the dry nanocomposite and the initial gas reservoir stage, the contribution of adsorbed CH4 is 20.2%
the one with high water content, the nanocomposite with low water in the dry shale nanocomposite; while that of dissolved CH4 (40.5%) and
content has the maximal total CH4 recovery (I–IV). This observation can free CH4 (39.3%) are close to each other, which are the major CH4 states
provide a significant implication for CO2 sequestration with enhanced in the dry nanocomposite. After pressure depletion, dissolved CH4
CH4 recovery. (45.1–46.2%) becomes the dominant CH4 state in the dry nano­
Fig. 5b presents the gas amount during the gas recovery process with composite; the contribution of free CH4 (25.6–27.1%) is close to that of
different water contents. The total gas amount after CO2 injection is adsorbed CH4 (27.6–28.2%). In the dry nanocomposite, the injected CO2
larger than the CH4 amount before CO2 injection. The ratio of CH4 mainly exists in the form of dissolved state (~50%), the adsorbed and
replacement by CO2 is around 1:2, namely two CO2 molecules replace free CO2 each takes up ~25%. With the increase of water content, the
one CH4 molecule. Both the amount of CH4 and CO2 decrease in the contribution of adsorbed CH4 remains basically the same; the contri­
second pressure depletion stage. In the dry nanocomposite, the amount bution of dissolved CH4 decreases slightly, while that of free CH4 in­
of produced CH4 is close to that of CO2, while in the wet nanocomposite, creases slightly. The contributions of various CO2 states remained
the amount of produced CO2 is more than that of CH4. This observation basically unchanged under the studied water contents, which indicated
implies that water molecules in the nanocomposite can facilitate the that both the adsorbed, dissolved, and free states of CO2 decreased with
production of CO2, which is similar to our previous conclusion in the wet a similar rate. The decrease of both CO2 states is related to the water
kerogen matrix [34]. This is because water molecules can occupy the distribution characteristics in Section 3.2. The water clusters developed
polar functional groups (see Fig. 3), which are potentially preferential on the clay surface can reduce the dissolved state of CO2, the water
adsorption sites for CO2 [34]. The strong affinity of CO2 with these molecules adsorbed around the polar functional groups can reduce the
functional groups can be attributed to the strong non-bonded interaction adsorbed state of CO2, and the water clusters aggregating in the kerogen
including both the VDW and electrostatic interaction. slit can reduce the free state of CO2. The similar decreasing rate of the
Fig. 5c, d shows the CH4 and CO2 states in the shale nanocomposite three states suggests that water has resembling partition characteristics
during the gas recovery process with different water contents. The with CO2 in the dissolved, adsorbed, and free regions of the

8
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

nanocomposite. 3.4. Effect of shale gas components and kerogen structure flexibility
The evolution of CH4 and CO2 density distribution in the dry nano­
composite is shown in Figs. 6a and A3, respectively. Combined with Fig. 8a shows the density profile of shale gas components (CH4
Fig. 5c, d, it can be observed that the free CH4 shows a drastic reduction excluded) along the z direction. The distribution of shale gas compo­
during the first pressure depletion stage, while the dissolved CH4 is the nents (CH4 excluded) in the shale nanocomposite during the gas re­
dominant produced state during the CO2 injection stage. In the second covery process is presented in Fig. A4. The shale gas components (CH4
pressure depletion stage, both the CH4/CO2 dissolved state, free state, excluded) exist mainly in the forms of dissolved state and adsorbed state
and adsorbed state decrease slightly, and the relative proportion of in the nanocomposite. Affected by the matrix pore size and the molec­
various gas states remains basically unchanged. Fig. 6b shows the CH4 ular size of each component, most of the components remain in the
density distribution in the nanocomposite under various water contents. ultraconfining kerogen matrix in the form of dissolved state. Some
Combined with Fig. 5c, it can be seen that the increasing water content components diffuse into the kerogen slit from the kerogen matrix pores
reduces both the dissolved, adsorbed, and free CH4 in the nano­ but are still adsorbed on the surfaces of the kerogen slit. At different
composite, which can be explained by the water distribution charac­ production stages, the density peak sites of shale gas components in the
teristics in Section 3.2. Some water molecules form into clusters on the dissolved region of the shale nanocomposite remain relatively fixed, and
clay surface and the cluster develops perpendicular to the clay surface, the change of density peak heights is small. This observation indicates
extending to the kerogen matrix, which can reduce the dissolved CH4; that the dissolved state of shale gas components in the kerogen matrix
some water molecules are adsorbed around the polar functional groups changes slightly during the recovery process. By contrast, the density
on the surface of the kerogen slit, which can reduce the adsorbed CH4; peaks of adsorbed shale gas components in the kerogen slit show
some water molecules aggregate into clusters in the kerogen slit, which noticeable change during the recovery process. The shale gas compo­
can reduce the free CH4. nents exchange between the two adsorption layers on the surfaces of the
Fig. 7 presents the total recovery of various gas states in shale kerogen slit.
nanocomposite during the whole gas recovery process under different Fig. 8b, c shows the CH4 and CO2 states in the shale nanocomposite
water contents. In the dry nanocomposite, CH4 free state has the during the gas recovery process. Wherein symbol A and B denote the
maximal recovery at 62.1%, followed by CH4 dissolved state with a shale nanocomposite with/without shale gas components, respectively.
recovery of 37.8%. The recovery of the CH4 adsorbed state is only For both case A and case B, the clay sheet and the kerogen matrix are
24.3%. Compared with CH4 free and dissolved states, CH4 adsorbed kept fixed. Symbol C also denotes the shale nanocomposite without
state, attracted by the surfaces of nanopores, is more difficult to be shale gas components. For case C, the clay sheet is fixed, but the kerogen
extracted. Thus CH4 adsorbed state is the potential target for enhanced matrix is flexible. The effect of shale gas components on gas production
gas recovery in the shale reservoir. During the production after the in­ can be studied by comparing case A and case B. In the first pressure
jection of CO2, the recoveries of various CO2 states are close to each depletion (I–II) and CO2 injection (II–III) stages, shale gas components
other. Compared with the dry nanocomposite, the recovery of the CH4 can reduce the amount of CH4 dissolved state by 50%, since most of
adsorbed state is smaller, while that of the CO2 adsorbed state is bigger these components dissolve in the kerogen matrix. After the second
in the wet nanocomposite. Moreover, the nanocomposite with the lower pressure depletion (IV), the CH4 adsorbed state is reduced by ~50%. For
water content has the maximal recovery of CH4 dissolved state. The CO2 states, shale gas components show a small effect on CO2 free and
effect of water content can be attributed to water distribution charac­ dissolved states, but a noticeable effect on CO2 adsorbed state. Shale gas
teristics. At the lower water content, some water molecules can occupy components are mainly in the forms of dissolved and adsorbed states,
the preferential adsorption sites for CO2 on the surfaces of kerogen slit, thus exerting a small effect on CO2 free state. As system pressure de­
weakening the interplay between CO2 and kerogen slit surfaces. Also, creases, massive CH4 dissolved state is produced from the kerogen ma­
the reduction of accessible adsorption sites for CO2 compels more CO2 trix, which provides additional spaces for CO2 dissolved state. On the
molecules to compete with CH4 molecules in the dissolved region, thus surfaces of the kerogen slit, shale gas components compete with CO2,
boosting the recovery of CH4 dissolved state. reducing the amount of CO2 adsorbed state. In the late production stage,
shale gas components have a bigger effect on CO2 adsorption, the

Fig. 6. CH4 density distribution in the nanocomposite. (a) Evolution during the gas recovery process under dry condition. (b) Effect of water content at 202 atm. The
dashed lines represent the bulk density distributions; the black lines indicate the boundaries of dissolved regions. The densities are averaged among 100
configurations.

9
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

Fig. 7. Recovery of various gas states in shale nanocomposite during the gas recovery process with different water contents. (a) CH4; (b) CO2.

Fig. 8. (a) Density distribution of shale gas components (CH4 excluded), composition of CH4 states (b) and CO2 states (c), and CH4 recovery (d) in nanocomposite
during the gas recovery process.

10
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

stability of sequestrated CO2 can be greatly weakened. By comparing the work reported in this paper.
case B and case C, the effect of kerogen structure flexibility on gas states
can be investigated. Both the amount of CH4 and CO2 dissolved states Acknowledgments
can be greatly increased by kerogen structure flexibility. Compared with
the gas dissolved state, the effect of kerogen structure flexibility on the This work was supported by the National Natural Science Foundation
gas free and adsorbed states is relatively small. of China (Grant Nos. 51774298, 41972137 and 42002157) and the In­
Fig. 8d presents the CH4 recovery in the shale nanocomposite during dependent project of the State Key Laboratory of Oil and Gas Reservoir
the gas recovery process. In the pressure depletion stage, CH4 free state Geology and Exploitation (Grant No. CDL2019001). Computer time was
has the maximal recovery at ~53%, followed by CH4 dissolved state at provided by the computing clusters of the College of Chemistry at the
18.1–27.7%. The recovery of the CH4 adsorbed state is negative, which University of California, Berkeley. We appreciate the support of the
is because the massive reduction of CH4 free state increases the relative China Scholarship Council (CSC No. 201806440095).
amount of CH4 adsorbed state. In the CO2 injection stage, the recovery of
the CH4 adsorbed state ranks the first, followed by the dissolved state,
Appendix A. Supplementary data
then the free state. By comparing case A and case B, shale gas compo­
nents improve the recovery of CH4 dissolved state during pressure
Supplementary data to this article can be found online at https://doi.
depletion and that of CH4 adsorbed state during CO2 injection. By
org/10.1016/j.cej.2021.128423.
comparing case B and case C, kerogen structure flexibility can reduce the
recovery of CH4 dissolved state during pressure depletion. In the CO2
References
injection stage, the effect of kerogen structure flexibility on the CH4
recovery in various states is insignificant. As shown in Fig. 8b, c, kerogen [1] R.A. Kerr, Natural gas from shale bursts onto the scene, Nature 328 (2010)
structure flexibility has little effect on the free and adsorbed states of 1624–1626.
CH4 and CO2, thus resulting in a neglected change of CH4 recovery in the [2] R.D. Vidic, S.L. Brantley, J.M. Vandenbossche, D. Yoxtheimer, J.D. Abad, Impact of
shale gas development on regional water quality, Science 340 (6134) (2013)
free and adsorbed states during CO2 injection. By contrast, both the
1235009.
dissolved states of CH4 and CO2 can be greatly increased by kerogen [3] L. Cueto-Felgueroso, R. Juanes, Forecasting long-term gas production from shale,
structure flexibility. During pressure depletion, the CH4 recovery in the Proc. Natl. Acad. Sci. U.S.A. 110 (49) (2013) 19660–19661.
[4] V. Kuuskraa, S.H. Stevens, K.D. Moodhe, Technically Recoverable Shale Oil and
dissolved state is reduced due to the increasing dissolved state of CH4. As
Shale Gas Resources: An Assessment of 137 Shale Formations in 41 Countries
CO2 is injected, kerogen structure flexibility facilitates more CO2 mol­ Outside the United States, US Energy Information Administration, US Department
ecules to access the dissolved region and replace the CH4 molecules, thus of Energy, 2013.
increasing the CH4 recovery in the dissolved state and making it com­ [5] R.S. Middleton, J.W. Carey, R.P. Currier, et al., Shale gas and non-aqueous
fracturing fluids: opportunities and challenges for supercritical CO2, Appl. Energy
parable to that in the fixed nanocomposite. 147 (2015) 500–509.
[6] J. Liu, L. Xie, Y. Yao, et al., Preliminary study of influence factors and estimation
4. Conclusions model of the enhanced gas recovery stimulated by carbon dioxide utilization in
shale, ACS Sustainable Chem. Eng. 7 (24) (2019) 20114–20125.
[7] S.R. Etminan, F. Javadpour, B.B. Maini, et al., Measurement of gas storage
The characteristics of fluid states during pressure depletion and CO2 processes in shale and of the molecular diffusion coefficient in kerogen, Int. J. Coal
sequestration in the shale gas reservoir are clarified by molecular sim­ Geol. 123 (2014) 10–19.
[8] Z. Jin, A. Firoozabadi, Thermodynamic modeling of phase behavior in shale media,
ulations within a shale clay-organic nanocomposite. The effects of water SPE J. 21 (01) (2016) 190–207.
content, shale gas components, and kerogen structure flexibility are [9] F. Chen, S. Lu, X. Ding, et al., Evaluation of the density and thickness of adsorbed
discussed. CO2 sequestration can not only replace CH4 but also elevate methane in differently sized pores contributed by various components in a shale
gas reservoir: a case study of the Longmaxi Shale in Southeast Chongqing, China,
the system pressure, wherein the improved CH4 recovery by replace­ Chem. Eng. J. 367 (2019) 123–138.
ment (11.9%) is around three times more than that by pressure elevation [10] D. Chai, G. Yang, Z. Fan, et al., Gas transport in shale matrix coupling multilayer
(4.1%). CH4 free state shows a drastic reduction during pressure adsorption and pore confinement effect, Chem. Eng. J. 370 (2019) 1534–1549.
[11] P. Chareonsuppanimit, S.A. Mohammad, R.L. Robinson Jr., K.A.M. Gasem, High-
depletion, while CH4 dissolved state is the dominant produced state
pressure adsorption of gases on shales: measurements and modeling, Int. J. Coal
during CO2 injection. In the late production stage, CH4 adsorbed state Geol. 95 (2012) 34–46.
has minimal recovery, which is the potential target for enhanced gas [12] Z. Sun, X. Li, W. Liu, et al., Molecular dynamics of methane flow behavior through
recovery in the shale reservoir. Water molecules exist in three forms in realistic organic nanopores under geologic shale condition: pore size and kerogen
types, Chem. Eng. J. 398 (2020) 124341.
the nanocomposite: being adsorbed on the clay surface in a cluster form; [13] T. Zhang, G.S. Ellis, S.C. Ruppel, et al., Effect of organic-matter type and thermal
occupying the slit pore in a cluster form; being adsorbed around polar maturity on methane adsorption in shale-gas systems, Org. Geochem. 47 (2012)
functional groups of kerogen in a dispersed form. Increasing water 120–131.
[14] M. Zhang, S. Zhan, Z. Jin, Recovery mechanisms of hydrocarbon mixtures in
content can reduce the CH4 recovery during pressure depletion, but organic and inorganic nanopores during pressure drawdown and CO2 injection
improve the CH4 recovery during CO2 injection. In the CO2 injection from molecular perspectives, Chem. Eng. J. 382 (2020), 122808.
stage, a low water content can increase the recovery of CH4 dissolved [15] J. Zhou, Z. Jin, K.H. Luo, Insights into recovery of multi-component shale gas by
CO2 injection: a molecular perspective, Fuel 267 (2020), 117247.
state, shale gas components can promote the recovery of CH4 adsorbed [16] J. Li, Z. Chen, K. Wu, et al., A multi-site model to determine supercritical methane
state, and kerogen structure flexibility can massively boost the dissolu­ adsorption in energetically heterogeneous shales, Chem. Eng. J. 349 (2018)
tion of CH4 and CO2. 438–455.
[17] S. Wang, Q. Feng, F. Javadpour, et al., Competitive adsorption of methane and
All the results in this work were based on physical interactions in the
ethane in montmorillonite nanopores of shale at supercritical conditions: a grand
nanocomposite and fluid system. In shale gas reservoirs, there are also canonical Monte Carlo simulation study, Chem. Eng. J. 355 (2019) 76–90.
complicated chemical interactions such as reactions between CO2 and [18] J. Zhou, Z. Jin, K.H. Luo, Effects of moisture contents on shale gas recovery and
CO2 sequestration, Langmuir 35 (26) (2019) 8716–8725.
water, carbonic acid and shale, inorganic and organic interface, which
[19] J. Liu, W.G. Chapman, Thermodynamic modeling of the equilibrium partitioning of
may affect the fluid behaviors. To represent more closely the realistic hydrocarbons in nanoporous kerogen particles, Energy Fuels 33 (2) (2019)
fluid behavior in shale gas reservoir, the chemical interactions may be 891–904.
incorporated to get further insight in a future publication. [20] H.M.N. Faisal, K.S. Katti, D.R. Katti, Modeling the behavior of organic kerogen in
the proximity of calcite mineral by molecular dynamics simulations, Energy Fuels
34 (3) (2020) 2849–2860.
Declaration of Competing Interest [21] C. Bousige, C.M. Ghimbeu, C. Vix-Guterl, et al., Realistic molecular model of
kerogen’s nanostructure, Nat. Mater. 15 (5) (2016) 576–582.
[22] X. Hu, H. Deng, C. Lu, et al., Characterization of CO2/CH4 competitive adsorption
The authors declare that they have no known competing financial in various clay minerals in relation to shale gas recovery from molecular
interests or personal relationships that could have appeared to influence simulation, Energy Fuel 33 (9) (2019) 8202–8214.

11
L. Huang et al. Chemical Engineering Journal 411 (2021) 128423

[23] J. Berthonneau, O. Grauby, M. Abuhaikal, et al., Evolution of organo-clay [47] L. Huang, A. Khoshnood, A. Firoozabadi, Swelling of Kimmeridge kerogen by
composites with respect to thermal maturity in type II organic-rich source rocks, normal-alkanes, naphthenes and aromatics, Fuel 267 (2020), 117155.
Geochim. Cosmochim. Acta 195 (2016) 68–83. [48] G. Hantal, L. Brochard, M.N. Dias Soeiro Cordeiro, et al., Surface chemistry and
[24] K.N. Alstadt, K.S. Katti, D.R. Katti, Nanoscale morphology of kerogen and in situ atomic-scale reconstruction of kerogen–silica composites, J. Phys. Chem. C 118 (5)
nanomechanical properties of green river oil shale, J. Nanomech. Micromech. 6 (1) (2014) 2429–2438.
(2016) 04015003. [49] S.M. Pradhan, K.S. Katti, D.R. Katti, Multiscale model of collagen fibril in bone:
[25] K.N. Alstadt, D.R. Katti, K.S. Katti, An in situ FTIR step-scan photoacoustic elastic response, J. Eng. Mech. 140 (2014) 454–461.
investigation of kerogen and minerals in oil shale, Spectrochim. Acta Part A Mol. [50] P. Ghosh, D.R. Katti, K.S. Katti, Mineral proximity influences mechanical response
Biomol. Spectrosc. 89 (2012) 105–113. of proteins in biological mineral− protein hybrid systems, Biomacromolecules 8 (3)
[26] G. Hantal, L. Brochard, R.J.M. Pellenq, et al., Role of interfaces in elasticity and (2007) 851–856.
failure of clay-organic nanocomposites: toughening upon interface weakening? [51] D. Sikdar, S.M. Pradhan, D.R. Katti, et al., Altered phase model for polymer clay
Langmuir 33 (42) (2017) 11457–11466. nanocomposites, Langmuir 24 (2008) 5599–5607.
[27] D.R. Katti, K.B. Thapa, K.S. Katti, Modeling molecular interactions of sodium [52] N.T. Skipper, F.R.C. Chang, G. Sposito, Monte Carlo simulations of interlayer
montmorillonite clay with 3D kerogen models, Fuel 199 (2017) 641–652. molecular structure in swelling clay minerals. 1. Methodology, Clays Clay Miner.
[28] D.R. Katti, H.B. Upadhyay, K.S. Katti, Molecular interactions of kerogen moieties 43 (3) (1995) 285–293.
with Na-montmorillonite: an experimental and modeling study, Fuel 130 (2014) [53] P. Ungerer, J. Collell, M. Yiannourakou, Molecular modeling of the volumetric and
34–45. thermodynamic properties of kerogen: influence of organic type and maturity,
[29] T. Wu, Q. Xue, X. Li, et al., Extraction of kerogen from oil shale with supercritical Energy Fuel 29 (1) (2014) 91–105.
carbon dioxide: molecular dynamics simulations, J. Supercrit. Fluids 107 (2016) [54] S. Tesson, A. Firoozabadi, Deformation and swelling of kerogen matrix in light
499–506. hydrocarbons and carbon dioxide, J. Phys. Chem. C 123 (48) (2019) 29173–29183.
[30] Z. Jin, A. Firoozabadi, Effect of water on methane and carbon dioxide sorption in [55] T.A. Ho, Y. Wang, A. Ilgen, et al., Supercritical CO2-induced atomistic lubrication
clay minerals by Monte Carlo simulations, Fluid Phase Equilib. 382 (2014) 10–20. for water flow in a rough hydrophilic nanochannel, Nanoscale 10 (42) (2018)
[31] Y. Hu, D. Devegowda, A. Striolo, et al., Microscopic dynamics of water and 19957–19963.
hydrocarbon in shale-kerogen pores of potentially mixed wettability, SPE J. 20 [56] T. Wu, A. Firoozabadi, Mechanical properties and failure envelope of kerogen
(2014) 112–124. matrix by molecular dynamics simulations, J. Phys. Chem. C 124 (4) (2020)
[32] J. Li, X. Li, K. Wu, et al., Water sorption and distribution characteristics in clay and 2289–2294.
shale: effect of surface force, Energy Fuel 30 (11) (2016) 8863–8874. [57] T. Wu, A. Firoozabadi, Fracture toughness and surface energy density of kerogen
[33] J. Li, X. Li, X. Wang, et al., Water distribution characteristic and effect on methane by molecular dynamics simulations in tensile failure, J. Phys. Chem. C 124 (29)
adsorption capacity in shale clay, Int. J. Coal Geol. 159 (2016) 135–154. (2020) 15895–15901.
[34] L. Huang, Z. Ning, Q. Wang, et al., Effect of organic type and moisture on CO2/CH4 [58] P. Billemont, B. Coasne, G.D. Weireld, Adsorption of carbon dioxide, methane, and
competitive adsorption in kerogen with implications for CO2 sequestration and their mixtures in porous carbons: effect of surface chemistry, water content, and
enhanced CH4 recovery, Appl. Energy 210 (2018) 28–43. pore disorder, Langmuir 27 (3) (2011) 1015–1024.
[35] L. Huang, Z. Ning, Q. Wang, et al., Molecular simulation of adsorption behaviors of [59] M. Chen, B. Coasne, R. Guyer, et al., Role of hydrogen bonding in hysteresis
methane, carbon dioxide and their mixtures on kerogen: effect of kerogen maturity observed in sorption-induced swelling of soft nanoporous polymers, Nat. Commun.
and moisture content, Fuel 211 (2018) 159–172. 9 (1) (2018) 3507.
[36] L. Huang, Z. Ning, Q. Wang, et al., Enhanced gas recovery by CO2 sequestration in [60] D.L. George, E.B. Bowles, Shale gas measurement and associated issues, Pipeline
marine shale: a molecular view based on realistic kerogen model, Arab. J. Geosci. Gas J. 238 (7) (2011) 38–41.
11 (15) (2018) 404. [61] S. Plimpton, Fast parallel algorithms for short-range molecular dynamics,
[37] L. Huang, Z. Ning, Q. Wang, et al., Molecular simulation of CO2 sequestration and J. Comput. Phys. 117 (1) (1995) 1–19.
enhanced gas recovery in gas rich shale: an insight based on realistic kerogen [62] A.T. Hagler, S. Lifson, P. Dauber, Consistent force field studies of intermolecular
model, in: SPE Abu Dhabi International Petroleum Exhibition & Conference, forces in hydrogen-bonded crystals. 2. A benchmark for the objective comparison
Society of Petroleum Engineers, Abu Dhabi, UAE, Nov. 13–16, 2017, SPE-188216. of alternative force fields, J. Am. Chem. Soc. 101 (18) (1979) 5122–5130.
[38] Q. Wang, L. Huang, Molecular insight into competitive adsorption of methane and [63] R.T. Cygan, J.-J. Liang, A.G. Kalinichev, Molecular models of hydroxide,
carbon dioxide in montmorillonite: effect of clay structure and water content, Fuel oxyhydroxide, and clay phases and the development of a general force field,
239 (2019) 32–43. J. Phys. Chem. B 108 (4) (2004) 1255–1266.
[39] L. Huang, Z. Ning, Q. Wang, et al., Thermodynamic and structural characterization [64] M.G. Martin, J.I. Siepmann, Transferable potentials for phase equilibria. 1. United-
of bulk organic matter in Chinese Silurian Shale: experimental and molecular atom description of n-alkanes, J. Phys. Chem. B 102 (1998) 2569–2577.
modeling studies, Energy Fuel 31 (5) (2017) 4851–4865. [65] J.J. Potoff, J.I. Siepmann, Vapor–liquid equilibria of mixtures containing alkanes,
[40] L. Huang, Z. Ning, Q. Wang, et al., Microstructure and adsorption properties of carbon dioxide, and nitrogen, AIChE J. 47 (7) (2001) 1676–1682.
organic matter in Chinese Cambrian gas shale: experimental characterization, [66] H.J.C. Berendsen, J.R. Grigera, T.P. Straatsma, The missing term in effective pair
molecular modeling and molecular simulation, Int. J. Coal Geol. 198 (2018) 14–28. potentials, J. Phys. Chem. 91 (24) (1987) 6269–6271.
[41] S. Tesson, A. Firoozabadi, Methane adsorption and self-diffusion in shale kerogen [67] H.A. Lorentz, Ueber die Anwendung des Satzes vom Virial in der kinetischen
and slit nanopores by molecular simulations, J. Phys. Chem. C 122 (41) (2018) Theorie der Gase, Ann. Phys. 248 (1) (1881) 127–136.
23528–23542. [68] D. Frenkel, B. Smit, Understanding Molecular Simulation: From Algorithms to
[42] T. Wu, H. Zhao, S. Tesson, et al., Absolute adsorption of light hydrocarbons and Applications, 2002.
carbon dioxide in shale rock and isolated kerogen, Fuel 235 (2019) 855–867. [69] J.C. Shelley, G.N. Patey, Boundary condition effects in simulations of water
[43] H. Sui, J. Yao, Effect of surface chemistry for CH4/CO2 adsorption in kerogen: a confined between planar walls, Mol. Phys. 88 (2) (1996) 385–398.
molecular simulation study, J. Nat. Gas Sci. Eng. 31 (2016) 738–746. [70] S. Nosé, A unified formulation of the constant temperature molecular dynamics
[44] T.A. Ho, Y. Wang, L.J. Criscenti, Chemo-mechanical coupling in kerogen gas methods, J. Chem. Phys. 81 (1) (1984) 511–519.
adsorption/desorption, Phys. Chem. Chem. Phys. 20 (18) (2018) 12390–12395. [71] L.D. Gelb, K.E. Gubbins, Pore size distributions in porous glasses: a computer
[45] L. Huang, Z. Ning, Q. Wang, et al., Molecular insights into kerogen deformation simulation study, Langmuir 15 (2) (1999) 305–308.
induced by CO2/CH4 sorption: effect of maturity and moisture, Energy Fuel 33 (6) [72] G. Chen, S. Lu, K. Liu, Investigation of pore size effects on adsorption behavior of
(2019) 4792–4805. shale gas, Mar. Pet. Geol. 109 (2019) 1–8.
[46] L. Huang, Z. Ning, Q. Wang, et al., Kerogen deformation upon CO2/CH4
competitive sorption: implications for CO2 sequestration and enhanced CH4
recovery, J. Petrol. Sci. Eng. 183 (2019), 106460.

12

You might also like