Next Article in Journal
15N Natural Abundance Characteristics of Ammonia Volatilization from Soils Applied by Different Types of Fertilizer
Next Article in Special Issue
Factors Influencing the CO2 Corrosion Pattern of Oil–Water Mixed Transmission Pipeline during High Water Content Period
Previous Article in Journal
Ozone-Induced Lung and Bronchial Injury: A Mouse Model Study
Previous Article in Special Issue
Study on Enhancing Shale Oil Recovery by CO2 Pre-Pad Energized Fracturing in A83 Block, Ordos Basin
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molecule Simulation of CH4/CO2 Competitive Adsorption and CO2 Storage in Shale Montmorillonite

1
School of Geoscience and Technology, Southwest Petroleum University, Chengdu 610500, China
2
College of Energy, Chengdu University of Technology, Chengdu 610059, China
3
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Chengdu University of Technology, Chengdu 610059, China
4
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation, Southwest Petroleum University, Chengdu 610500, China
*
Author to whom correspondence should be addressed.
Atmosphere 2022, 13(10), 1565; https://doi.org/10.3390/atmos13101565
Submission received: 28 July 2022 / Revised: 6 September 2022 / Accepted: 21 September 2022 / Published: 25 September 2022
(This article belongs to the Special Issue CO2 Geological Storage and Utilization)

Abstract

:
The main source of production in the middle and late stages of shale gas extraction is the adsorbed gas in shale, and the adsorbed gas in shale mainly comes from organic matter casein and clay minerals in shale; therefore, this paper uses sodium-based montmorillonite to characterize the clay minerals in shale and study the CH4 adsorption law in clay minerals, and this study has certain guiding significance for shale gas extraction. In addition, this paper also conducts a study on the competitive adsorption law of CH4 and CO2, and at the same time, predicts the theoretical sequestration of CO2 in shale clay minerals, which is a reference value for the study of CO2 burial in shale and is beneficial to the early realization of carbon neutral. In this paper, the slit model of sodium-based montmorillonite and the fluid model of CH4 and CO2 were constructed using Materials Studio software, and the following two aspects were studied based on the Monte Carlo method: Firstly, the microscopic adsorption behavior of CH4 in sodium-based montmorillonite was studied, and the simulations showed that the adsorption capacity of montmorillonite decreases with increasing temperature, increases and then decreases with increasing pressure, and decreases with increasing pore size. CH4 has two states of adsorption and free state in the slit. The adsorption type of CH4 in montmorillonite is physical adsorption. Secondly, the competitive adsorption of CH4 and CO2 in sodium-based montmorillonite was studied, and the simulations showed that the CO2 repulsion efficiency increased with increasing CO2 injection pressure, and the CO2/CH4 competitive adsorption ratio decreased with increasing pressure. The amount of CO2 storage decreased with increasing temperature and increased with increasing CO2 injection pressure.

1. Introduction

Shale oil and gas are kinds of unconventional oil and gas resources with considerable geological reserves. The adsorbed state is one of the main states of CH4 occurrence in shales (the other state is the free state). The proportion of adsorbed gas in different shale gas reservoirs varies from 20% to 85% of the whole reservoir, with an average of about 50% [1]. The pore structure of shale gas reservoirs is complex, with the most developed micro- and nano-pores; therefore, the storage and transport of shale gas are very different compared to conventional gas. Conventional core experiments do not accurately describe the gas adsorption characteristics and mechanisms in shale, while molecular simulations can accurately study the adsorption patterns of molecules under micro- and nano-pores and make quantitative characterization for the adsorption studies of shale gas.
Liu, Y., Liu, B., Xiong, J., Srinivas, G., et al. conducted molecular simulations of CH4 adsorption in graphite oxide slits using graphene oxide instead of kerogen and summarized the effects of temperature, pressure, pore size, and functional group type on CH4 adsorption [2,3,4,5]. Jin Z.H., Xu C.X., Xiong, J., Liang, L.X., et al. constructed molecular models of clastic minerals such as quartz and studied the microscopic adsorption patterns of shale gas in the pores of clastic minerals [6,7,8,9]. Babatunde, K.A., Tian, S.C., Sui, H., Katti, D.R., Tang, X., Shi, Y., Collell, J., Ru, X., Ungerer, P. et al. developed molecular models of different types of kerogen by molecular simulation software, analyzed the influence of variables such as temperature and pressure on the adsorption capacity of caseinates, ranked the adsorption capacity of different types of caseinates, and revealed the influence of water content changes on the adsorption of multi-component gases in caseinates. The effect of water content on the adsorption of multicomponent gases in the roots was revealed [10,11,12,13,14,15,16,17,18]. Xiong, J., Feng, D., Lv, Z.L., Huang, T., Ren, J.H., Greathouse, J.A., Li, W., et al. developed molecular models of clay minerals such as montmorillonite, illite, and kaolinite by molecular simulation software, and the effects of pore size, temperature, water content, and different compositions on the adsorption behavior of CH4 and CO2 in the slit model of each mineral type were discussed, and the strength of the adsorption capacity of the three clay minerals was summarized [19,20,21,22,23,24,25]. In summary, initially, some scholars proposed to replace the dry casein with graphene-containing oxygen functional groups to make it closer to the organic matter structure of shale. In recent years, some scholars have established specific kerogen models and clay mineral models to study the adsorption pattern and mechanism of CH4 in them and analyze the effects of pore size, temperature, and pressure on adsorption.
At the present stage, scholars in China and abroad have mainly conducted research on the adsorption of organic matter and clay minerals in shale, and these studies did not delve into the effects of factors such as temperature and pressure on the competitive adsorption of CH4 and CO2. Therefore, this paper focuses on the microscopic adsorption behavior of CH4 in sodium-based montmorillonite and the competitive adsorption behavior of CH4 and CO2. The study of the microscopic adsorption behavior of CH4 provides a theoretical basis for the understanding of shale gas seepage patterns and storage characteristics; the study of the competitive adsorption behavior of CH4 and CO2 not only provides a theoretical basis for the enhanced extraction of shale gas by CO2 injection but also reveals the sequestration capacity of shale clay minerals for CO2 and provides a research idea for the geological storage of CO2, and geological carbon sequestration is a way of CO2 emission reduction.

2. Simulation Model and Potential Energy Parameters

2.1. Model Building

Shale has the following three main components: organic matter, clay minerals, and non-clay minerals, while clay minerals are dominated by illite, montmorillonite, and illite/smectite formation [26,27,28,29,30]. In this paper, a montmorillonite cell model (shown in Figure 1a) was constructed by using the molecular simulation software, Material Studio, with sodium-based montmorillonite as the skeleton. The specific parameters of this cell model are a = 5.23 Å, b = 9.06 Å, c = 12.5 Å, α = 90°, β = 99°, and γ = 90°, and the atomic coordinate parameters are shown in Table 1 [31]. The montmorillonite slit model was then established by the steps of the tessellation of the cell, establishing the interface model, and performing supercell (shown in Figure 1c), where H denotes the pore size of the slit model, and the values of H taken in this paper are 1 nm, 2 nm, 3 nm, 4 nm, 5 nm, 6 nm, 8 nm, and 10 nm, respectively. The specific parameters of the final slit model are x = 52.30 Å, y = 45.30 Å, and z varies from 32 Å to 122 Å with different slit aperture sizes. To study the competitive adsorption pattern of CH4 and CO2 in sodium-based montmorillonite, this paper also established CH4 and CO2 models (shown in Figure 1d,e) based on the establishment of sodium-based montmorillonite. The bond length and bond angle parameters of the two fluid models are labeled in Figure 1d,e.

2.2. Potential Energy Selection and Parameters

In this paper, the L-J potential function is used for the simulation, and the various potential and charge parameters that appear in the simulation are shown in Table 2 [32]. Both the montmorillonite model and the two-fluid models are assumed to be rigid bodies during the simulation because both CH4 molecules and CO2 molecules are electrically neutral, but the sodium ions in the montmorillonite model are positively charged, so the interaction between the two fluids and the atoms in montmorillonite needs to consider long-range charge Coulomb forces in addition to short-range van der Waals forces, and the L-J potential function is described as [33]:
E = ε i j [ ( δ i j r i j ) 12 ( δ i j r i j ) 6 ] + q i q j 4 π ε 0 r i j
where: δ and ε denote the potential energy parameter, which is related to the type of interacting atoms; r i j denotes the distance between two atoms, nm; q i and q j denote the atomic charge in the system, C; ε 0 denotes the dielectric constant, 8.854 × 10−12 F/m.

2.3. Model Verification

A low-pressure isothermal nitrogen adsorption simulation was performed using the established montmorillonite model, and the results obtained from the simulation were compared with those obtained from sodium-rich montmorillonite nitrogen adsorption experiments in Wyoming, USA [33] to verify the authenticity of the model. The results of the comparison are shown in Figure 2 (both experimental and simulated conditions are T = 334 K and P = 0–30 MPa), and it can be concluded that although there is an error between the isothermal adsorption simulation results and the experimental results, the error between the two is small, within 10%, indicating that the established sodium-rich montmorillonite model is more realistic and can represent the actual situation in the subsurface. The reason for the error is that the real sodium-rich montmorillonite in the subsurface has a more complex pore structure compared with the model.

3. Study on Adsorption Law of CH4

3.1. Simulation Methods and Conditions

In this paper, two molecular simulation methods are used: the first one is the giant regular Monte Carlo simulation, and the second one is the molecular dynamics simulation, based on which the Sorption module of the Material Studio software is used to realize the simulation of CH4 adsorption by montmorillonite. The total number of simulation steps was 1 × 106, of which the first 1.0 × 105 steps were used for the equilibrium simulation of the system and the second 9.0 × 105 steps were used for the process simulation. Six temperature points (278 K, 298 K, 318 K, 338 K, 358 K, and 378 K), six pressure points (5 MPa, 10 MPa, 15 MPa, 20 MPa, 25 MPa, and 30 MPa), and eight pore size points (1 nm, 2 nm, 3 nm, 4 nm, 5 nm, 6 nm, 8 nm, and 10 nm) were set for the simulation.
The critical temperature and pressure of CH4 are 190.55 K and 4.59 MPa, and the temperature and pressure points simulated in this paper are higher than the critical temperature and pressure of CH4. The CH4 is in a supercritical state during the simulation, and there is a large gap between the total adsorbed gas and the excess adsorbed gas in this state. Therefore, the total adsorbed gas needs to be converted into excess adsorbed gas [34]. The conversion formula is as follows:
n e x = N ρ g ( V a + V g ) × 10 3 / M S
where n e x is the excess adsorption capacity, mmol/m2; N is the simulated total gas volume, mmol/m2; ρ g is the gas phase density, g/cm3; V a + V g is the free space volume of the adsorption system, cm3; M is the molecular mass of the gas, g/mol; S is the specific surface area of the cell, m2.

3.2. Analysis of Simulation Results

(1)
Effect of pressure on CH4 adsorption under different pore sizes of montmorillonite slit
Figure 3 shows the total adsorption and excess adsorption of CH4 at different pressures with different slit pore sizes, respectively. From Figure 3b, it can be found that the adsorption capacity gradually increases from 0.0 MPa to 10.0 MPa when the pressure increases and reaches the peak at 10 MPa, then shows a decreasing trend in the interval from 10.0 MPa to 30.0 MPa. Since the adsorption of CH4 by montmorillonite slit is an interfacial phenomenon, the excess adsorption capacity will rise first and then fall. During the adsorption process, two kinds of forces exist at the interfacial interface between the montmorillonite slit wall and CH4. One is the mutual gravitational force between CH4 molecules, which is called gravitational force A. The other is the gravitational force of the montmorillonite slit wall on CH4, which is called gravitational force B. When gravitational force A is smaller than gravitational force B, a potential trap occurs, and most of the CH4 accumulates on both sides of the montmorillonite slit under the action of gravitational force B. This is due to the imbalance of gravitational forces. This phenomenon of the uneven density distribution of adsorbents due to the unbalanced gravitational force is the adsorption phenomenon. When the pressure is greater than 10.0 MPa, the bulk phase density of CH4 becomes larger, so the gravitational force A gradually becomes larger, and the corresponding gravitational force B starts to decrease. This change in the gravitational force relationship directly affects the density distribution of CH4 and the excess adsorption of montmorillonite slits. It is also shown in Figure 3b that the larger the pore size of the montmorillonite slit, the lower the capacity of the excess adsorption of CH4. The three-dimensional configuration of the montmorillonite slit for some pore sizes is shown in Figure 4.
(2)
Effect of temperature on CH4 adsorption under different pore sizes of montmorillonite slits
Figure 5 shows the values of CH4 adsorption at different temperatures for different montmorillonite slit pore sizes, respectively. Figure 5b shows that the excess adsorption decreases with increasing temperature. This is because, when the temperature increases, the effect of molecular kinetic energy change on the adsorption capacity of montmorillonite takes the major part, and the adsorption capacity is inversely proportional to the change of temperature, with the higher the temperature, the lower the adsorption capacity instead. From Figure 5b, it is easy to find that the adsorption capacity peak at the temperature of 278.0 K. From the comparison of the excess adsorption capacity of different montmorillonite slit pore sizes, the best adsorption effect of montmorillonite on CH4 was observed at the pore size of 1 nm and the worst at 10 nm.
(3)
Density variation under different pore sizes of montmorillonite slit
Figure 6 shows the density comparison of CH4 under different pore sizes of the montmorillonite slit. It can be found in Figure 6 that the peak of CH4 density distribution is located near the wall of the montmorillonite slit. Moreover, when the width of the oxidized montmorillonite slit pore size is 1 nm, most of the CH4 is attached to both sides of the slit and exists in the adsorption state. When the slit width is 2~5 nm, most of the CH4 is still attached to the slit wall, but some of it is also dispersed in the pores in the middle of the slit. When the slit width is 6~10 nm, the density of CH4 dispersed in the pores increases some more compared with 2~5 nm. Therefore, there are two occurrences of CH4 in the slit of montmorillonite with different pore sizes. In the small pore size, CH4 is mainly in the adsorbed state, and as the slit width increases, the free state of CH4 becomes more, and the two occurrences in the slit will coexist.
(4)
Heat of adsorption
The heat generated by the adsorption process is the heat of adsorption, and the value of the heat of adsorption can measure the degree of the strength of adsorption. Higher values of the heat of adsorption indicate stronger adsorption. Figure 7 represents the average equivalent heat of adsorption of CH4 at different montmorillonite slit pore sizes. From Figure 7, it can be found that the larger the slit pore size is, the less the average equivalent heat of adsorption of CH4 is, and the decreasing trend of the heat of adsorption tends to level off gradually. When the slit pore size was 1 nm, the average equivalent heat of adsorption of CH4 was 23.458 KJ/mol, and when the slit pore size was 10 nm, the average equivalent heat of adsorption of CH4 was 8.849 KJ/mol, both of which were lower than the standard for chemisorption (42.0 KJ/mol), indicating that the adsorption of CH4 on different montmorillonite slit pore sizes was physical adsorption. In addition, the adsorption heat curve also shows that the adsorption capacity of the small pore size slit is stronger than that of the large pore size slit, and the adsorption capacity of the slit decreases with the increase in pore size.

4. Study of Competitive Adsorption Pattern of CH4 and CO2

4.1. Simulation Method and Conditions

The pore size distribution curve of montmorillonite samples (Figure 8) shows that the pore size of montmorillonite is relatively small, dominated by micropores and mesopores less than 10 nm, and peaks between 5 nm and 6 nm, which indicates that the size of montmorillonite pores in shale is mainly concentrated in 5~6 nm [26]. Therefore, in this paper, a montmorillonite slit of 5 nm was used to conduct a simulation study of the competitive adsorption of CH4 and CO2. The simulation method used is consistent with Section 2.1. Two sets of conditions were set for the simulation: one set with a constant temperature of 298 K, the constant initial pressure of 5 MPa for CH4, and eight pressure points (0 MPa to 7 MPa) for CO2 pressure, whose group purpose was to analyze the CO2 adsorption capacity in the mixed phase of CH4 and CO2 and to predict the theoretical CO2 sequestration in montmorillonite. In another group with a constant temperature of 298 K and the initial pressures of CH4 and CO2 set to the same value, with 6 pressure points (5~10 MPa), the purpose of this group is to calculate the CO2/CH4 competitive adsorption ratio.

4.2. Analysis of Simulation Results

(1)
CO2 adsorption capacity in mixed-phase
Figure 9 shows the total adsorption and excess adsorption curves of the two fluids with a constant initial pressure of CH4 and varying initial pressure of CO2. From Figure 9, it can be found that the excess adsorption of CH4 gradually decreases (from 0.019 mmol/m2 to 0.006 mmol/m2) and the excess adsorption of CO2 gradually increases (from 0.00 mmol/m2 to 0.052 mmol/m2) as the initial pressure of CO2 increases. In addition, the CO2-repelling CH4 efficiency curve (shown in Figure 10) was obtained by computational processing in this paper, and Figure 10 shows that the CO2 repelling efficiency increases with the increase in the initial pressure of CO2. Figure 11 shows the comparison curves of the CO2 storage volume at different temperatures and different CO2 injection pressures, from which it can be concluded that the CO2 storage volume gradually decreases with the increase in temperature, and the higher the temperature, the smaller the decrease in storage volume. With the increase in CO2 injection pressure, the storage volume gradually increases, and the higher the pressure, the smaller the increase in storage volume.
(2)
CO2/CH4 competitive adsorption ratio
The simulation was carried out under the condition that the initial pressures of CH4 and CO2 were equal and increased synchronously. The total and excess adsorption of CH4 and CO2 at each pressure point are shown in Figure 12. As the total pressure gradually increased from 10 MPa to 20 MPa, the excess sorption of CH4 increased from 0.007 mmol/m2 to 0.013 mmol/m2, the excess sorption of CO2 increased from 0.014 mmol/m2 to 0.019 mmol/m2, and the excess sorption of both CH4 and CO2 increased with the increase in pressure. Figure 13 shows the scatter plot of the pressure versus CO2/CH4 competitive adsorption ratio and its regression curve, from which it can be concluded that the CO2/CH4 competitive adsorption ratio gradually decreases with the increase in pressure.

5. Conclusions

(1)
At constant temperature, the excess adsorption of CH4 in the montmorillonite slit increases first and then decreases with increasing pressure and reaches a peak between 10.0 MPa and 15.00 MPa. At constant pressure, the excess adsorption decreases gradually with increasing temperature and is at a peak at 278.0 K. The adsorption effect decreases as the pore size becomes larger, with the maximum adsorption of 1 nm pore size.
(2)
Comparing the changes in the CH4 density at the 1 nm~10 nm pore size, it was found that the peak of the CH4 density appeared at the slit wall after the simulation of the adsorption process was performed. The CH4 in the slit of montmorillonite has two occurrences and is mainly in the adsorption state. With the increase in the slit width, the free state of CH4 becomes more and the slit appears to have two occurrences coexisting.
(3)
The adsorption of CH4 on the slit of montmorillonite is physical adsorption; the adsorption capacity of the slit decreases with increasing pore size.
(4)
When the adsorbent is the coexistence of CH4 and CO2, the initial pressure of CH4 is constant, and the repulsion efficiency of CO2 increases with the increase in the initial pressure of CO2; it is also analyzed that the amount of CO2 storage decreases with the increase in temperature and increases with the increase in injection pressure.
(5)
When the initial pressure of CH4 and CO2 are equal, it can be concluded from the curve of pressure and the CO2/CH4 competitive adsorption ratio that the competitive adsorption ratio is negatively correlated with the pressure, and the higher the pressure the lower the competitive adsorption ratio.

Author Contributions

Data curation, X.Q.; Funding acquisition, H.T. and L.S.; Investigation, F.G., H.T. and X.Q.; Methodology, D.H. and H.T.; Project administration, J.G. and L.S.; Software, F.G. and X.Q.; Supervision, J.G. and L.S.; Validation, D.H. and J.G.; Writing—original draft, D.H.; Writing—review & editing, D.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by China Postdoctoral Foundation, No. 2017M612995; Applied Basic Research Project of Sichuan Provincial Science and Technology Department, No. 2021YJ0352; State Key Laboratory of Oil and Gas Reservoir Geology and Development Engineering Free Exploration Project, No. CZ201910.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets supporting the conclusions of this article are private and came from the Chengdu University of Technology, Chengdu, China.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Curtis, J.B. Fractured shale-gas systems. AAPG Bull. 2002, 86, 1921–1938. [Google Scholar]
  2. Liu, Y.Y.; Wilcox, J. Effects of surface heterogeneity on the adsorption of CO2 in microporous carbons. Environ. Sci. Technol. 2012, 46, 1940–1947. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, B.; Shi, J.Q. A molecular dynamics simulation of methane adsorption in graphite slit-pores. Chin. J. Comput. Phys. 2013, 30, 692–699. [Google Scholar]
  4. Xiong, J.; Liu, X.J.; Liang, L.X.; Zeng, Q. Methane adsorption on carbon models of the organic matter of organic-rich shales. Energy Fuels 2017, 31, 1489–1501. [Google Scholar] [CrossRef]
  5. Srinivas, G.; Burress, J.W.; Ford, J.; Yildirim, T. Porous graphene oxide frameworks: Synthesis and gas sorption properties. J. Mater. Chem. 2011, 21, 11323–11329. [Google Scholar] [CrossRef]
  6. Jin, Z.H.; Firoozabadi, A. Methane and carbon dioxide adsorption in clay-like slit pores by Monte Carlo simulations. Fluid Phase Equilibria 2013, 360, 456–465. [Google Scholar] [CrossRef]
  7. Xiong, J.; Liu, X.; Liang, L.; Zeng, Q. Adsorption of methane in organic-rich shale nanopores: An experimental and molecular simulation study. Fuel 2017, 200, 299–315. [Google Scholar] [CrossRef]
  8. Liang, L.; Luo, D.; Liu, X.; Xiong, J. Experimental study on the wettability and adsorption characteristics of Longmaxi Formation shale in the Sichuan Basin, China. J. Nat. Gas Sci. Eng. 2016, 33, 1107–1118. [Google Scholar] [CrossRef]
  9. Xu, C.X.; Xue, H.T.; Li, H.B. Microscopic adsorption mechanism difference in the mineral pore of shale gas reservoir. Spec. Oil Gas Reserv. 2020, 27, 79–84. [Google Scholar]
  10. Tian, S.C.; Wang, T.Y.; Li, G.S.; Sheng, M.; Ren, W.X. Molecular simulation of methane adsorption behavior in different shale kerogen types. Nat. Gas Ind. 2017, 37, 18–25. [Google Scholar]
  11. Sui, H.G.; Yao, J. Effect of surface chemistry for CH4/CO2 adsorption in kerogen: A molecular simulation study. J. Nat. Gas Sci. Eng. 2016, 31, 738–746. [Google Scholar] [CrossRef]
  12. Katti, D.R.; Thapa, K.B.; Katti, K.S. Modeling molecular interactions of sodium montmorillonite clay with 3D kerogen models. Fuel 2017, 199, 641–652. [Google Scholar] [CrossRef]
  13. Tang, X.; Zhou, X.Y.; Peng, Y.J. Molecular simulation of methane adsorption within illite minerals in the Longmaxi Formation shale based on a grand canonical Monte Carlo method and the pore size distribution in southeastern Chongqing, China. J. Nat. Gas Geosci. 2019, 4, 111–119. [Google Scholar] [CrossRef]
  14. Shi, Y.; Yang, X.N.; Li, S.G.; Yan, M.; Shuang, H.Q.; Xue, J.H. Effect of moisture on adsorption and diffusion of multi-component gas in kerogen: A molecular simulation study. J. Xi’an Shiyou Univ. 2021, 36, 50–57. [Google Scholar]
  15. Collell, J.; Galliero, G.; Gouth, F.; Montel, F.; Pujol, M.; Ungerer, P.; Yiannourakou, M. Molecular simulation and modelisation of methane/ethane mixtures adsorption onto a microporous molecular model of kerogen under typical reservoir conditions. Microporous Mesoporous Mater. 2014, 197, 194–203. [Google Scholar] [CrossRef]
  16. Ru, X.; Cheng, Z.; Song, L.; Wang, H.; Li, J. Experimental and computational studies on the average molecular structure of Chinese Huadian oil shale kerogen. J. Mol. Struct. 2012, 1030, 10–18. [Google Scholar] [CrossRef]
  17. Ungerer, P.; Collell, J.; Yiannourakou, M. Molecular Modeling of the Volumetric and Thermodynamic Properties of Kerogen: Influence of Organic Type and Maturity. Energy Fuels 2015, 29, 91–105. [Google Scholar] [CrossRef]
  18. Babatunde, K.A.; Negash Mojid, B.M.; Ahmed MR, T.Y.; Jufar, S.R. Molecular simulation study of CO2/CH4 adsorption on realistic heterogeneous shale surfaces. Appl. Surf. Sci. 2021, 543, 148789. [Google Scholar] [CrossRef]
  19. Xiong, J.; Liu, X.J.; Liang, L.X. Molecular simulation on the adsorption behaviors of methane in montmorillonite slit pores. Acta Pet. Sin. 2016, 37, 1021–1029. [Google Scholar]
  20. Feng, D.; Li, X.F.; Wang, X.Z.; Li, J.; Shi, J.T.; Zhang, T.; Li, P.H.; Chen, Y. Pore distribution characteristics and methane adsorption capacity of clay under different water content conditions. J. China Coal Soc. 2017, 42, 2402–2413. [Google Scholar]
  21. Lu, Z.L.; Ning, Z.F.; Wang, Q.; Huang, L.; Meng, H.; Yu, X.F.; Qin, H.B. Molecular simulation of methane adsorption behavior on clay minerals in shale. J. China Coal Soc. 2019, 44, 3117–3124. [Google Scholar]
  22. Huang, T.; Liu, D.H.; Sun, J. Simulation on adsorption of methane in micro-pore of clay minerals. Sci. Technol. Eng. 2020, 20, 141–145. [Google Scholar]
  23. Ren, J.H.; Ren, X.H.; Song, H.Q.; Han, D.L.; Wang, C.C.; Sheng, G.L.; Lu, W.F. Adsorption and diffusion characteristics of methane in nanopores based on molecular simulation. Acta Pet. Sin. 2020, 41, 1366–1375. [Google Scholar]
  24. Greathouse, J.A.; Cygan, R.T.; Fredrich, J.T.; Jerauld, G.R. Adsorption of Aqueous Crude Oil Components on the Basal Surfaces of Clay Minerals: Molecular Simulations Including Salinity and Temperature Effects. J. Phys. Chem. C 2017, 121, 22773–22786. [Google Scholar] [CrossRef]
  25. Li, W.; Pang, X.; Snape, C.; Zhang, B.; Zheng, D.; Zhang, X. Molecular Simulation Study on Methane Adsorption Capacity and Mechanism in Clay Minerals: Effect of Clay Type, Pressure, and Water Saturation in Shales. Energy Fuels 2019, 33, 765–778. [Google Scholar] [CrossRef]
  26. Xiong, J. Investigation of the Influences of the Methane Adsorption Capacity on the Shales; Southwest Petroleum University: Chengdu, China, 2015. [Google Scholar]
  27. Bowker, K.A. Recent developments of the Barnett Shale play, Fort Worth basin. West Tex. Geol. Soc. Bull. 2003, 42, 4–11. [Google Scholar]
  28. Ross, D.J.K.; Bustin, R.M. Characterizing the shale gas resource potential of Devonian–Mississippian strata in the Western Canada sedimentary basin: Application of an integrated formation evaluation. AAPG Bull. 2008, 92, 87–125. [Google Scholar] [CrossRef]
  29. Liang, C.; Jiang, Z.X.; Yang, Y.T.; Wei, X.J. Shale lithofacies and reservoir space of the Wufeng–Longmaxi formation, Sichuan Basin, China. Pet. Explor. Dev. 2012, 39, 736–743. [Google Scholar] [CrossRef]
  30. Li, J.; Yu, B.; Zhang, J.; Li, Y.; Wu, J. Reservoir characteristics and their influence factors of the Lower Cambrian dark shale in northern Guizhou. Oil Gas Geol. 2012, 33, 364–374. [Google Scholar]
  31. Zhang, Y.Y.; Chen, M.; Deng, Y.; Jin, Y.; Lu, Y.H.; Xia, Y. Molecular dynamics simulation of temperature and pressure effects on hydration characteristics of montmorillonites. J. Chin. Ceram. Soc. 2018, 46, 1489–1498. [Google Scholar]
  32. Wang, J.; Wang, J.X.; Zeng, F.G.; Wu, X.L. Molecular simulations of crystal structure conformation X-Ray diffraction and infra-red spectrum in montmorillonites. Acta Mineral. Sin. 2011, 31, 133–136. [Google Scholar]
  33. Wang, Z.H.; Zhang, Y.P.; Guo, P.; Li, F.H.; Bai, Y. Research on methane adsorption of pure clay minerals and evaluation of adsorption models. Sci. Technol. Eng. 2020, 20, 2209–2215. [Google Scholar]
  34. Xiong, J.; Liu, X.J.; Liang, L.X. Molecular simulation study on the adsorption behaviors of methane in slit-like clay mineral pore. J. China Coal Soc. 2017, 42, 959–968. [Google Scholar]
Figure 1. Molecular model diagram of competitive adsorption of CH4 and CO2 in sodium-based montmorillonite (a): Sodium-based montmorillonite crystal, (b): Sodium-based montmorillonite supercell crystal, (c): Montmorillonite slit, (d): Methane model, (e): Carbon dioxide model.
Figure 1. Molecular model diagram of competitive adsorption of CH4 and CO2 in sodium-based montmorillonite (a): Sodium-based montmorillonite crystal, (b): Sodium-based montmorillonite supercell crystal, (c): Montmorillonite slit, (d): Methane model, (e): Carbon dioxide model.
Atmosphere 13 01565 g001
Figure 2. Comparison between simulated and experimental results.
Figure 2. Comparison between simulated and experimental results.
Atmosphere 13 01565 g002
Figure 3. Methane adsorption capacity at different pore sizes (a): Total methane adsorption capacity, (b): Excess methane adsorption capacity.
Figure 3. Methane adsorption capacity at different pore sizes (a): Total methane adsorption capacity, (b): Excess methane adsorption capacity.
Atmosphere 13 01565 g003
Figure 4. Partial pore size montmorillonite 3D conformation (a): 1 nm, (b): 2 nm, (c): 3 nm, (d): 4 nm, (e): 5 nm.
Figure 4. Partial pore size montmorillonite 3D conformation (a): 1 nm, (b): 2 nm, (c): 3 nm, (d): 4 nm, (e): 5 nm.
Atmosphere 13 01565 g004
Figure 5. Methane adsorption capacity at different pore sizes (a): Total methane adsorption capacity, (b): Excess methane adsorption capacity.
Figure 5. Methane adsorption capacity at different pore sizes (a): Total methane adsorption capacity, (b): Excess methane adsorption capacity.
Atmosphere 13 01565 g005
Figure 6. Comparison of the density of methane at different pore sizes.
Figure 6. Comparison of the density of methane at different pore sizes.
Atmosphere 13 01565 g006
Figure 7. Average equivalent heat of adsorption at different pore sizes.
Figure 7. Average equivalent heat of adsorption at different pore sizes.
Atmosphere 13 01565 g007
Figure 8. Pore size distribution curve of montmorillonite samples ((a) obtained by BJH method, (b) obtained by NLDFT method) [18]. Reproduced with permission from Xiong J., Doctoral thesis of Southwest Petroleum University; published by Southwest Petroleum University, 2015.
Figure 8. Pore size distribution curve of montmorillonite samples ((a) obtained by BJH method, (b) obtained by NLDFT method) [18]. Reproduced with permission from Xiong J., Doctoral thesis of Southwest Petroleum University; published by Southwest Petroleum University, 2015.
Atmosphere 13 01565 g008
Figure 9. The adsorption capacity of two fluids (a): Total adsorption capacity, (b): Excess adsorption capacity.
Figure 9. The adsorption capacity of two fluids (a): Total adsorption capacity, (b): Excess adsorption capacity.
Atmosphere 13 01565 g009
Figure 10. Carbon dioxide repulsion efficiency.
Figure 10. Carbon dioxide repulsion efficiency.
Atmosphere 13 01565 g010
Figure 11. Relation curve between CO2 storage capacity and injection pressure.
Figure 11. Relation curve between CO2 storage capacity and injection pressure.
Atmosphere 13 01565 g011
Figure 12. The adsorption capacity of two fluids (a): Total adsorption capacity, (b): Excess adsorption capacity.
Figure 12. The adsorption capacity of two fluids (a): Total adsorption capacity, (b): Excess adsorption capacity.
Atmosphere 13 01565 g012
Figure 13. Competitive adsorption ratio versus pressure.
Figure 13. Competitive adsorption ratio versus pressure.
Atmosphere 13 01565 g013
Table 1. Spatial coordinates of each atom and cation in the cell.
Table 1. Spatial coordinates of each atom and cation in the cell.
Atomsc = 1.25 nmc = 1.53 nm
xyzxyz
Al03.0212.503.0215.5
Si0.4721.519.580.4721.5112.58
O0.12209.040.122012.04
O−0.6862.6159.24−0.6862.61512.04
O0.7721.5111.20.7721.5114.2
O(OH)0.8084.5311.250.8084.5314.25
H(OH)−0.1034.5310.812−0.1034.5313.182
Na+04.536.2504.539.25
Table 2. Potential energy parameters and charges of each atom [32] Reproduced with permission from Wang J, Acta Mineral. Sin.; published by the Chinese Academy of Sciences, 2011.
Table 2. Potential energy parameters and charges of each atom [32] Reproduced with permission from Wang J, Acta Mineral. Sin.; published by the Chinese Academy of Sciences, 2011.
TypesAtom Types(ε/KB)/Kσ/nmq/c
CH4CH4148.10.3730
CO2C28.1290.27570.6512
O80.5070.3033−0.3256
MontmorilloniteAl003
Mg002
Si31530.1841.2
O1560.317−1
Na+1000.2591
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hou, D.; Gong, F.; Tang, H.; Guo, J.; Qiang, X.; Sun, L. Molecule Simulation of CH4/CO2 Competitive Adsorption and CO2 Storage in Shale Montmorillonite. Atmosphere 2022, 13, 1565. https://doi.org/10.3390/atmos13101565

AMA Style

Hou D, Gong F, Tang H, Guo J, Qiang X, Sun L. Molecule Simulation of CH4/CO2 Competitive Adsorption and CO2 Storage in Shale Montmorillonite. Atmosphere. 2022; 13(10):1565. https://doi.org/10.3390/atmos13101565

Chicago/Turabian Style

Hou, Dali, Fengming Gong, Hongming Tang, Jianchun Guo, Xianyu Qiang, and Lei Sun. 2022. "Molecule Simulation of CH4/CO2 Competitive Adsorption and CO2 Storage in Shale Montmorillonite" Atmosphere 13, no. 10: 1565. https://doi.org/10.3390/atmos13101565

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop