You are on page 1of 17

Chemical Engineering Journal 425 (2021) 130292

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Molecular insights into carbon dioxide enhanced multi-component shale


gas recovery and its sequestration in realistic kerogen
Sen Wang a, b, *, Xinyu Yao a, b, Qihong Feng a, b, Farzam Javadpour c, Yuxuan Yang a, b,
Qingzhong Xue d, Xiaofang Li d
a
School of Petroleum Engineering, China University of Petroleum (East China), Qingdao 266580, China
b
Key Laboratory of Unconventional Oil & Gas Development, Ministry of Education, Qingdao 266580, China
c
Bureau of Economic Geology, Jackson School of Geosciences, The University of Texas at Austin, University Station, Box X, Austin, TX 78713, USA
d
State Key Laboratory of Heavy Oil Processing, School of Material Science & Engineering, China University of Petroleum (East China), Qingdao 266580, China

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

Keywords: Understanding the governing processes of CO2 huff-n-puff in shales is essential for enhancing shale gas recovery
Shale gas and CO2 sequestration. However, existing studies have not fully accounted for the chemical composition and
Adsorption geometry of kerogen nanopores and the reality that natural gas is a multi-component mixture. We used grand
Kerogen
canonical Monte Carlo (GCMC) simulations to study the competitive adsorption between CO2 and typical hy­
CO2 sequestration
Molecular simulation
drocarbon components (CH4, C2H6, and C3H8). We further studied the recovery mechanisms of CO2 huff-n-puff
within kerogenic circular nanopores at reservoir conditions. We probed the effects of pressure, pore geometry,
and size on gas recovery and CO2 sequestration efficiency. Although pressure drop readily exploits CH4 in the
adsorption layer, the recovery due to CO2 injection primarily occurs within the kerogen matrix for pure CH4.
Injecting CO2 facilitates the recovery of heavier hydrocarbons, whereas pressure drawdown exhibits better
performance for lighter components. CO2 huff-n-puff may serve as a promising method for gas exploitation in
circular pores, whereas pressure drop favors the production in kerogen slit. The tremendously different gas
adsorption and recovery behavior in distinct pore geometries and compositions necessitate the study using
realistic shale kerogen models. Moreover, enlarging the pore size improves the recovery of each component
during pressure drawdown but restrains the performance of CO2 injection; meanwhile, the total gas recovery and
CO2 sequestration efficiency increase. This study provides a more in-depth understanding of multi-component
gas recovery mechanisms within realistic shale kerogen nanopores and sheds light on the CO2 sequestration in
shale reservoirs.

1. Introduction the shale gas recovery using advanced technologies has been an urgent
issue to be solved [10,11].
Shale gas has emerged as an important alternative clean energy Injecting CO2 into shale formations (through huff-n-puff) is recog­
source to conventional fossil fuels because of its low emissions and high nized as a promising solution because it enhances shale gas productivity
energy efficiency [1-4]. Recently, the exploration and development of and mitigates global warming via geological CO2 sequestration [12-16].
shale gas have attracted global focus. As reported by the Energy Infor­ CO2 huff-n-puff generally contains three procedures: CO2 injection
mation Administration (EIA), gas production in the Eagle Ford shale of (huff), soaking, and pressure drawdown (puff). There are recent studies
the United States has surged from 0.119 billion cubic feet per day (Bcf/ on the transport of gas mixtures in nanomaterials, such as carbon
d) in 2010 to 4.405 Bcf/d in 2020 [5]. Meanwhile, China, Australia, and nanotube (CNT) and metal–organic framework (MOF) [17-22]. How­
several European countries have started strong programs to exploit hy­ ever, the underlying mechanisms are different in amorphous nano­
drocarbons in shale [6]. The permeability of shale matrix is 6 to 9 orders porous kerogen that motivates a thorough study of the CO2 huff-n-puff in
of magnitude smaller than that of conventional reservoirs, leading to the shales.
rapid decrease of well production rates [7-9]. Therefore, how to improve Some experimental studies have been reported on the competitive

* Corresponding author.
E-mail address: fwforest@gmail.com (S. Wang).

https://doi.org/10.1016/j.cej.2021.130292
Received 20 February 2021; Received in revised form 4 May 2021; Accepted 7 May 2021
Available online 21 May 2021
1385-8947/© 2021 Elsevier B.V. All rights reserved.
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

adsorption of CH4 and CO2 in shales [23,24]. Using a manometric setup,


Khosrokhavar et al. [23] found the preferential adsorption of CO2 over
CH4 on a carboniferous shale from Belgium. The experimental results of
Charoensuppanimit et al. [24] on Woodford and Caney shales samples
suggested that the ratio of adsorption amount between CO2 and CH4
may vary from 2.1 to 8.6 depending on the total organic carbon (TOC).
Note that the produced gas from shale formations contains heavier hy­
drocarbon components, e.g., up to 20% molar fraction of C2H6 and C3H8
[25]. Hence, the adsorption of multi-component alkanes in shale is
becoming a focus of recent studies [26-30]. Cheng and Huang [26]
measured the adsorption selectivity of C1–C6 components in clays, coal,
and shale at pressures below 0.3 MPa. Kong et al. [27] used the ther­
mogravimetric method to study the excess adsorption/desorption iso­
therms of C1 and C2 on shales and estimated the absolute adsorption at
different temperatures. The sorption of light hydrocarbons (C1-C4) in
shale samples and isolated kerogens has also been explored by Fir­
oozabadi and his coworkers [29,30]. However, owing to the heteroge­
neous compositions of shale samples, it is complicated—through
experimental measurements—to reveal the molecular interactions be­
tween hydrocarbon mixtures and CO2 under confinement and elucidate
Fig. 1. Scanning electron microscopy (SEM) image showing the morphology of
the underlying mechanisms of CO2 injection. Moreover, the pressure and
organic nanopores in shale. The experimental sample is adopted from the
temperature of a realistic shale reservoir are too high [31], which re­ Longmaxi shale of Sichuan Province, China. The black and gray areas represent
quires more sophisticated equipment to tackling the challenge. the organic matter (OM) and inorganic minerals, respectively. The darkest spots
Therefore, molecular simulations, such as grand canonical Monte in the OM correspond to organic nanopores.
Carlo (GCMC) and molecular dynamics (MD), have been utilized as a
viable and powerful tool to provide deeper insights into the gas
adsorption behavior at the nanoscale [32-34]. The results can not only Table 1
complement the experimental analyses but also be beneficial for the Comparison of the molecular models used in previous literatures and this work.
equipment design. For example, on the basis of Configurational-biased No. References Surface Fluid Simulation Limitations
GCMC and MD, Falk et al. [33] probed the effect of the molecular method
length (from C1 to C12) and pore accessibility on the adsorption of n- 1 Zhang et al. Pure CH4, C2H6, GCMC Cannot
alkane in kerogen matrix. Tesson and Firoozabadi [32] created kerogen [41] carbon C3H8, and account for the
slits having different surface roughness and analyzed their effect on slit, K- CO2 influences of
Illite slit heteroatoms in
methane adsorption and self-diffusion. Using MD, Wu et al. [34] studied
2 Wu et al. Pure CH4 and MD kerogen
the recovery process of methane and ethane mixtures in nanopores [34] carbon C2H6
under a pressure gradient drive. They suggested that ethane shows su­ slits
perior adsorption capacity over methane, and this phenomenon is more 3 Ho et al. Kerogen CH4, CO2, MD
pronounced in narrow pores. Recently, Wang et al. [35] simulated the [40] matrix, H2O
CNTs,
competitive adsorption of CH4 and C2H6 in clay nanopores and reported Kerogen
that the favorable adsorption of C2H6 fails at elevated pressures. slit
However, most of the existing studies only concentrate on the 4 Kazemi Kerogen CH4, C2H6, GCMC, MD Pores within
adsorption of a pure-component alkane or its competitive adsorption et al. [61] matrix H2O, and the kerogen
CO2 matrix is too
with CO2 in shales. In contrast, the studies on the interactions between
5 Tesson et al. Kerogen CH4 GCMC small to
hydrocarbon mixtures and CO2 are scarcely reported [36-45]. Huang [32] matrix, represent a
et al. [36,37] simulated the competitive adsorption between CH4 and Kerogen realistic shale
CO2 on dry and moist kerogen matrix and highlighted the more slit kerogen, and
remarkable impact of moisture on CO2. Similar results were also re­ 6 Huang et al. Kerogen CH4, CO2, GCMC the organic
[36,37] matrix H2O shale pores are
ported by Jin et al. [38]. GCMC simulations were performed by Vasi­
7 Vasileiadis Kerogen CH4, C2H6, GCMC usually
leiadis et al. [39] to study the adsorption of a quaternary mixture (C1, C2, et al. [39] matrix n-C4H10 and circular
CO2, and N2) in the kerogen matrix at various temperatures and pres­ CO2 instead of slit-
sures. They suggested that the confinement effect plays a dominant role 8 Yang et al. Kerogen Multi- MD shaped.
[62] slit component
in the mixture adsorption. Some other simulations have been conducted
shale oil
to study CO2 sequestration and enhanced gas recovery at the nanoscale. 9 Zhou et al. Kerogen CH4, C2H6, GCMC
Ho et al. [40] used carbon nanotubes (CNTs) to mimic the kerogen [38,42] slit C3H8, and
nanopore and studied the retention and release of CH4 and CO2. They CO2
reported that CO2 could diffuse through the water layer in the kerogen 10 Sun et al. Circular CH4 MD Only studied
[63] kerogen the adsorption
matrix and facilitate methane desorption. Zhang et al. [41] used pure-
nanopore and transport
carbon and K-illite to construct the organic and inorganic slits, respec­ of CH4 within
tively, to probe the recovery mechanisms of multi-component shale gas organic
during CO2 injection. Zhou et al. [42] extended this study to kerogen nanopores
11 This work Circular CH4, C2H6, GCMC /
slits and observed a more favorable performance of enhanced gas re­
kerogen C3H8, and
covery in narrower slits. nanopore CO2
Even though several advances have been achieved to understand the
recovery mechanisms of CO2 huff-n-puff and sequestration in shale gas
reservoirs from the molecular perspective, there still exists three unre­
solved problems. First, the pore diameter (<1.4 nm) of the kerogen

2
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 2. (a) Molecular model of the II-D kerogen unit; (b) equilibrium configuration of the kerogen-based circular nanopores. Color code: black, C; gray, H; red, O;
blue, N; yellow, S. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

matrix models, which were widely employed to study shale gas


Table 2
adsorption and transport [46-50], is much smaller than that of realistic
NVT/NPT relaxation procedure to create the kerogen matrix.
kerogen (2–100 nm). Considering the tremendous impact of pore size on
gas storage and diffusion, more realistic kerogen models are required to No. Ensemble T (K) Time (ps) P (MPa)

clarify the actual behavior of shale gas. Second, recent studies on the 1 NVT 900 400 –-
shale pore structure characterization via advanced techniques, such as 2 NPT 900 200 20
3 NPT 700 200 20
atom force microscopy (AFM) [51] and scanning electron microscope
4 NPT 500 200 20
(SEM) [52], suggested that the pores located within inorganic minerals 5 NPT 300 400 20
of shales are primarily slit-shaped [53,54]. In contrast, the organic pores
tend to exhibit circular cross-sectional geometry owing to the thermal
conversion of kerogen into hydrocarbons [55], as shown in Fig. 1. the circular organic nanopore (Fig. 2). This type of kerogen corresponds
However, most of the relevant studies focused on kerogen slit to typical unconventional reserves with high potential in producing
[38,41,56]. A multi-layer adsorption model developed by Qajar et al. shale gas such as the Barnett shale [66]. The detailed composition and
[57] suggested that shale was better characterized using a network of structural parameters of the kerogen unit match fairly well with the
interconnected cylindrical pores. Moreover, because of the greater experimental data of solid-state 13C nuclear magnetic resonance (NMR)
sorption energy and the larger surface area, cylindrical kerogen pore and X-ray photoelectron spectroscopy [64]. Besides, II-D kerogen has
dominates the adsorption capacity of shale gas [58], which necessitates been employed to study the competitive adsorption of CO2 and CH4 in
the study in kerogen pores having circular geometry. The last problem shales [32,38,40].
lies in the influence of heteroatoms. Carbon nanotubes (CNT) have been The technique proposed by Collell et al. [67] is utilized to construct
used as the shale skeleton without accounting for the heteroatoms (i.e., the molecular model of kerogen matrix and then cut out a cylindrical
Oxygen, Nitrogen, and Sulphur) in kerogen [40,59,60]. To manifest the pore. This process is achieved through a series of MD simulations in the
differences of this work from current literature, we summarize the NVT (canonical) and NPT (isobaric-isothermal) ensembles [37,65].
characteristics of the molecular models used in relevant studies in First, we randomly put 72 relaxed kerogen units (a total of 21,024
Table 1. atoms) into a box with an initial density of 0.1 g/cm3 [67,68], and
Combining MD and GCMC simulations, we studied the recovery performed 400 ps NVT simulations at 900 K to relax the structure. Then,
mechanisms of pure CH4 and hydrocarbon mixture (C1/C2/C3) confined we conducted NPT simulations with four stepwise decreasing tempera­
in kerogen-based circular nanopores through CO2 huff-n-puff and pres­ tures from 900 K to 700 K, 500 K and eventually 300 K at 20 MPa. For
sure drawdown. We discuss the effects of pressure, pore geometry, and each step, the duration of our simulation is long enough to reach the
pore size on hydrocarbon recovery and CO2 sequestration efficiency. To equilibration.
the best of our knowledge, this work is the first attempt to explore the Consequently, we obtained the stable molecular structures of the
physics of pressure drawdown and CO2 injection in shale kerogen with kerogen matrix. We summarized the whole simulation procedure in
circular nanopores and multi-component gas. Table 2 and showed the variation of kerogen density in Fig. 3. As the
temperature decreases, the kerogen is condensed and approaching the
2. Models and methods stratigraphic conditions. In Section 3, we will show that the density of
the kerogen matrix agrees well with the experimental measurements,
2.1. Construction of Kerogen-based circular nanopore which justify the validity of our model. Then we deleted those atoms
located within a cylinder having the predefined diameter and added
On the basis of the comprehensive experimental results from Kele­ hydrogens atoms to the dangling bonds to get the circular kerogen
men et al. [64], Ungerer et al. [65] developed molecular structures of nanopore, as shown in Fig. 2b.
kerogen units having distinct thermal maturities. We used the II-D
kerogen unit (chemical formula: C175H102O9N4S2) [65] to construct

3
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 3. Variation of the kerogen density during the construction of kerogen matrix. The snapshots at different time are also included to show the compression process.

2.2. Force field fixed [41]. After shale gas production, we recorded the composition of
confined fluid at the equilibrium state. In Fig. 4, we show the simulation
Three light hydrocarbons (CH4, C2H6, and C3H8) are taken into ac­ snapshots in 2-nm kerogen-based nanopores at different stages.
count because of their high proportions in shale gas reservoirs such as We estimate the total amount of shale gas molecules in the kerogen
Barnett and Marcellus [69]. According to the data from the Mississip­ nanopore at each simulation stage through GCMC [73,74]. Molecular
pian Barnett Shales [70], our hydrocarbon mixture is composed of C1, dynamics is not employed here because GCMC is much computationally
C2, and C3 with the bulk mole ratio of 80:15:5. A widely used all-atom cheaper to study the equilibrium properties, from which the ultimate gas
force field, COMPASS (Condensed-phase Optimized Molecular Poten­ recovery can be readily estimated [41]. MD is more favorable to simu­
tials for Atomistic Simulation Studies) [71], is employed to estimate the late the dynamic process and time-dependent properties because each
intramolecular interactions of kerogen and alkanes. This force field new configuration is obtained by solving Newton’s second law. How­
describes the dispersion-repulsion energy and the Coulombic in­ ever, GCMC generates an ensemble of configurations under specific
teractions through Lennard-Jones 6–9 potential and point charges. conditions (fugacity) through random perturbations, from which the
COMPASS has been successfully employed by Huang et al. [36] to study relevant thermophysical properties may be estimated although the time-
the competitive adsorption of CH4 and CO2 within kerogen matrix. We evolving information cannot be obtained [75]. Thus, the application of
maintain the temperature constant using the Nose/Hoover thermostat. A GCMC saves the computational resources required for gas adsorption
cutoff distance of 12.5 Å is used when estimating the van der Waals and simulation. Random insertion, translation, rotation, and deletion moves
electrostatic interactions. are implemented for the gas molecules in each GCMC cycle [76]. As
suggested by Vasileiadis et al. [39] and Jin et al. [38], the kerogen
2.3. Simulation details structure is kept rigid during our simulations. For each pore pressure, we
first calculate the fugacity and then estimate the number of gas mole­
We explore two different processes of hydrocarbon production cules contained in the nanopore using GCMC. In other words, the pore
pressure is controlled by the different numbers of gas molecules, whose
(pressure drawdown and CO2 injection) at constant temperature (T =
353 K) to assess the efficiency of shale gas recovery and CO2 storage amounts are determined by GCMC. A periodic boundary condition is
imposed in each direction. For each point, our simulation consists of 1 ×
[38]. The initial pore pressure is P0 = 30 MPa, which is given according
to the realistic reservoir condition. First, we reduce the pore pressure 107 steps, in which the first half is performed for equilibrium, while the
other steps are utilized to estimate the gas storage. The absolute
from P0 = 30 MPa to P1 = 15 MPa to mimic the shale gas production
caused by primary pressure drawdown, supposing that the hydrocarbon adsorption amount of hydrocarbons in the nanopore Nab is given by
mixture in pores is in chemical equilibrium with that in an infinite 〈Ni 〉
volume. Then, we inject CO2 into the kerogen nanopore (huff) until the Nab = (1)
V⋅NA
mole fraction of CO2 reaches 25%. During this process, because the pore
Here, <Ni > is the number of gas molecules in our kerogen model;
volume in the external bulk reservoir—which acts like the fractures
NA is the Avogadro constant, i.e., 6.022 × 1023 1/mol; and V denotes the
connected to the nanopores–maintains constant, the density of each
accessible pore volume of the kerogen. We use helium as the probe to
component in the bulk mixture of C1/C2/C3/CO2 at the equilibrium
detect V because its atom is small enough to probe the pores, and it is
pressure P2 is equal to that in C1/C2/C3 mixture at P1. Thus, according to
hardly adsorbed on kerogen [77,78].
the molar composition of the mixture after CO2 injection, we assume
A parameter introduced by Zhou et al. [42], CE, is used to evaluate
different pressures P2 for this system and calculate the alkane densities
the confinement effect on the different types of hydrocarbons,
using the Peng-Robinson equation of state (PR-EOS) until the density of
each component at P2 is equal to that at P1 [72]. Using this method, we yi,pore
CE = (2)
estimate P2 = 19.1 MPa. Then another pressure drawdown process is yi,bulk
conducted from P2 to P3 = 10 MPa while the bulk composition remains

4
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 4. Simulation snapshots showing the recovery of C1/C2/C3 mixture during CO2 huff-n-puff in a 2 nm kerogen nanopore (T = 353 K): (a) initial reservoir (P0 = 30
MPa), (b) after pressure drawdown (P1 = 15 MPa), (c) CO2 huff (P2 = 19.1 MPa), and (d) CO2 puff (P3 = 15 MPa).

where yi pore and yi bulk denote the molar fraction of component i within 3. Model validation
kerogen nanopores and in the bulk fluid, respectively. A value of CE
greater than 1 means a stronger affinity. Prior to the GCMC simulations, we validate our model of kerogen
To quantify the efficiencies of pressure drawdown and CO2 injection, nanopore. The compositions and functional groups of the kerogen unit
we define the hydrocarbon recovery ηi as the ratio of released gas agree well with the experimental data of Kelemen et al [64]. The density
quantity to the initial adsorption amount during a single-stage of our kerogen matrix is 1.24 ± 0.01 g/cm3, which is consistent with the
[41,79,80], experimental results (1.18–1.25 g/cm3) of Type II kerogen reported by
Okiongbo et al. [81]. To estimate the porosity, we detected the free pore
nreleased volume using a helium molecule with a given Connolly radius as the
ηi = i
× 100% (3)
ninitial
i probe. The probe molecules inserted into the kerogen roll on the van der
Waals surface of the skeleton, from which the area wrapped by the
where nreleased
i is the amount of component i recovered from the kerogen, surface is defined as the free pore volume [82]. The estimated porosity
and ninitial
i is the initial adsorption amount of component i before of the kerogen matrix (19.17 ± 0.1%) also lies in the range of mea­
production. surement values (4.45%–22.50%) in the Barnett shale [55].
Moreover, we also computed the excess adsorption capacity of
methane Г ex using GCMC to make a comparison with experimental data,

5
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

adsorbent [92].
After CO2 injection, there are both CH4 and CO2 molecules confined
in the kerogen nanopore. Therefore, the selectivity of CO2 against CH4,
SCO2/CH4, is introduced to quantify the competitive adsorption [35],
xCO2 /xCH4
SCO2 /CH4 = (5)
yCO2 /yCH4

where, xi and yi denote the mole fractions of component i (CO2 or CH4) in


the adsorbed and bulk phases, respectively. They can be readily esti­
mated from GCMC simulations. If the selectivity is greater than 1, CO2 is
preferentially adsorbed [93]. A higher selectivity signifies a stronger
adsorption capacity of CO2 with respect to CH4. In Fig. 6c, we show the
selectivity corresponding to the bulk mole fraction of yCO2 = 0.25. The
selectivity is greater than unity but decreases with pressure, indicating a
more considerable amount of CO2 to be adsorbed on the kerogen and the
adsorption capacity differences between CO2 and CH4 become smaller at
elevated pressure. A similar trend has also been reported for CO2/CH4
mixtures within coal and kerogen matrix [36,44]. Because the energy of
kerogen surface is heterogeneous, the adsorption sites with higher en­
ergy are favorably occupied by CO2 at low pressure; however, with the
Fig. 5. Comparison of the simulated excess adsorption isotherms in a 1.2 nm pressure increasing, once the higher energy positions have been filled,
kerogen-based nanopore (T = 338 K) and the experimental results reported by CO2 and CH4 begins to competitively occupy the lower energy positions,
Gasparik et al. [51] (Barnett shale) and Zhang et al. [84] (Blakely shale). The thus leading to the smaller selectivity at high pressures.
samples from Barnett shale have distinct total organic carbon contents (TOCs) Fig. 6b shows the absolute adsorption isotherms of each component
and vitrinite reflectance values (VRs): VR = 2.2%, TOC = 3.5% (red spheres) in CH4/C2H6/C3H8 mixtures at 353 K. As the pressure increases, the
and VR = 1.0%, TOC = 4.5% (blue triangles). (For interpretation of the ref­ adsorption capacity of CH4 is enhanced, whereas the adsorption amount
erences to color in this figure legend, the reader is referred to the web version of
of C2H6 and C3H8 remains almost unchanged. To better elucidate the
this article.)
adsorption behavior, the parameter, CE (Eq. (2)), is used for the simu­
lations involved with gas mixtures. Fig. 6d reveals that the CE of C3H8
〈Ni 〉 ρbulk V and CO2 are always greater than 1, indicating that the preferential
Γex = − (4)
NA M adsorption of C3H8 and CO2 upon the kerogen surface. Meanwhile, the
In this equation, M represents molar mass (g/mol); and ρbulk (g/cm3) CE value of C3H8 decreases rapidly with pressure, and it is even smaller
is the density of the bulk fluid, which is obtained from the National than that of CO2 when P greater than 20 MPa, because C3H8 confined in
Institute of Standards and Technology (NIST) database [83]. the nanopore saturates at lower pressure. In contrast, the density of the
In Fig. 5, we compare the estimated excess adsorption isotherms of bulk phase continually increases. Similar behavior is also observed for
CH4 within our model and the experimental results. The experimental C2H6. The variation of pore pressure shows a negligible effect on the CE
data are adopted from the Barnett shales measured by Gasparik et al. of CO2, revealing that CO2 always exhibits a superior adsorption ca­
[51] and Blakely shale measured by Zhang et al. [84]. Fig. 5 indicates pacity in the mixture.
that the simulated and measured data exhibit the same trend, which first To disclose the underlying mechanisms of competitive adsorption,
increases quickly and then reaches a plateau. Furthermore, the esti­ we calculate the isosteric heat, Qst, using the formula developed by
mated excess adsorption capacity is also reasonably close to the exper­ Karavias and Myers [94]. The Clausius-Clapeyron equation [95] was not
imental data. The deviations may be attributed to the discrepancy of adopted here because it assumes the adsorbed volume is negligible and
sample properties, such as the pore radius. Bae et al. [85,86] suggested the bulk phase is regarded as ideal gases [96], which contradicts the
that the impenetrable pore necks existing in the realistic shale kerogen high-pressure condition. As shown in Fig. 7a, the isosteric heat of CO2 is
may also contribute the differences. Thus, the validity of our constructed approximately 1.54 times greater than that of CH4, manifesting a
kerogen nanopore has been justified and we will use it to study the re­ stronger affinity between kerogen and CO2. This result explains the
covery of shale gas during pressure drawdown and CO2 injection. preferential adsorption of CO2 over CH4 within the kerogen nanopore
(Fig. 6c). The isosteric heat of C3H8 is much larger than that of CH4 and
4. Results and discussion C2H6, suggesting that longer alkane has much superior interaction with
kerogen (Fig. 7b). This result gets in line with the recent work of Liu
4.1. Adsorption and recovery of CO2-hydrocarbon mixtures within et al. [97], who reported that the heavier component is strongly retained
kerogen nanopore by kerogen matrix. We also note that the isosteric heat of adsorption for
CO2 is smaller than that of C3H8 but greater than CH4 and C2H6. Because
To investigate the effect of CO2 injection on the shale gas recovery, the isosteric heat extrapolated to the limit of zero coverage characterizes
the adsorption behavior of pure and multi-component alkanes in the interactions between adsorbate and adsorbent, we infer that the
kerogen nanopores is analyzed first. We probe the adsorption behavior affinity of these components with kerogen decreases in the order: C3H8,
of pure CH4 within kerogen-based nanopores under the conditions of CO2, C2H6, and CH4. This conclusion agrees well with the resulting CEs
either CO2 presence or absence at 353 K using the GCMC method. As at lower pressures (Fig. 6d). However, as the pressure increases, the
shown in Fig. 6a, the absolute adsorption amount of CH4 increases interactions among adsorbate molecules become stronger and the larger
gradually with the increasing pressures and tends to approach a con­ molecule sizes of C3H8 cause a greater repulsive force, thus leading to
stant. This trend is consistent with the adsorption characteristics of CH4 the restrained adsorption capacity of C3H8 in comparison to CO2
in other nanoporous materials, such as CNTs [87-89] and clays (Fig. 6d).
[35,90,91]. According to the recommended classification of IUPAC (the We further study the recovery process of shale gas using CO2 huff-n-
International Union of Pure and Applied Chemistry), the isotherm cor­ puff. Both the pure CH4 and the CH4/C2H6/C3H8 mixture are taken into
responds to the Langmuir type—a typical behavior for nanoporous account. In Fig. 8, we present the number density distributions of each

6
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 6. (a) Absolute adsorption isotherms, Nab, of pure CH4 in a 2 nm kerogen-based nanopore at T = 353 K; (b) Absolute adsorption isotherms of CH4/C2H6/C3H8
mixture; (c) adsorption selectivity of the CH4/CO2 binary mixture; (d): Confinement effect of each component in the CO2/CH4/C2H6/C3H8/ mixture.

Fig. 7. Variation of isosteric heats (Qst) against the pore pressure: (a) CH4/CO2, (b) CH4/C2H6/C3H8, and (c) CO2/CH4/C2H6/C3H8.

component in 2-nm kerogen nanopores. The green dashed lines repre­ and puff stages, respectively. The reduction of CH4 adsorption peak
sent the interface boundary between the cylinder pore and kerogen during pressure drawdown is greater than that during CO2 injection,
matrix. A pronounced density peak (also termed adsorption layer indicating that decreasing the pore pressure primarily produces adsor­
[98,99]) forms at the interfacial area, and some gas molecules absorb bed CH4 molecules. This phenomenon should be attributed to the
within the kerogen matrix. Dominated by the inherent surface proper­ weaker adsorption capacity at lower pressures (Fig. 6a). Meanwhile, for
ties, the location of adsorption layers remains unchanged. the adsorbed gas within the kerogen matrix, the CH4 quantity produced
Fig. 8a shows the variation of CH4 number density profiles during the during CO2 injection is greater than that exploited during pressure
whole production process. The pressure variations from P0 to P1, P1 to drawdown, demonstrating that the adsorbed gas is favorable to be
P2, and P2 to P3 correspond to the pressure drawdown stage, CO2 huff extracted using CO2 injection. Because the tiny space located within the

7
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 8. Number density profiles of each hydrocarbon and CO2 in a 2-nm kerogen nanopore under different stages (T = 353 K): (a) pure CH4; (b) CH4/C2H6/
C3H8 mixture.

kerogen matrix is much smaller than the outside (cylinder) pore, and outer surfaces of the kerogen matrix, respectively. During the CO2
CO2—which shows a much superior adsorption capacity in smaller puff stage (P2 → P3), there is a decrease in the CO2 adsorption layer, but
pores—preferentially occupies the sorption sites of CH4 and lead to the no significant variation is observed for the CO2 density profile in the
gas release [38]. This conclusion is also supported by Huang et al. [36], kerogen matrix (Fig. 8a, right panel). It reveals that due to the strong
who reported that CO2 and CH4 are favorably adsorbed upon the inner fluid–solid interactions, CO2 stored in the kerogen matrix is difficult to

8
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 9. Absolute adsorption isotherms of pure CH4 in CNT, kerogen slit, and kerogen-based nanopore having the same dimension (2 nm).

molecules tend to move from the kerogen matrix to the surface in the
second pressure drop. As mentioned before, C3H8 shows a much stronger
affinity to kerogen than CH4, C2H6, and CO2. Thus, during this stage,
when shorter alkanes are released from the matrix, some C3H8 molecules
enter the kerogen matrix, which is demonstrated by the higher adsorp­
tion density of C3H8 at P3 than P2 (lower left panel, Fig. 8b). Therefore,
these C3H8 molecules occupy the original adsorption sites of CO2 within
the matrix and drive CO2 molecules into the inner pore.

4.2. Effect of pore structure on shale gas recovery

Existing studies always use CNTs [89], graphene slit [41,100], and
kerogen slit [42] to mimic the shale kerogen to study the adsorption and
transport of hydrocarbons. To examine the validity of these models, in
this section, we study the effect of pore structure on the gas recovery
using CO2 huff-n-puff. Owing to the similarity of chemical composition
between CNT and graphene, we account for three different models:
kerogen slit (Fig. 9a), CNT (Fig. 9b), and kerogen-based nanopore. The
kerogen slit is composed of two kerogen cubic matrices, each of which
consists of 12 kerogen units. The aperture of the kerogen slit is 2 nm,
which is as same as the diameters of CNT and our kerogen nanopore. The
Fig. 10. Molecular structure of (a) kerogen slit and (b) CNT having the same
pore size. estimated pore volumes for kerogen slit, carbon nanotube, and circular
kerogen nanopore are 20.413 nm3, 19.195 nm3, and 28.067 nm3,
respectively. Because our kerogen-based nanopore consists of a 2-nm
exploit; organic-rich shale shows good potential for CO2 sequestration.
cylindrical pore and a kerogen matrix, the available pore volume is
Meanwhile, adsorbed CO2 molecules are released from the surface
greater than that of CNT even they have the same inner diameter.
because of the relatively smaller interactions in the inner pore.
Firstly, we compute the adsorption isotherms of pure CH4 in the
In Fig. 8b, we show the density profile of each component when CO2
three models. As shown in Fig. 10, all the absolute adsorption isotherms
is injected into the CH4/C2H6/C3H8 mixture in kerogen. As the pressure
increase monotonically with growing pressure and approach a plateau at
drops (P0 → P1), the density of CH4 in the adsorption layer decreases, but
higher pressure. However, when the pressure is lower, the adsorption
C2H6 doesn’t show significant variations in the density profiles. How­
isotherms within different geometries show distinct increasing rates.
ever, we observe the enrichment of C3H8 near the interfacial region at
The adsorption capacities in kerogen slit and cylinder pore are almost
lower pressures, which is in contrast to CH4 (Fig. 8b, lower-left panel).
identical when P < 8 MPa but much smaller than that in a carbon
We attribute that the adsorption sites on the kerogen surface are not
nanotube. Besides, the maximum adsorption amount in CNT also ex­
fully occupied at low pressures; therefore, the stronger interactions be­
ceeds that in kerogen nanopores. For example, the adsorption capacity
tween kerogen and C3H8 cause the preferential absorption of more C3H8
of CH4, Nab, in CNT is 0.00842 mmol/m2 at P = 40 MPa, which is ~ 1.35
molecules upon the interface [33], leading to an increasing density.
times as larger as that in kerogen nanopores (0.00623 mmol/m2). Both
However, at elevated pressures, the adsorption sites are almost filled,
the surface roughness and chemical composition contribute to this
and the entropic effect becomes more pronounced, causing a decreased
phenomenon [32]. On the one hand, the CNT surface is ultrasmooth,
density of C3H8 [42]. After CO2 injection, the amount of the adsorbed
whereas the surface of kerogen nanopores is rough. Tesson et al. [32]
hydrocarbons decreases, and more carbon dioxide molecules are
suggested that a smooth slit adsorbs 42% gas molecules more than a
adsorbed in the kerogen because of the greater attractions. Fig. 8b also
rough slit. On the other hand, the carbon nanotube is totally composed
suggests that the CH4 recovery during CO2 injection (P1 → P2) is
of carbon atoms, CH4 can be readily adsorbed anywhere on the surfaces;
significantly lower than that obtained during pressure drawdown (P0 →
however, the composition of kerogen matrix is intricate, including a
P1); however, the variation of C2H6 shows the opposite trend. Further­
large number of hydrogen and carbon atoms and a few nitrogen, sulfur,
more, the number of C3H8 molecules in the adsorption layer and kerogen
and oxygen atoms. These heteroatoms exert a distinct interaction on the
matrix decreases remarkably (P1 → P2), which is contrary to the stages of
gas molecules compared to the pure carbon skeleton [47,100]. Fig. 10
pressure drawdown with CO2 presence (P2 → P3) or absence (P0 → P1).
also indicates that if the pressure is greater than 8 MPa, more CH4
Meanwhile, the density peak of CO2 at P3 (after 2nd pressure down) is
molecules are adsorbed within the kerogen slit than in the kerogen
slightly higher than that at P2 (after CO2 injection), revealing that CO2
nanopore with the same pore size. Although the chemical composition of

9
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 11. Absolute adsorption isotherms (left panel) and CE values (right panel) of (a, d) CH4, (b, e) C2H6, and (c, f) C3H8 in their mixtures confined in different pores.

these two models is similar, they show different pore curvatures. The chemical composition, the specific adsorption amounts are different,
curvature of a kerogen slit is eliminated, whereas the kerogen-based especially for methane. Caused by the pore geometry effect, the
nanopore having a greater curvature leads to the decrease of potential adsorption capacity rises faster in the slit-shaped pores (Fig. 11a). We
well depth, causing less adsorption [59]. Similar conclusions have also also note that the adsorption behavior of each component in carbon
been reported in carbon nanotube and graphene slit [101]. nanotubes are tremendously different from those in kerogen cylindrical
We also analyze the effect of pore geometry on the adsorption pores. Therefore, in order to elucidate the adsorption and transport
behavior of CH4, C2H6, and C3H8 mixtures in the three different models. behavior of gas confined in shale kerogen, cylindrical kerogen por­
In general, the adsorption isotherms of CH4 increase with pressure, es—instead of the carbon nanotube and kerogen slit—should be utilized,
whereas C2H6 adsorption capacity remains almost unchanged when P because the other models cannot account for the influence of hetero­
greater than 10 MPa, and the adsorption of C3H8 tends to be suppressed atoms, amorphons structure, and circular nanopores (diameter: 2–100
(Fig. 11). Because the molar ratio of CH4, C2H6, and C3H8 is 80:15:5, the nm) of realistic kerogen. This conclusion sheds light on the constructions
amount of adsorbed CH4 exhibits the largest value in each pore followed of molecular models involved with shale resources and also suggests that
by C2H6 and C3H8. For example, the adsorption amounts of CH4, C2H6, some results of the existing researches may be questionable. A detailed
and C3H8 in the kerogen-based nanopore are 0.0044, 0.00082, and comparison between our work and previous studies is summarized in
0.00036 mmol/m2 at 40 MPa, respectively. Moreover, the adsorption Table 1.
amounts of CH4 in kerogen slit are larger than others, whereas the ca­ To further assess the selectivity of hydrocarbons in different pore
pacities of C2H6 and C3H8 in CNT are significantly greater than the other structures, we show CE values in Fig. 11b. They exhibit the same trend
two models. Although the adsorption isotherms in the cylindrical and that the CE values of C2H6 and C3H8 decrease with pressures but that of
slit-shaped kerogen pore show similar trends because of the same CH4 grow slightly, indicating that the competitive adsorption capacity of

10
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

CH4 enhances whereas those of C2H6 and C3H8 become weaker. The
reason is that longer alkanes reach adsorption saturation quickly at
lower pressures; however, at higher pressures, the confined fluids
become denser, and the differences in molecular chains tend to be less
significant [102]. For most of the simulation conditions, the CE values of
both C2H6 and C3H8 are greater than unity, suggesting that heavier
components are more favorable for adsorption.
Moreover, the CE value of each component in kerogen slit and cir­
cular nanopore gets close to unity at P = 40 MPa. As reported by Denayer
et al. [102], the adsorption behavior of hydrocarbons in the nanopore is
controlled by the combination of entropic and enthalpic effects, and the
less significant differences among molecule types can be found in the
densely confined space under higher pressures. Owing to the influences
of pore geometry and chemical composition, the CE value of each
component in CNT is tremendously distinct from that in kerogen slit and
nanopore: for CH4, the CE value is lower in CNT, whereas for C2H6 and
C3H8, the values are greater.
Fig. 12 shows the estimated recovery efficiency of each component in
different pores during CO2 huff and puff. For pure CH4 confined in the
kerogen slit, the highest recovery appears during the first pressure
drawdown process because of the steepest adsorption isotherm (Fig. 10).
However, a comparison with the results in circular pores suggests that
the gas recovery behavior is also influenced by pore geometry: CO2 in­
jection stage produced the greatest recovery ratio in CNT and kerogen
nanopores (Fig. 12a). Owing to the stronger interactions between CO2
and kerogen, adsorbed CO2 molecules primarily occupy the inner sur­
faces of the kerogen matrix, while CH4 is mainly located upon the outer
surface and larger pores [37]. Thus, when CO2 is injected, CH4 mole­
cules tend to be expelled from the matrix surface. Because CNT is
exclusively composed of carbon atoms and the surface is ultra-smooth,
the gas releasement is favorable after CO2 injection, causing a higher
recovery ratio.
Moreover, owing to the decreasing adsorption capacity with
reducing pressure (Fig. 11a), CH4 recovery in the mixtures during each
stage reduces gradually with the production process (P0 → P1 → P2 →
P3), indicating that pressure drawdown is a better way to retrieve CH4 in
the mixtures (Fig. 12b). However, CO2 injection is more effective for
exploiting C2H6 and C3H8 (Fig. 12b). This observation can be explained
using the flat isotherms of C2H6 and steeply decreased isotherms of C3H8
(Fig. 11c). In CNT and kerogen-based nanopore, the recovery ratio of
hydrocarbons during CO2 injection increase with the number of carbon
atoms, whereas the pressure drop stages show the opposite trend. The
recoveries of C2H6 and C3H8 during the pressure drawdown stage are
even negative because the adsorption amount of C2H6 remains almost
unchanged with pressure, but that of C3H8 reduces. The “negative re­
covery ratio” means that no C3H8 modules are released during this stage
because the adsorption capacity of C3H8 decreases with pressure when P
greater than 8 MPa (Fig. 11). Zhou et al. [42] reported a similar behavior
in kerogen slit. Owing to the stronger affinity with kerogen (Fig. 7), C3H8
achieves saturation at lower pressure, whereas an elevated pressure
facilitates the adsorption of lighter alkanes. Therefore, shorter compo­
nents are favorably produced during pressure drop, but C3H8 is more
likely to be left in the pore.

4.3. Effect of pore size on shale gas recovery

In Fig. 13, we present the variation of CH4 absolute adsorption iso­


therms against the sizes of different kerogen cylindrical pores. As ex­
pected, the overall trends of these isotherms are the same, i.e., the
adsorption capacity increases with elevated pressure. However, the
Fig. 12. Shale gas recovery in different pore models during CO2 huff-n-puff: (a)
different units (such as mmol/m2 and mmol/cm3) of adsorption iso­
pure CH4; (b) CH4/C2H6/C3H8 mixture.
therms lead to distinct conclusions on the effect of pore size. If Nab is
characterized using per unit surface area (mmol/m2), the adsorption is
higher in larger pores because the greater size facilitates the formation
of layered structures [80]. Therefore, Nab expressed in mmol/m2 is
positively correlated with the pore radius [73,103,104]. However, the

11
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 13. Absolute adsorption isotherms of pure CH4 within kerogen-based nanopores having different radii. The adsorption capacities are shown in different units:
(a) mmol/m2; (b) mmol/cm3.

fluid–solid adhesion energy becomes greater (deeper) with pore size preferential adsorption capacity of CO2. Because the pressure drop from
decreasing [60,84,97.104]; thus, the adsorption amount of CH4 in per P0 to P1, 15 MPa, is larger than that from P2 to P3, 4.9 MPa, the recovery
volume of a narrower pore is larger, leading to the opposite variation ratio in the second pressure drop is much smaller than that in the first
tendencies of the adsorption capacities with respect to radius. pressure drop stage. As shown in Fig. 15b, for each component in the
Fig. 14 shows the absolute adsorption isotherms of each component hydrocarbon mixture, the recovery ratio during CO2 injection in the 1.2-
in the shale gas mixtures confined in kerogen-based nanopores having nm kerogen pore is higher than the other stages, which means injecting
different radii (T = 353 K). We represent the adsorption amount using CO2 can efficiently recover the hydrocarbons confined in much smaller
the per accessible volume. Because of the weaker fluid–solid in­ pores. During pressure drawdown, the recovery ratio of C3H8 presents
teractions, their adsorption capacities in a 5.4-nm kerogen pore are negative values, indicating that no C3H8 modules are released out during
much lower than 2-nm and 1.2-nm nanopores. CH4 in the hydrocarbons this stage because of the decreasing adsorption capacity with pressure
mixtures exhibit a similar trend of adsorption isotherms with pure CH4; (Fig. 14c). In a 2-nm nanopore, the recovery ratio of CH4 in the first
however, the exact values are much smaller because C2H6 and C3H8 pressure drop (25.51%) is slightly higher than that of the CO2 huff stage
would occupy some adsorption sites on the kerogen surface. Fig. 14a (18.21%). In contrast, for C2H6 and C3H8, the recovery during CO2 in­
also indicates that the adsorption isotherms of CH4 increase with pres­ jection is much higher, i.e., 29.08% and 36.70%, respectively. Similar
sure in each pore. However, the adsorption capacities of C2H6 and C3H8 conclusions can also be reached in the 5.4 nm kerogen nanopore,
increase at lower pressures and then decrease in a 1.2-nm kerogen-based revealing that CO2 huff-n-puff serves as a promising technology for
nanopore but remain almost constant in a 2-nm pore and exhibit a slight exploiting heavier hydrocarbons.
increase in a 5.4-nm pore. This observation suggests that the adsorption
behavior of longer alkanes in the mixture is pore size dependent;
enlarging the pore causes the shifts of the saturation pressure to elevated 4.4. Implications for shale gas recovery and CO2 sequestration
values.
The estimated CE values are imposed in Fig. 14 (right panel) to study In the above sections, we have analyzed the competitive adsorption
the selectivity adsorption of hydrocarbons mixture in kerogen nano­ behaviors of hydrocarbon within shale kerogen nanopores and explored
pores with different radii. It is evident that the CE value for each the gas recovery mechanisms during each stage influenced by various
component decreases in the order of C3H8, C2H6, and CH4. More pre­ factors, e.g., pressure, pore structure, and pore size. To gain a better
cisely, the CE values of CH4 are typically smaller than unity, while those insight into the field applications, we show the total gas recovery and
of C2H6 and C3H8 are greater than unity. This phenomenon further CO2 sequestration ratio corresponding to distinct hydrocarbon compo­
demonstrates the preferential adsorption of the heavier components in sitions (i.e., pure CH4 and hydrocarbon mixture) in Fig. 16.
kerogen-based nanopores, which should be attributed to the stronger The total gas recovery is defined as the ratio between the total
fluid–solid interactions. For the gas mixture, the confinement effect of number of gas molecules released from P0 (30 MPa) to P3 (after CO2 huff-
C2H6 and C3H8 is higher in the vicinity of the solid surface, and more n-puff) and the initial loadings. Fig. 16 indicates that the recovery ratio
CH4 molecules are forced to stay in the central pore where the affinities of CH4 in pure gas is almost identical to that in the mixture, revealing
from the kerogen are weaker. Fig. 14 also suggests that owing to the that the existence of other components, such as C2H6 and C3H8, has no
weakened confinement effect, the deviations of the CE values from the significant influences on the CH4 recovery. A comparison among the
unity decrease when enlarging the pore size. As the pressure increased to total recoveries within distinct pores suggests that, in the carbon
40 MPa, the difference among the CE values of these three components nanotube, the recovery of CH4 is highest (56%), whereas that of C2H6
tend to be negligible because the confined space gets more densely (27%) and C3H8 (-21%) are lowest. This result agrees well with the
packed at higher pressures, thus leading to the less remarkable differ­ aforementioned adsorption behavior. As shown in Fig. 11, caused by the
ences among the molecular types [102]. differences in surface roughness and thermophysical properties, we
With the increasing pore size, the recovery ratio of pure CH4 during observe much greater adsorption amounts of C2H6 and C3H8 confined in
the pressure drop process increases while that corresponding to CO2 CNT, implying the stronger adsorption capacity as compared to kerogen
injection decreases (Fig. 15a). We attribute the reason to the reducing slit and nanopore. Therefore, during CO2 huff and puff, fewer C2H6 and
affinities from the kerogen surface, which leads to the diminishing C3H8 molecules will be released in CNT.
Given that the chemical composition of CNT is distinct from the real

12
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 14. Absolute adsorption isotherms (left panel) and CE values (right panel) of (a, d) CH4, (b, e) C2H6, and (c, f) C3H8 in the mixtures confined in kerogen-based
nanopores having different radii.

shale kerogen and its ultrasmooth surface may lead to an erroneous observe a greater CO2 sequestration ratio in the hydrocarbon mixture,
understanding of the fluid flow behavior, we suggest that the carbon revealing that more CO2 can be stored in shale reservoirs with the in­
nanotube model should not be utilized to mimic shale kerogen. As ex­ crease of other components. For pure CH4, the sequestration efficiency
pected, increasing the pore size will diminish the interactions from the in CNT is much larger than the other two models, which accounts for the
kerogen surface and further facilitate shale gas recovery (Fig. 16b). highest recovery ratio of CH4 (39.19%), as shown in Fig. 12a. However,
Among these components, the highest recovery ratio of CH4 (52%) for the hydrocarbon mixture, the maximum sequestration appears in the
manifests that the lighter components are preferentially recovered kerogen slit (7.05%). When we inject the same amount of CO2 into a
because of the weakest adsorption capacity (Fig. 7). shale reservoir, a greater hydrocarbon recovery also means a higher
The effect of pore structures and sizes on CO2 sequestration ratio, sequestration efficiency of CO2 because the total pore volume of the
ξCO2, for the fluids having different compositions are presented in reservoir remains almost constant. Fig. 16c also suggests that the
Fig. 16b. We define this parameter as, sequestration ratio gets a remarkable increase with pore radius because
of the larger storage space and weaker affinity from the solid surface.
nsequestrated These results indicate that the shale reservoirs having larger pores can
ξCO2 = CO2
× 100% (6)
ninitial
CO2 be a good location for the geological sequestration of CO2, which not
only facilitates the adequate replacement of shale gas, but also ensure a
where nsequestrated
CO2 is the number of CO2 molecules sequestrated in the greater sequestration ratio.
pore, and ninitial
CO2 is the total number of CO2 molecules contained in the
model at the initial stage of the second pressure drawdown process. We

13
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 15. Recovery of shale gas confined in shale nanopores having different radii during CO2 huff-n-puff: (a) pure CH4; (b) recovery of each component in the
hydrocarbon mixture.

14
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

Fig. 16. (a) Effect of pore structure on the total recovery of each component at 353 K (P0 → P3); (b) Effect of pore size on the total recovery of each component within
the kerogen-based nanopore at 353 K (P0 → P3); (c) CO2 sequestration ratio in kerogen pores having different geometries and sizes. Both the results for pure methane
and CH4/C2H6/C3H8 mixtures (80:15:5) are included..

5. Conclusions diminishing the CO2 injection effect; however, both the total gas re­
covery and CO2 sequestration ratio will increase. Our study sheds new
Using GCMC simulations, we studied the hydrocarbon recovery light on understanding the fundamental mechanisms of enhanced shale
mechanisms and CO2 sequestration in realistic shale kerogen during CO2 gas recovery and CO2 sequestration through CO2 huff-n-puff, which
huff-n-puff. We explored the effects of pressure, pore structure, and pore would benefit the efficient design of shale gas development strategies.
size on gas recovery and sequestration efficiency, accounting for pure
CH4 and hydrocarbon mixture. A comparative study among carbon
nanotube, kerogen slit, and kerogen cylindrical nanopore suggests that Declaration of Competing Interest
the different chemical compositions and surface roughness induce the
tremendously different hydrocarbon adsorption and CO2 sequestration The authors declare that they have no known competing financial
behavior in these pores. CO2 injection may serve as an efficient method interests or personal relationships that could have appeared to influence
for circular pores, whereas the pressure drawdown technology is more the work reported in this paper.
efficient for slit-shaped kerogen nanopore. We recommend that when
one wants to study the adsorption and transport of fluids through shale Acknowledgements
kerogen using molecular modeling, it is better to use the cylindrical
kerogen-based nanopore instead of carbonaceous models (such as car­ This work is supported partly by the National Natural Science
bon nanotube and graphene) as well as kerogen slit. Foundation of China [grant numbers 51704312, U1762213, and
Owing to the stronger interactions with the kerogen matrix, the 51974340]; the Applied Fundamental Research Project of Qingdao
adsorption capacity of CO2 surpasses other components (CH4/C2H6/ [grant number No. 19-6-2-21-cg]; and the NanoGeosciences Lab and
C3H8). Thus, the injected CO2 drives the adsorbed hydrocarbon mole­ Mudrock Systems Research Laboratory at the Bureau of Economic Ge­
cules from the surface and leads to an enhanced gas recovery. The ology, Jackson School of Geosciences, The University of Texas at Austin.
density distributions of pure CH4 in the cylindrical kerogen nanopore We would like to thank Dr. Shansi Tian for providing the SEM image.
suggest that decreasing pressure primarily exploits gas in the adsorption
layer, while CO2 injection releases CH4 absorbed within the kerogen References
matrix. For the mixture, CO2 huff-n-puff is more promising for the re­
covery of heavier hydrocarbons, while pressure drawdown favors the [1] R.D. Vidic, S.L. Brantley, J.M. Vandenbossche, D. Yoxtheimer, J.D. Abad, Impact
production of lighter hydrocarbons. Therefore, the oil company may of shale gas development on regional water quality, Science 340 (2013) 1235009-
1235009. https://doi.org/10.1126/science.1235009.
develop a shale gas reservoir through low-cost pressure drop technology [2] J.D. Hughes, Energy: A reality check on the shale revolution, Nature 494 (7437)
first, which could quickly recover the investment and then change it to (2013) 307–308, https://doi.org/10.1038/494307a.
CO2 huff-n-puff to improve the benefits. Moreover, enlarging the pore [3] A. Paylor, The social–economic impact of shale gas extraction: a global
perspective, Third World Q. 38 (2) (2017) 340–355, https://doi.org/10.1080/
size will improve the performance of pressure drawdown while 01436597.2016.1153420.

15
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

[4] J. Li, Z. Chen, K. Wu, K. Wang, J. Luo, D. Feng, S. Qu, X. Li, A multi-site model to [31] D.J.K. Ross, R. Marc Bustin, Impact of mass balance calculations on adsorption
determine supercritical methane adsorption in energetically heterogeneous capacities in microporous shale gas reservoirs, Fuel 86 (17-18) (2007)
shales, Chem. Eng. J. 349 (2018) 438–455, https://doi.org/10.1016/j. 2696–2706, https://doi.org/10.1016/j.fuel.2007.02.036.
cej.2018.05.105. [32] S. Tesson, A. Firoozabadi, Methane adsorption and self-diffusion in shale kerogen
[5] Energy Information Administration, Shale gas production drives world natural and slit nanopores by molecular simulations, J. Phys. Chem. C. 122 (41) (2018)
gas production growth. https://www.eia.gov/todayinenergy/detail.php? 23528–23542, https://doi.org/10.1021/acs.jpcc.8b07123.
id=27512, 2016 (accessed January 15, 2021). [33] K. Falk, R. Pellenq, F.J. Ulm, B. Coasne, Effect of chain length and pore
[6] S.R. Etminan, F. Javadpour, B.B. Maini, Z. Chen, Measurement of gas storage accessibility on alkane adsorption in kerogen, Energy Fuels 29 (12) (2015)
processes in shale and of the molecular diffusion coefficient in kerogen, Int. J. 7889–7896, https://doi.org/10.1021/acs.energyfuels.5b02015.
Coal Geol. 123 (2014) 10–19, https://doi.org/10.1016/j.coal.2013.10.007. [34] H. Wu, Y. He, R. Qiao, Recovery of multicomponent shale gas from single
[7] J.G. Speight, Chapter 2 - Shale Gas Resources, in: J.G. Speight (Ed.), Shale Gas nanopores, Energy Fuels 31 (8) (2017) 7932–7940, https://doi.org/10.1021/acs.
Production Processes, Gulf Professional Publishing, Boston, 2013, pp. 25–68, energyfuels.7b01013.
https://doi.org/10.1016/C2012-0-00596-0. [35] S. Wang, Q. Feng, F. Javadpour, Q. Hu, K. Wu, Competitive adsorption of
[8] Y. He, S. Cheng, Z. Sun, Z. Chai, Z. Rui, Improving Oil Recovery Through Fracture methane and ethane in montmorillonite nanopores of shale at supercritical
Injection and Production of Multiple Fractured Horizontal Wells, J. Energ. conditions: A grand canonical Monte Carlo simulation study, Chem. Eng. J. 355
Resour. Technol. 142 (2020) 1–19, https://doi.org/10.1115/1.4045957. (2019) 76–90, https://doi.org/10.1016/j.cej.2018.08.067.
[9] R.M. Slatt, Important geological properties of unconventional resource shales, [36] L. Huang, Z. Ning, Q. Wang, R. Qi, Y. Zeng, H. Qin, H. Ye, W. Zhang, Molecular
Cent. Eur. J. Geosci. 3 (2011) 435–448, https://doi.org/10.2478/s13533-011- simulation of adsorption behaviors of methane, carbon dioxide and their mixtures
0042-2. on kerogen: Effect of kerogen maturity and moisture content, Fuel 211 (2018)
[10] M. Rezaveisi, F. Javadpour, K. Sepehrnoori, Modeling chromatographic 159–172, https://doi.org/10.1016/j.fuel.2017.09.060.
separation of produced gas in shale wells, Int. J. Coal Geol. 121 (2014) 110–122, [37] L. Huang, Z. Ning, Q. Wang, W. Zhang, Z. Cheng, X. Wu, H. Qin, Effect of organic
https://doi.org/10.1016/j.coal.2013.11.005. type and moisture on CO2 /CH4 competitive adsorption in kerogen with
[11] J. Lee, K. Perry, Unconventional Gas Reservoirs—Tight Gas, Coal Seams, and implications for CO2 sequestration and enhanced CH4 recovery, Appl. Energy 210
Shales, Unconv. Gas. 29 (2007). (2018) 28–43, https://doi.org/10.1016/j.apenergy.2017.10.122.
[12] Y. Sun, G. Zhang, S. Li, S. Jiang, CO2/N2 injection into CH4+C3H8 hydrates for gas [38] J. Zhou, Z. Jin, K.H. Luo, Effects of Moisture Contents on Shale Gas Recovery and
recovery and CO2 sequestration, Chem. Eng. J. (2019), 121973, https://doi.org/ CO2 Sequestration, Langmuir 35 (26) (2019) 8716–8725, https://doi.org/
10.1016/j.cej.2019.121973. 10.1021/acs.langmuir.9b00862.
[13] R. Ershadnia, C.D. Wallace, M.R. Soltanian, CO2 geological sequestration in [39] M. Vasileiadis, L.D. Peristeras, K.D. Papavasileiou, I.G. Economou, Transport
heterogeneous binary media: Effects of geological and operational conditions, properties of shale gas in relation to kerogen porosity, J. Phys. Chem. C. 122 (11)
Adv. Geo-energ. Res. 4 (4) (2020) 392–405. (2018) 6166–6177, https://doi.org/10.1021/acs.jpcc.8b00162.
[14] J. Jiang, Z. Rui, R. Hazlett, J. Lu, An integrated technical-economic model for [40] T.A. Ho, Y. Wang, Y. Xiong, L.J. Criscenti, Differential retention and release of
evaluating CO2 enhanced oil recovery development, Appl. Energy, 247, 190-211. CO2 and CH4 in kerogen nanopores: Implications for gas extraction and carbon
https://doi.org/10.1016/j.apenergy.2019.04.025. sequestration, Fuel 220 (2018) 1–7, https://doi.org/10.1016/j.fuel.2018.01.106.
[15] T. Ahmed, H. Nasrabadi, A. Firoozabadi, Complex flow and composition path in [41] M. Zhang, S. Zhan, Z. Jin, Recovery mechanisms of hydrocarbon mixtures in
CO2 injection schemes from density effects, Energy Fuels 26 (7) (2012) organic and inorganic nanopores during pressure drawdown and CO2 injection
4590–4598, https://doi.org/10.1021/ef300502f. from molecular perspectives, Chem. Eng. J. 382 (2020) 122808, https://doi.org/
[16] J. Huang, T. Jin, M. Barrufet, J. Killough, Evaluation of CO2 injection into shale 10.1016/j.cej.2019.122808.
gas reservoirs considering dispersed distribution of kerogen, Appl. Energy 260 [42] J. Zhou, Z. Jin, K.H. Luo, Insights into recovery of multi-component shale gas by
(2020) 114285, https://doi.org/10.1016/j.apenergy.2019.114285. CO2 injection: A molecular perspective, Fuel 267 (2020) 117247, https://doi.org/
[17] P. Zuloaga, W. Yu, J. Miao, K. Sepehrnoori, Performance evaluation of CO2 Huff- 10.1016/j.fuel.2020.117247.
n-Puff and continuous CO2 injection in tight oil reservoirs, Energy 134 (2017) [43] M. Firouzi, J. Wilcox, Molecular modeling of carbon dioxide transport and storage
181–192, https://doi.org/10.1016/j.energy.2017.06.028. in porous carbon-based materials, Microporous Mesoporous Mater. 158 (2012)
[18] C. Song, D. Yang, Experimental and numerical evaluation of CO2 huff-n-puff 195–203, https://doi.org/10.1016/j.micromeso.2012.02.045.
processes in Bakken formation, Fuel 190 (2017) 145–162, https://doi.org/ [44] J. Zhang, K. Liu, M.B. Clennell, D.N. Dewhurst, M. Pervukhina, Molecular
10.1016/j.fuel.2016.11.041. simulation of CO2–CH4 competitive adsorption and induced coal swelling, Fuel
[19] A. Abedini, F. Torabi, Oil recovery performance of immiscible and miscible CO2 160 (2015) 309–317, https://doi.org/10.1016/j.fuel.2015.07.092.
huff-and-puff processes, Energy Fuels 28 (2) (2014) 774–784, https://doi.org/ [45] J. Zhang, N. Burke, S. Zhang, K. Liu, M. Pervukhina, Thermodynamic analysis of
10.1021/ef401363b. molecular simulations of CO2 and CH4 adsorption in FAU zeolites, Chem. Eng. Sci.
[20] X. Meng, Z. Meng, J. Ma, T. Wang, Performance evaluation of CO2 huff-n-puff gas 113 (2014) 54–61, https://doi.org/10.1016/j.ces.2014.04.001.
injection in shale gas condensate reservoirs, Energies 12 (2019) 42, https://doi. [46] J. Collell, G. Galliero, R. Vermorel, P. Ungerer, M. Yiannourakou, F. Montel,
org/10.3390/en12010042. M. Pujol, Transport of multicomponent hydrocarbon mixtures in shale organic
[21] F. Torabi, A.Q. Firouz, A. Kavousi, K. Asghari, Comparative evaluation of matter by molecular simulations, J. Phys. Chem. C. 119 (39) (2015)
immiscible, near miscible and miscible CO2 huff-n-puff to enhance oil recovery 22587–22595, https://doi.org/10.1021/acs.jpcc.5b07242.
from a single matrix–fracture system (experimental and simulation studies), Fuel [47] T. Zhao, X. Li, H. Zhao, M. Li, Molecular simulation of adsorption and
93 (2012) 443–453, https://doi.org/10.1016/j.fuel.2011.08.037. thermodynamic properties on type II kerogen: Influence of maturity and moisture
[22] P. Nguyen, J.W. Carey, H.S. Viswanathan, M. Porter, Effectiveness of supercritical content, Fuel 190 (2017) 198–207, https://doi.org/10.1016/j.fuel.2016.11.027.
CO2 and N2 huff-and-puff methods of enhanced oil recovery in shale fracture [48] K. Li, S. Kong, P. Xia, X. Wang, Microstructural characterisation of organic matter
networks using microfluidic experiments, Appl. Energy 230 (2018) 160–174, pores in coal-measure shale, Adv. Geo-energ. Res. 4 (4) (2020) 372–391, https://
https://doi.org/10.1016/j.apenergy.2018.08.098. doi.org/10.46690/ager.2020.04.04.
[23] R. Khosrokhavar, K. Wolf, H. Bruining, Sorption of CH4 and CO2 on a [49] Y. Tang, C. Hou, Y. He, Y. Wang, Y. Chen, Z. Rui, Review on pore structure
carboniferous shale from Belgium using a manometric setup, Int. J. Coal Geol. characterization and microscopic flow mechanism of CO2 flooding in porous
128 (2014) 153–161, https://doi.org/10.1016/j.coal.2014.04.014. media, Energy Technol. 9 (1) (2021) 2000787, https://doi.org/10.1002/
[24] P. Charoensuppanimit, S.A. Mohammad, K.A.M. Gasem, Measurements and ente.202000787.
modeling of gas adsorption on shales, Energy Fuels 30 (3) (2016) 2309–2319, [50] T.A. Ho, L.J. Criscenti, Y. Wang, Nanostructural control of methane release in
https://doi.org/10.1021/acs.energyfuels.5b02751. kerogen and its implications to wellbore production decline, Sci. Rep. 6 (2016)
[25] X. Tang, J. Zhang, X. Wang, B. Yu, W. Ding, J. Xiong, Y. Yang, L. Wang, C. Yang, 28053, https://doi.org/10.1038/srep28053.
Shale characteristics in the southeastern Ordos Basin, China: Implications for [51] M. Gasparik, P. Bertier, Y. Gensterblum, A. Ghanizadeh, B.M. Krooss, R. Littke,
hydrocarbon accumulation conditions and the potential of continental shales, Int. Geological controls on the methane storage capacity in organic-rich shales, Int. J.
J. Coal Geol. 128 (2014) 32–46, https://doi.org/10.1016/j.coal.2014.03.005. Coal Geol. 123 (2014) 34–51, https://doi.org/10.1016/j.coal.2013.06.010.
[26] A.-L. Cheng, W.-L. Huang, Selective adsorption of hydrocarbon gases on clays and [52] A. Afsharpoor, F. Javadpour, Liquid slip flow in a network of shale noncircular
organic matter, Org. Geochem. 35 (4) (2004) 413–423, https://doi.org/10.1016/ nanopores, Fuel 180 (2016) 580–590, https://doi.org/10.1016/j.
j.orggeochem.2004.01.007. fuel.2016.04.078.
[27] S. Kong, X. Huang, K. Li, X. Song, Adsorption/desorption isotherms of CH4 and [53] F.P. Wang, R.M. Reed, A. John, G. Katherine, Pore networks and fluid flow in gas
C2H6 on typical shale samples, Fuel 255 (2019) 115632, https://doi.org/ shales, in: Proc. - SPE Annu. Tech. Conf. Exhib., 2009. https://doi.org/10.2118/
10.1016/j.fuel.2019.115632. 124253-ms.
[28] Y.u. Wang, T.T. Tsotsis, K. Jessen, Competitive sorption of methane/ethane [54] R.J. Ambrose, R.C. Hartman, M. Diaz-Campos, I.Y. Akkutlu, C.H. Sondergeld,
mixtures on shale: Measurements and modeling, Ind. Eng. Chem. Res. 54 (48) Shale gas-in-place calculations part I: New pore-scale considerations, Spe J. 17
(2015) 12187–12195, https://doi.org/10.1021/acs.iecr.5b02850. (2012) 219–229, https://doi.org/10.2118/131772-PA.
[29] H. Zhao, T. Wu, A. Firoozabadi, High pressure sorption of various hydrocarbons [55] R.G. Loucks, R.M. Reed, S.C. Ruppel, D.M. Jarvie, Morphology, genesis, and
and carbon dioxide in Kimmeridge Blackstone and isolated kerogen, Fuel 224 distribution of nanometer-scale pores in siliceous mudstones of the mississippian
(2018) 412–423, https://doi.org/10.1016/j.fuel.2018.02.186. barnett shale, J. Sediment. Res. 79 (12) (2009) 848–861, https://doi.org/
[30] T. Wu, H. Zhao, S. Tesson, A. Firoozabadi, Absolute adsorption of light 10.2110/jsr.2009.092.
hydrocarbons and carbon dioxide in shale rock and isolated kerogen, Fuel 235 [56] T. Wang, S. Tian, G. Li, M. Sheng, W. Ren, Q. Liu, S. Zhang, Molecular simulation
(2019) 855–867, https://doi.org/10.1016/j.cej.2018.08.067. of CO2/CH4 competitive adsorption on shale kerogen for CO2 sequestration and
enhanced gas recovery, J. Phys. Chem. C. 122 (30) (2018) 17009–17018, https://
doi.org/10.1021/acs.jpcc.8b02061.

16
S. Wang et al. Chemical Engineering Journal 425 (2021) 130292

[57] A. Qajar, H. Daigle, M. Prodanovic, The effects of pore geometry on adsorption [81] K.S. Okiongbo, A.C. Aplin, S.R. Larter, Changes in type II kerogen density as a
equilibrium in shale formations and coal-beds: Lattice density functional theory function of maturity: Evidence from the kimmeridge clay formation, Energy Fuels
study, Fuel 163 (2016) 205–213, https://doi.org/10.1016/j.fuel.2015.09.061. 19 (2005) 2495–2499, https://doi.org/10.1021/ef050194+.
[58] Y.u. Pang, X. Hu, S. Wang, S. Chen, M.Y. Soliman, H. Deng, Characterization of [82] M.L. Connolly, Solvent-accessible surfaces of proteins and nucleic acids, Science
adsorption isotherm and density profile in cylindrical nanopores: Modeling and (80-.). 221 (1983) 709-713. https://doi.org/10.1126/science.6879170.
measurement, Chem. Eng. J. 396 (2020) 125212, https://doi.org/10.1016/j. [83] G. Chen, J. Zhang, S. Lu, M. Pervukhina, K. Liu, Q. Xue, H. Tian, S. Tian, J. Li, M.
cej.2020.125212. B. Clennell, D.N. Dewhurst, Adsorption behavior of hydrocarbon on illite, Energy
[59] X. Zhu, Y.-P. Zhao, Atomic mechanisms and equation of state of methane & Fuel 30 (11) (2016) 9114–9121, https://doi.org/10.1021/acs.
adsorption in carbon nanopores, J. Phys. Chem. C. 118 (31) (2014) energyfuels.6b01777.
17737–17744, https://doi.org/10.1021/jp5047003. [84] T. Zhang, G.S. Ellis, S.C. Ruppel, K. Milliken, R. Yang, Effect of organic-matter
[60] S. Riewchotisakul, I.Y. Akkutlu, Adsorption-enhanced transport of hydrocarbons type and thermal maturity on methane adsoption in shale-gas systems, Org.
in organic nanopores, Spe J. 21 (2016) 1960–1969, https://doi.org/10.2118/ Geochem. 47 (2012) 120–131, https://doi.org/10.1016/j.
175107-PA. orggeochem.2012.03.012.
[61] M. Kazemi, A. Takbiri-Borujeni, M. Mansouri-Boroujeni, T. Sun, Effect of water in [85] J.-S. Bae, S.K. Bhatia, High-pressure adsorption of methane and carbon dioxide on
trasnport and storage of non-aqueous species in kerogen, Proc. - SPE Annu. Tech. coal, Energy Fuels 20 (6) (2006) 2599–2607, https://doi.org/10.1021/
Conf. Exhib. (2017), https://doi.org/10.2118/187087-ms. ef060318y.
[62] Y. Yang, J. Liu, J. Yao, J. Kou, Z. Li, T. Wu, K. Zhang, L. Zhang, H. Sun, Adsorption [86] J.-S. Bae, S.K. Bhatia, V. Rudolph, P. Massarotto, Pore accessibility of methane
behaviors of shale oil in kerogen slit by molecular simulation, Chem. Eng. J. 387 and carbon dioxide in coals, Energy Fuels 23 (6) (2009) 3319–3327, https://doi.
(2020) 124054, https://doi.org/10.1016/j.cej.2020.124054. org/10.1021/ef900084b.
[63] Z. Sun, X. Li, W. Liu, T. Zhang, M. He, H. Nasrabadi, Molecular dynamics of [87] S. Wang, Q. Feng, F. Javadpour, M. Zha, R. Cui, Multiscale modeling of gas
methane flow behavior through realistic organic nanopores under geologic shale transport in shale matrix: An integrated study of molecular dynamics and rigid-
condition: pore size and kerogen types, Chem. Eng. J. 398 (2020) 124341, pore-network model, Spe J. (2020) 1416–1442, https://doi.org/10.2118/
https://doi.org/10.1016/j.cej.2020.124341. 187286-PA.
[64] S.R. Kelemen, M. Afeworki, M.L. Gorbaty, M. Sansone, P.J. Kwiatek, C.C. Walters, [88] H. Wu, J. Chen, H.e. Liu, Molecular dynamics simulations about adsorption and
H. Freund, M. Siskin, A.E. Bence, D.J. Curry, M. Solum, R.J. Pugmire, displacement of methane in carbon nanochannels, J. Phys. Chem. C. 119 (24)
M. Vandenbroucke, M. Leblond, F. Behar, Direct characterization of kerogen by x- (2015) 13652–13657, https://doi.org/10.1021/acs.jpcc.5b02436.
ray and solid-state **13C nuclear magnetic resonance methods, Energy Fuels 21 [89] K. Mosher, J. He, Y. Liu, E. Rupp, J. Wilcox, Molecular simulation of methane
(3) (2007) 1548–1561, https://doi.org/10.1021/ef060321h. adsorption in micro- and mesoporous carbons with applications to coal and gas
[65] P. Ungerer, J. Collell, M. Yiannourakou, Molecular modeling of the volumetric shale systems, Int. J. Coal Geol. 109–110 (2013) 36–44, https://doi.org/10.1016/
and thermodynamic properties of kerogen: Influence of organic type and j.coal.2013.01.001.
maturity, Energy Fuels 29 (1) (2015) 91–105, https://doi.org/10.1021/ [90] J. Zhang, S.K. Choi, Molecular dynamics simulation of methane in potassium
ef502154k. montmorillonite clay hydrates, J. Phys. B At. Mol. Opt. Phys. 39 (18) (2006)
[66] D.M. Jarvie, R.J. Hill, T.E. Ruble, R.M. Pollastro, Unconventional shale-gas 3839–3848, https://doi.org/10.1088/0953-4075/39/18/013.
systems: The Mississippian Barnett Shale of north-central Texas as one model for [91] L.u. Zhang, C. Liu, Q. Li, Molecular simulations of competitive adsorption
thermogenic shale-gas assessment, Am. Assoc. Pet. Geol. Bull. 91 (4) (2007) behavior between CH4-C2H6 in K-illite clay at supercritical conditions, Fuel 260
475–499, https://doi.org/10.1306/12190606068. (2020) 116358, https://doi.org/10.1016/j.fuel.2019.116358.
[67] J. Collell, P. Ungerer, G. Galliero, M. Yiannourakou, F. Montel, M. Pujol, [92] M.D. Donohue, G.L. Aranovich, Classification of Gibbs adsorption isotherms, Adv.
Molecular simulation of bulk organic matter in type II shales in the middle of the Colloid Interface Sci. 76 (1998) 137–152, https://doi.org/10.1016/S0001-8686
oil formation window, Energy Fuels 28 (2014) 7457–7466, https://doi.org/ (98)00044-X.
10.1021/ef5021632. [93] Y. Kurniawan, S.K. Bhatia, V. Rudolph, Simulation of binary mixture adsorption
[68] T. Wang, S. Tian, G. Li, M. Sheng, Selective adsorption of supercritical carbon of methane and CO2 at supercritical conditions in carbons, AIChE J. 52 (3) (2006)
dioxide and methane binary mixture in shale kerogen nanopores, J. Nat. Gas Sci. 957–967, https://doi.org/10.1002/(ISSN)1547-590510.1002/aic.v52:310.1002/
Eng. 50 (2018) 181–188, https://doi.org/10.1016/j.jngse.2017.12.002. aic.10687.
[69] S.L. Montgomery, D.M. Jarvie, K.A. Bowker, R.M. Pollastro, Mississippian Barnett [94] F. Karavias, A.L. Myers, Isosteric heats of multicomponent adsorption:
Shale, Fort Worth basin, north-central Texas: Gas-shale play with multi–trillion thermodynamics and computer simulations, Langmuir 7 (12) (1991) 3118–3126,
cubic foot potential, Am. Assoc. Pet. Geol. Bull. 89 (2) (2005) 155–175, https:// https://doi.org/10.1021/la00060a035.
doi.org/10.1306/09170404042. [95] H. Pan, J.A. Ritter, P.B. Balbuena, Examination of the approximations used in
[70] J. Zumberge, K. Ferworn, S. Brown, Isotopic reversal (’rollover’) in shale gases determining the isosteric heat of adsorption from the Clausius− Clapeyron
produced from the Mississippian Barnett and Fayetteville formations, Mar. Pet. equation, Langmuir 14 (21) (1998) 6323–6327, https://doi.org/10.1021/
Geol. 31 (1) (2012) 43–52, https://doi.org/10.1016/j.marpetgeo.2011.06.009. la9803373.
[71] H.J. Sun, COMPASS: An ab initio force-field optimized for condensed-phase [96] T. Vuong, P.A. Monson, Monte Carlo simulation studies of heats of adsorption in
applications overview with details on alkane and benzene compounds, J. Phys. heterogeneous solids, Langmuir 12 (22) (1996) 5425–5432, https://doi.org/
Chem. B. 102 (1998) 7338–7364, https://doi.org/10.1021/jp980939v. 10.1021/la960325m.
[72] D.-Y. Peng, D.B. Robinson, A new two-constant equation of state, Ind. Eng. Chem. [97] J. Liu, W.G. Chapman, Thermodynamic modeling of the equilibrium partitioning
Fundam. 15 (1) (1976) 59–64, https://doi.org/10.1021/i160057a011. of hydrocarbons in nanoporous kerogen particles, Energy Fuels 33 (2) (2019)
[73] A. Sharma, S. Namsani, J.K. Singh, Molecular simulation of shale gas adsorption 891–904, https://doi.org/10.1021/acs.energyfuels.8b03771.
and diffusion in inorganic nanopores, Mol. Simul. 41 (5-6) (2015) 414–422, [98] Y. Liu, J. Wilcox, Effects of surface heterogeneity on the adsorption of CO2 in
https://doi.org/10.1080/08927022.2014.968850. microporous carbons, Environ. Sci. Technol. 46 (3) (2012) 1940–1947, https://
[74] Y. Liu, H.A. Li, Y. Tian, Z. Jin, H. Deng, Determination of the absolute adsorption/ doi.org/10.1021/es204071g.
desorption isotherms of CH4 and n-C4H10 on shale from a nano-scale perspective, [99] S. Wang, F. Javadpour, Q. Feng, Fast mass transport of oil and supercritical
Fuel 218 (2018) 67–77, https://doi.org/10.1016/j.fuel.2018.01.012. carbon dioxide through organic nanopores in shale, Fuel 181 (2016) 741–758,
[75] E. Paquet, H.L. Viktor, Molecular dynamics, monte carlo simulations, and https://doi.org/10.1016/j.fuel.2016.05.057.
langevin dynamics: A computational review, BioMed Res. Int. 2015 (2015) 1–18, [100] P. Billemont, B. Coasne, G. De Weireld, Adsorption of carbon dioxide, methane,
https://doi.org/10.1155/2015/183918. and their mixtures in porous carbons: Effect of surface chemistry, water content,
[76] D. Dubbeldam, A. Torres-Knoop, K.S. Walton, On the inner workings of Monte and pore disorder, Langmuir 29 (10) (2013) 3328–3338, https://doi.org/
Carlo codes, Mol. Simul. 39 (14-15) (2013) 1253–1292, https://doi.org/10.1080/ 10.1021/la3048938.
08927022.2013.819102. [101] W. Song, J. Yao, J. Ma, A. Li, Y. Li, H. Sun, L. Zhang, Grand canonical Monte Carlo
[77] Z. Jin, A. Firoozabadi, Effect of water on methane and carbon dioxide sorption in simulations of pore structure influence on methane adsorption in micro-porous
clay minerals by Monte Carlo simulations, Fluid Phase Equilib. 382 (2014) 10–20, carbons with applications to coal and shale systems, Fuel 215 (2018) 196–203,
https://doi.org/10.1016/j.fluid.2014.07.035. https://doi.org/10.1016/j.fuel.2017.11.016.
[78] Y. Tian, C. Yan, Z. Jin, Characterization of methane excess and absolute [102] J.F. Denayer, I. Daems, G.V. Baron, Adsorption and reaction in confined spaces,
adsorption in various clay nanopores from molecular simulation, Sci. Rep. 7 Oil Gas Sci. Technol. 61 (4) (2006) 561–569, https://doi.org/10.2516/ogst:
(2017) 12040, https://doi.org/10.1038/s41598-017-12123-x. 2006025a.
[79] H. Sun, H. Zhao, N.a. Qi, Y. Li, Molecular insights into the enhanced shale gas [103] D. Keffer, H.T. Davis, A.V. McCormick, The effect of nanopore shape on the
recovery by carbon dioxide in kerogen slit nanopores, J. Phys. Chem. C 121 (18) structure and isotherms of adsorbed fluids, Adsorption 2 (1) (1996) 9–21, https://
(2017) 10233–10241. doi.org/10.1007/BF00127094.
[80] K. Bui, I.Y. Akkutlu, Hydrocarbons recovery from model-kerogen nanopores, Spe [104] D. Cao, J. Wu, Self-diffusion of methane in single-walled carbon nanotubes at
J. (2017) 854–862, https://doi.org/10.2118/185162-pa. suband supercritical conditions, Langmuir 20 (9) (2004) 3759–3765, https://doi.
org/10.1021/la036375q.

17

You might also like