10.8
CiteScore
 
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
10.8
CiteScore
5.3
Impact Factor
Generic selectors
Exact matches only
Search in title
Search in content
Post Type Selectors
Search in posts
Search in pages
Filter by Categories
Corrigendum
Current Issue
Editorial
Erratum
Full Length Article
Full lenth article
Letter to Editor
Original Article
Research article
Retraction notice
Review
Review Article
SPECIAL ISSUE: ENVIRONMENTAL CHEMISTRY
View/Download PDF

Translate this page into:

Original article
03 2021
:15;
103665
doi:
10.1016/j.arabjc.2021.103665

Investigation into the adsorption of CO2, N2 and CH4 on kaolinite clay

Yunnan Key Laboratory of Sino-German Blue Mining and Utilization of Special Underground Space, Faculty of Land Resources Engineering, Kunming University of Science and Technology, Kunming, Yunnan 650093, China
Key Laboratory of Safety and High-efficiency Coal Mining, Ministry of Education (Anhui University of Science and Technology), Huainan 232001, China
Sinohydro Engineering Bureau 8 Co. LTD., POWERCHINA, Changsha 410004, China
Department of Hydraulic Engineering, Tsinghua University, Beijing 100084, China
School of Civil and Transportation Engineering, Guangdong University of Technology, Guangzhou, Guangdong 510006, China
State Key Laboratory of Mountain Bridge and Tunnel Engineering, Chongqing Jiaotong University, Chongqing 400074, China

⁎Corresponding author. PangDD1989@126.com (Dongdong Pang)

Disclaimer:
This article was originally published by Elsevier and was migrated to Scientific Scholar after the change of Publisher.

Peer review under responsibility of King Saud University.

Abstract

Investigating the adsorption characteristics of CO2, N2 and CH4 on kaolinite clay is beneficial for enhanced shale gas recovery by gas injection. In this paper, the experiments of CO2, N2 and CH4 adsorption at 288 K, 308 K and 328 K on kaolinite clay were conducted, and the thermodynamics analysis of adsorption of three gases was performed. The findings reveal that the order of the uptakes of three gases on kaolinite clay is as follows: N2 < CH4 < CO2. Reducing temperature enlarges the separation coefficients of CO2 over CH4 (αCO2/CH4), CO2 over N2 (αCO2/N2), and CH4 over N2 (αCH4/N2). The value of αCO2/CH4 greater than one validates that CO2 is capable to directly replace the pre-adsorbed CH4. The spontaneity of CO2 adsorption is the highest, while N2 has the lowest adsorption spontaneity. Injecting N2 into gas-bearing reservoir can cause CH4 desorption by lowering the spontaneity of CH4 adsorption. Adsorbed CO2 molecules form a most ordered rearrangement on kaolinite surface. The decrease rate of entropy loss for N2 adsorption is higher than those for CO2 and CH4 adsorption.

Keywords

Adsorption
Shale gas
Kaolinite
Thermodynamics
CO2 sequestration
1

1 Introduction

Urgent demand for natural gas resources has stimulated the exploitation of unconventional gas, such as shale gas, coalbed gas and tight gas (Du et al., 2018, 2021a; Huang et al., 2018). Shale gas, occurring as free gas in pore fracture network or adsorbed gas on the matrix surface, has increasingly considered as the significant alternative for conventional gas (Du et al., 2017, 2019a, Qin et al., 2021a,b). At present, CO2 geological sequestration is one of the most prospective techniques for the reduction of CO2 emissions. Given that the higher adsorption capacity of CO2 over CH4 on shale reservoir, utilization of CO2 can promote the recovery of additional CH4 from gas shale formations with an added benefit of the creation of valuable sequestration option for greenhouse gas. Meanwhile, the injection of N2 into gas-bearing sediments also has the ability to trigger the desorption of substantial amounts of adsorbed CH4 by reducing CH4 partial pressure and breaking the equilibrium condition, and taking N2 as injection agent is also effective to improve gas production (Jessen et al., 2008). Therefore, owing to the feasibility, recovery cost and environment remediation, the technologies of enhanced shale gas recovery using CO2 (CO2-ESGR) and N2 (N2-ESGR) have gained lots of attentions (Du et al., 2019b; Huo et al., 2017; Liu et al., 2019; Zhao et al., 2020).

For shale reservoir, inorganic minerals and organic matter are the two main distinct media. As the important component of inorganic minerals, clay minerals have a remarkable effect on the adsorption and migration properties of fluids due to the larger surface area and complex pore structure (Jin and Firoozabadi, 2014). To optimize the implementation strategy of enhanced hydrocarbon recovery by gas injection, the reliable and accurate adsorption equilibrium data of CO2, N2 and CH4 on clay minerals (kaolinite, montmorillonite and illite) over a broad range of pressures and temperatures are necessary. Among the clay minerals, kaolinite clay is very common within shale formation (Zhang et al., 2018). Especially for some kaolinite-rich shale reservoir, the influence of kaolinite clay on fluid adsorption is significant. Therefore, comprehending the adsorption behaviors of fluids on kaolinite clay is meaningful for shale gas development and CO2 sequestration.

Recently, many contributions have been conducted to study the fluid adsorption on kaolinite clay by laboratory experiment and molecular simulation. For example, Heller and Zoback (2014) measured the isothermal sorption curves of CH4 and CO2 on kaolinite clay and discovered that the adsorption capacity of CO2 is obviously higher than that of CH4. Song et al. (2018) used molecular simulation method to investigate the effect of pore size on the adsorption of CH4 and CO2. They found that the uptakes of CH4 and CO2 on kaolinite clay both decrease with increasing pore size. Zhou et al. (2019) researched the competitive sorption behaviors of injected CO2 and pre-adsorbed CH4 through molecular simulation. Their results showed that CH4 adsorption on kaolinite clay can be obviously suppressed by injected CO2. Kang et al. (2020) analyzed the influences of temperature and pressure on CH4/CO2 adsorption on kaolinite clay and indicated that the interaction of CH4 with kaolinite is much lower than that of CO2 with kaolinite. Evidently, the relevant researches about fluid adsorption on kaolinite clay mainly report CH4 and CO2 equilibrium adsorption data. The physical interaction of N2 with kaolinite clay receives less attention, and the available equilibrium data for N2 adsorption on kaolinite clay remains quite scattered. Therefore, it is needed to thoroughly study the interaction of N2 and kaolinite clay and systematically compare the sorption characteristics of CO2, N2 and CH4 on kaolinite clay.

The better understanding of fluid adsorption on clay minerals must be relied on thermodynamics. Many thermodynamics variables can provide valuable physical sense to explain the interaction of adsorbent-adsorbate systems (Du et al., 2020b). The investigation about the sorption characteristics of CO2, N2 and CH4 on kaolinite clay also needs to take entropy loss (ΔS), surface potential (ΔΩ), enthalpy change (ΔH), change of Gibbs free energy (ΔG) and isosteric heat of adsorption (Qst) into consideration. Hitherto, the published literatures are concentrated on Qst of CH4 on kaolinite clay, neglecting the equally important parameters of ΔS, ΔΩ and ΔG (Xiong et al., 2017; Zhang et al., 2018). Meanwhile, the studies about CO2 and N2 adsorption on kaolinite clay from the point of view of thermodynamics are very lacking. Therefore, to better understand the adsorption mechanism of three gases on kaolinite clay, it is necessary to perform the thermodynamics analysis of CO2, N2 and CH4 adsorption.

The purpose of the current work is to compare the adsorption behaviors and investigate the competitive adsorption mechanism of CO2, N2 and CH4 on kaolinite clay. To address this issue, we report the adsorption measurements conducted on kaolinite clay with CO2, N2 and CH4 at 288–328 K. The thermodynamics variables of Qst, ΔΩ, ΔG and ΔS of three gases were discussed. We hope that the findings can serve as a foundation for the research of shale gas exploration and CO2 geological storage in kaolinite-rich shale formations.

2

2 Materials

The kaolinite clay was sourced from Shanghai Macklin Biochemical Technology Co., Ltd (Shanghai, China). The three gases adopted in this research were acquired from Tiange Gas Co., Ltd (Chongqing, China) at the purities of 99.999% for CH4, 99.999% for CO2 and 99.99% for N2.

3

3 Experiments

3.1

3.1 Microscopic structural characterization

The pysisorption of low-pressure CO2 and N2 using an ASAP 2460 analyzer (Micromeritics, Georgia, USA) was performed to obtain the microscopic structure of kaolinite clay. Due to the pore shrinkage effect at ultra-low temperature of 77 K, N2 molecule cannot enter into the ultrafine pore space (Wang et al., 2015). On the contrary, CO2 molecule has the ability to diffuse into smaller micropore. Thereby, CO2 molecule can more accurately probe the micropore information, and the micropore morphology of kaolinite clay was determined by low-pressure CO2 adsorption. The low-pressure N2 adsorption was adopted to provide the mesopore and macropore properties of kaolinite clay. According to the obtained adsorption data, the pore size distribution, pore volume and specific surface area can be determined.

3.2

3.2 Adsorption isotherm measurement

The adsorption isotherm measurements for CO2, N2 and CH4 were conducted using a high-resolution gravimetric analyzer (IGA-100B, U.K.). Unlike the traditional volumetric method, gravimetric method is a direct experiment method to measure the adsorption quantity. As demonstrated by Liu et al. (2019), for the isothermal adsorption measurement conducted on shale, gravimetric system has a higher accuracy than volumetric system. Hence, gravimetric method was chosen to conduct the adsorption experiments of CO2, N2 and CH4 on kaolinite clay.

In addition, due to the cation-exchange property of clay sheets and the full/partial changes of clay atoms, clay minerals always have the strong hydrophilicity (Wang and Huang, 2019). Accordingly, in order to remove the effect of moisture on the sorption performances of CO2, N2 and CH4 on kaolinite clay, the experiment system was degassed at 378 K for 10 h before the sorption measurement. Then, the experiment pressure was enhanced to the set value and the adsorption quantity versus time was automatically recorded. This experiment process was repeated until the maximum experiment pressure was reached.

4

4 Modeling

4.1

4.1 Adsorption model

The isothermal sorption curve can not only give insight into the adsorptive potential for an adsorbent but also provide the information about the surface properties of adsorbent and the interaction between adsorbate-adsorbent systems (Du et al., 2021b; Hameed et al., 2017; Kamboh et al., 2014; Song et al., 2017). At present, many adsorption models, such as Langmuir model, Brunauer-Emmett-Teller model and pore-filling model, have been proposed to model the adsorption isotherms (Tang et al., 2017). Among these models, Langmuir model is the most commonly used model based on the simplicity and the reasonable explanation to the model variables. Langmuir model assumes that the adsorption is occurring on the chemically, energetically and structurally homogeneous surface by monolayer sorption (Zhou et al., 2012). The isothermal sorption curves of CO2, N2 and CH4 were matched by Langmuir model. Eq. (1) gives the form of Langmuir model.

(1)
n = n m b P 1 + b P where nm is the maximum monolayer adsorption amount; b is the Langmuir constant; n is the uptake; and P is pressure.

The fitness of Langmuir equation to experiment results can be analyzed by correlation coefficient (R2) and average relative error (ARE) (Song et al., 2017). The average relative error was calculated by Eq. (2).

(2)
ARE ( % ) = 100 m i = 1 m n exp - n mol n exp i where nmol is the simulated data; and nexp is the testing data.

4.2

4.2 Separation coefficient

The separation coefficient can be used as an indicator to estimate the competitive adsorption of fluids on adsorbent (Merkel et al., 2015). A greater separation coefficient reveals that the difference in interaction strengths of fluids is more remarkable and the pre-adsorbed component with weaker adsorption can be easily and effectively replaced by the component with stronger adsorption. The separation coefficient is expressed by Eq. (3) (Song et al., 2017).

(3)
α i / j = n m , i b i n m , j b j where bi and bj are the Langmuir constants of component i and component j, respectively; and nm,i and nm,j are the maximum sorption capacities of component i and component j, respectively.

4.3

4.3 Henry’s coefficient

When the experiment pressure is small, the adsorption quantity increases linearly with pressure, which is known as Henry’s law (Tang et al., 2015). Over the Henry’s region, each adsorbate molecule can be independently in contact with the whole material surface. The interaction force between adsorbate and material surface is dominated at the low pressure condition (Du and Wang, 2021). Consequently, Henry’s coefficient (KH) can be identified as a useful index to estimate the adsorption affinity. The larger KH corresponds to the stronger adsorption affinity.

At the low-pressure region, the relation of uptake and pressure can be described by virial equation, as shown in Eq. (4) (Du et al., 2020a; Ning et al., 2012).

(4)
l n ( ( P / M P a ) / ( q / m m o l / g ) ) = K 0 + K 1 n + K 2 n 2 + K 3 n 3 + · · ·

where K0, K1, K2 and K3 are the first, second, third and fourth viral coefficients, respectively; K0 is related to Henry’s coefficient, and KH = exp(−K0).

Considering the smaller adsorption amount, the higher-order term in Eq. (4) can be removed, and Eq. (5) can be gained.

(5)
l n ( ( P / M P a ) / ( q / m m o l / g ) ) = K 0 + K 1 n

Through matching the linear part of ln(P/q) as a function of n, K0 can be acquired, and then KH can be calculated.

4.4

4.4 Adsorption thermodynamics

The main thermodynamics parameters discussed in this work are Qst, ΔΩ, ΔG and ΔS. Qst is important in assessing the heat evolved/consumed during the adsorption/desorption processes (Sircar, 1992). Qst at a given uptake (qi) can be calculated using Clausius-Clapeyron equation, as expressed in Eq. (6) and Eq. (7) (Xue et al., 2018).

(6)
Q st = R T 2 ln ( P / M P a ) T n i
(7)
ln ( P / M P a ) = - Q st RT + C
where R is the universe gas coefficient; T is temperature; and C is the integration constant.

ΔΩ is adopted to estimate the required work to reach adsorption equilibrium (Ridha and Webley, 2010). ΔG is utilized to determine the spontaneity of sorption process. Ω and ΔG can be calculated by Eq. (8) and Eq. (9), respectively (Du et al., 2020b).

(8)
Δ Ω = - R T 0 P n P d P
(9)
Δ G = Δ Ω n = - RT 0 P n P d P n

ΔS is useful to understand the variation in the disorderliness of adsorbate-adsorbent systems, which can be gained by Eq. (10) and Eq. (11) (Song et al., 2017).

(10)
Δ S R = Δ h RT - μ RT - Δ Ω nRT
(11)
Δ h = 1 n 0 n Q st d n
where Δh is the molar enthalpy of adsorption; and μ is the chemical potential of the adsorbed gas.

5

5 Results and discussion

5.1

5.1 Pore structure of kaolinite clay

5.1.1

5.1.1 Mesopore and macropore morphology of kaolinite clay

The N2 adsorption and desorption isothermal curves at 77 K are plotted in Fig. 1(a), and the pore size distribution (PSD) analyzed by N2 adsorption is described in Fig. 1(b). It is clear that when the relative pressure is high, adsorption isotherm and desorption isotherm do not coincide. The hysteresis loop given in Fig. 1(a) belongs to type H3, illustrating that kaolinite clay contains a large amount of mesopores and these mesopores are slit-shaped (Hwang and Pini, 2019). The slit-shaped mesopores are mainly formed by the particles stacking of platy mineral. In addition, as plotted in Fig. 1(b), kaolinite clay has a continuous PSD ranging from 1.5 nm to 38 nm.

Low-pressure N2 adsorption and desorption isotherms (a) and pore size distribution (b).
Fig. 1
Low-pressure N2 adsorption and desorption isotherms (a) and pore size distribution (b).

Table 1 reports the pore structure parameters of kaolinite clay and shale obtained from N2 adsorption. Clearly, kaolinite clay does not have macropore, and the proportion of mesopore volume to total pore volume (Vmes/Vt) is up to 90.00%. Accordingly, the BET specific surface area (SBET) and Vt of kaolinite clay are mainly contributed by mesopore. SBET and Vt for shale are both higher than those for kaolinite clay. The higher SBET and Vt are beneficial to the adsorption of fluid on shale. Meanwhile, unlike kaolinite clay, shale also has a certain amount of macropores. These macropores can offer more flow paths and facilitate the fluid movement. Therefore, fluid flow on actual shale reservoirs is easier than that on kaolinite clay.

Table 1 Pore structure of kaolinite clay and shale obtained from N2 adsorption measurement.
Material Vt (cm3/g) Vmac (cm3/g) Vmes (cm3/g) Vmic (cm3/g) Vmac/Vt (%) Vmes/Vt (%) Vmic/Vt (%) SBET (m2/g) Data
Kaolinite clay 0.0160 0 0.0144 0.0016 0 90.00 10.00 9.11 This study
Nanchuan shale 0.0349 0.0038 0.0302 0.0009 10.89 86.53 2.58 35.41 Gu et al., 2017
Longmaxi shale 0.0239 0.0034 0.0169 0.0036 14.23 70.71 15.06 26.87 Gu et al., 2017

Note: Vmac is the macropore volume; Vmes is the mesopore volume; Vmic is the micropore volume; and Vt is the total pore volume.

5.1.2

5.1.2 Micropore morphology of kaolinite clay

The measured adsorption isotherm of CO2 at 273.15 K is described in Fig. 2(a), and the PSD of micropore of kaolinite clay is presented in Fig. 2(b). As given in Fig. 2(b), the PSD of micropore of kaolinite clay is not continuous. The micropore of kaolinite clay is concentrated at 0.68 nm, 0.75 nm and 0.81 nm.

Adsorption isotherm of CO2 (a) and micropore size distribution of kaolinite clay (b).
Fig. 2
Adsorption isotherm of CO2 (a) and micropore size distribution of kaolinite clay (b).

The comparison of micropore structure of kaolinite clay and shale is recorded in Table 2. The specific surface area (SCO2) and volume (VCO2) of micropore of kaolinite clay obtained by the adsorption of CO2 are 1.85 m2/g and 0.00077 cm3/g, respectively. SCO2 and VCO2 for kaolinite clay are significantly lower than SBET and Vt for kaolinite clay, respectively. Mesopore of kaolinite clay are more developed than micropore, and the studied kaolinite clay is mainly mesoporous material. As reported in Table 2, SCO2 and VCO2 for shale are both higher than those for kaolinite clay, indicating that shale has more developed micropore structure than kaolinite clay. With pore size decreases, the interaction energies on multiple surfaces superpose, leading to the noticeable improvement of attraction (Heller and Zoback, 2014). Consequently, micropore is more favorable for gas adsorption, and shale with well-developed micropore structure has greater adsorption capacity than kaolinite clay. In addition to clay minerals, shale also contains a certain amount of organic matter. The organic matter is responsible for the more developed micropore structure of shale.

Table 2 Micropore characterization results of kaolinite clay and shale.
Material VCO2 (cm3/g) SCO2 (m2/g) Data
Kaolinite clay 0.00077 1.85 This study
Jiaoshiba shale 0.00370 9.32 Huo et al., 2017
Longmaxi shale 0.00964 9.20 Cheng et al., 2020

5.2

5.2 Adsorption behaviors of CH4, N2 and CO2 in kaolinite clay

The obtained isothermal sorption curves of CO2, N2 and CH4 on kaolinite clay are displayed in Fig. 3(a), 3(b) and 3(c), respectively. Obviously, as the equilibrium pressure enhances, the uptakes for three gases all improve, suggesting that high pressure promotes CO2 storage and the accumulation of CH4-in-place. Raising equilibrium pressure can give adsorbate molecules greater opportunity to collide with matrix surface, resulting in more adsorbate molecules to be retained in the adsorbed phase. Meanwhile, Jin and Fioozabadi (2014) found that CH4 and CO2 can also be trapped in the middle of bigger pores with increasing pressure when the micropores are occupied by previously injected CH4 and CO2. As a result, high pressure can increase the uptakes of three gases on kaolinite clay.

Isothermal adsorption curves at different temperatures: (a) CO2, (b) N2 and (c) CH4.
Fig. 3
Isothermal adsorption curves at different temperatures: (a) CO2, (b) N2 and (c) CH4.

From Fig. 3, it can be found that the uptakes of three gases all decrease monotonically with temperature. A higher experiment temperature indicates the larger average kinetic energy of adsorbate molecules (Tang et al., 2015). With improving temperature, the injected adsorbate molecules have the adequate energy to overcome the electrostatic force and van der Walls force and stay in the gaseous phase (Zhou et al., 2012). The negative relation of adsorption quantity and temperature suggests that CO2, N2 and CH4 adsorption on kaolinite clay is an exothermic process.

By comparing the sorption performances of CO2, N2 and CH4 on kaolinite clay, we can see that the shape of adsorption isotherm of N2 is different from those of CO2 and CH4. N2 adsorption isotherm is linear, while the isotherms of CO2 and CH4 belong to Type I isotherm according to International Union of Pure and Applied Chemistry (IUPAC) classification. The shape of isothermal sorption curves of CO2 and CH4 on studied kaolinite clay is similar to the results of Heller and Zoback (2014) about the sorption of CH4 and CO2 on pure kaolinite sample. The difference in the isotherm shapes for three gases denotes that the interaction of N2-kaolinite is distinct from those of CO2-kaolinite and CH4-kaolinite. When equilibrium pressure is small, the isothermal sorption curves of CO2 and CH4 are steeper, and their sorption rates are faster. As the pressure enlarges, the sorption rates of CO2 and CH4 gradually decline. At the initial stage, the available micropores for CO2 and CH4 are very plentiful. Meanwhile, previous study has pointed out the sorption potential will become deeper with the pore size decreases (Broom and Thomas, 2013). Hence, the abundant micropores with deeper sorption potential give rise to the quicker sorption of CO2 and CH4 at the early phase. However, for N2 adsorption on kaolinite clay, the sorption rate almost keeps constant over the whole pressure range. The linear relation of the uptake of N2 with pressure manifests that N2 adsorption quantity is smaller and the interaction of N2-kaolinte is weaker. This may be due to the pore size occupied by adsorbed N2 is bigger than that occupied by adsorbed CO2 and CH4. The bigger pore has the shallower adsorption potential, which leads to the slow and constant sorption rate of N2.

As depicted in Fig. 3, the order of the uptakes of three gases on kaolinite clay is as follows: CO2 > CH4 > N2, agreeing with the findings of Xiong et al. (2017). When researching the sorption of three gases on New Albany shale, Chareonsuppanimit et al. (2012) found that the ratio of adsorption capacities for N2, CH4 and CO2 is 1:3.2:9.3. The ratios of CO2 adsorption amount (qCO2) to N2 adsorption amount (qN2) and qCO2 to CH4 adsorption amount (qCH4) on studied kaolinite clay are both lower than those on New Albany shale. The greater value of qCO2 than qCH4 on kaolinite clay supports the technical feasibility of applying CO2-ESGR project in kaolinite-rich shale reservoirs.

The difference in adsorption behaviors of CO2, N2 and CH4 on porous material is mainly due to the difference in physical characteristics of three gases, e.g., shape, polarity, kinetic diameter, dipole moment and quadrupole moment (Charriere et al., 2010). Table 3 summarizes the physical characteristics of CO2, N2 and CH4. The linear shape and smaller kinetic diameter promote CO2 to enter into more confined micropore, while these ultramicropores are only partly accessible to CH4 owing to the molecule sieve effect. The research results of Liu et al. (2013) show that the available pore space for CH4 adsorption within the organic matter is 40% smaller than that for the adsorption of CO2. Meanwhile, the boiling point of CH4 is lower than that of CO2, which means that adsorbed CH4 is easier to escape from the matrix by “evaporation” (Ettinger et al., 1966). Furthermore, CO2 molecule possesses a permanent quadrupole moment, while CH4 molecule is charge neutral. The quadrupole moment enhances the interaction of CO2-kaolinite, bringing about the stronger adsorbed layer for CO2 as confirmed by Jin and Firoozabadi (2014). Thereby, the studied kaolinite clay preferentially adsorbs CO2 over CH4. The smaller adsorption quantity of N2 than CH4 on kaolinite clay is mainly attributed to the lower polarizability of N2 molecule, which leads to the weaker interaction between N2 and kaolinite clay.

Table 3 Physical characteristics of CO2, N2 and CH4.
Gas Shape Boiling point (K) Kinetic diameter (nm) Polarizability(10−24 cm3) Dipole moment (10−18 esu cm) Quadrupole moment (10−26 esu cm2)
CO2 linear 194.65 0.330 2.65 0 4.30
CH4 tetrahedral 111.65 0.380 2.60 0 0
N2 linear 77.30 0.364 1.76 0 1.52

Fig. 3 presents the matching results of the sorption isotherms using Langmuir model, and the corresponding fitting parameters in Langmuir model are detailed in Table 4. It is evident that the simulated data are consistent with the experimental results. The consistency between experimental data and modeling results over a broad pressure and temperature range implies the excellent applicability of Langmuir model to simulate three gases adsorption on kaolinite clay.

Table 4 Langmuir fitting parameters of three gases on kaolinite clay.
Gas T (K) b (MPa−1) nm (mmol/g) ARE (%) R2
288 1.7070 0.1479 5.7993 0.9956
CO2 308 1.3320 0.1310 5.7394 0.9966
328 0.9221 0.1298 6.7149 0.9980
288 2.5078 0.0760 6.1598 0.9918
CH4 308 1.7616 0.0753 4.8197 0.9967
328 1.2197 0.0752 3.6734 0.9978
288 0.1437 0.2031 5.4168 0.9987
N2 308 0.1234 0.1995 5.4650 0.9963
328 0.1056 0.1975 9.9870 0.9918

5.3

5.3 Separation coefficients of CH4, N2 and CO2 in kaolinite clay

The calculated separation coefficients of CO2 over CH4 (αCO2/CH4), CO2 over N2 (αCO2/N2) and CH4 over N2 (αCH4/N2) are recorded in Table 5. The values of αCO2/CH4 are all bigger than one, demonstrating that the adsorbed CH4 can be directly displaced by injected CO2 on kaolinite clay. As listed in Table 5, the order of separation coefficients on kaolinite clay is as follows: αCO2/N2 > αCH4/N2 > αCO2/CH4. As the outstanding adsorption materials, clay minerals have been widely used for gas separation (Broom and Thomas, 2013). For CH4 enrichment from extracted coalbed gas, whose main components are CH4 and N2, addition of kaolinite clay to modify the adsorption materials can be considered. The values of αCO2/CH4, αCO2/N2 and αCH4/N2 all exhibit a decreasing trend with temperature. Low temperature not only increases the uptake but also improves the separation efficiency. The negative relationship of αCO2/CH4 with temperature has also been found on other adsorbents (Liu et al., 2018). Liu et al. (2018) thought that the induced desorption of adsorbate from matrix surface is the main reason for the decrease in αCO2/CH4 at high temperature. Meanwhile, in the previous study about CO2 and CH4 competitive adsorption on kerogen, αCO2/CH4 also continuously declines with pressure (Anh et al., 2018). Therefore, the shallow shale reservoirs with lower temperature and pressure condition boost the replacement of adsorbed CH4 by CO2, agreeing with the findings of Wang et al (2018).

Table 5 Separation coefficients on kaolinite clay and shale.
Material P (MPa) T (K) αCO2/CH4 αCO2/N2 αCH4/N2 Data
0–1.80 288 1.67 8.65 6.53 This study
Kaolinite clay 0–1.80 308 1.65 7.08 5.39
0–1.80 328 1.61 5.74 4.40
Wufeng shale 0–1.80 278–318 4.98–7.04 Gu et al., 2017
Longmaxi shale 0–1.80 278–318 2.87–4.33
Nanchuan shale 0–1.80 278–318 3.82–6.63
Fuling shale 0–1.80 278–318 3.29–4.77
Marcellus shale 0–27.50 313 4.76 Pei et al., 2015
Eagle Ford shale 0–41.30 313 4.42

The values of αCO2/CH4 on shale are also shown in Table 5. It can be found from Table 5 that αCO2/CH4 values on shale are bigger than those on kaolinite clay, elucidating that the actual shale formation is more suitable for the extraction of CH4 by CO2 injection. The findings of Wang and Huang (2019) pointed out αCO2/CH4 reduces as the pore size enlarges. As discussed in Section 5.1, shale has a more developed micropore structure. Accordingly, even though increasing kaolinite clay content can improve the adsorption ability of shale, it damages the displacement efficiency of CO2 on shale. Hence, to select the target reservoir for implementation of CO2-ESGR project, the content of kaolinite clay in shale sediment needs to be paid more attention.

5.4

5.4 Adsorption affinities of CH4, N2 and CO2 in kaolinite clay

The gained KH of CH4, N2 and CO2 in kaolinite clay are listed in Table 6. CO2 has a largest KH, while the KH value of N2 is the smallest. The adsorption affinity of CO2 on kaolinite clay is biggest, and N2 has a lowest affinity. For the adsorbate-adsorbent pair, the interaction strength depends on dipole moment-field gradient interaction energy, quadrupole moment-field gradient and induction energy (Deng et al., 2012). As given in Table 3, the highest quadrupole moment and polarizability make the strongest interaction of CO2-kaolinite and the largest KH for CO2. Although N2 molecule has a quadrapole moment, CH4 polarizability is higher. The combined effect results in the larger KH for CH4 than N2. In addition, temperature has a negative impact on KH for three gases, revealing that increasing temperature can decrease the adsorption affinity.

Table 6 Henry’s coefficients on kaolinite clay and shale.
Material Gas T (K) KH Gas T (K) KH Gas T (K) KH Data
Kaolinite clay 288 0.39 288 0.19 288 0.05 This study
CO2 308 0.26 CH4 308 0.18 N2 308 0.03
328 0.16 328 0.12 328 0.01
Youyang shale 278 1.13 278 0.12 Duan, 2017
CO2 298 0.65 CH4 298 0.11
318 0.44 318 0.07
Weiyuan shale 278 0.86 278 0.11 Duan, 2017
CO2 298 0.54 CH4 298 0.09
318 0.32 318 0.07

Note: the unit of KH is mmol/g/MPa.

The comparison of KH for CH4 and CO2 on kaolinite clay and shale is elucidated in Table 6. The difference in KH for CH4 and CO2 on Youyang shale and Weiyuan shale is more apparent than that on kaolinite clay. Because KH has a positive correlation with adsorption quantity, the difference in the uptakes of CH4 and CO2 on kaolinite clay is smaller. This smaller difference in the uptakes of CH4 and CO2 on kaolinite clay is not conducive to the enlargement of αCO2/CH4. Therefore, αCO2/CH4 on shale is greater than that on kaolinite clay.

From Table 6, we can find that KH values of CO2 on two shale samples are both higher than that on kaolinite clay, which reflects that the adsorption affinity and adsorption capacity of CO2 are larger on shale. For example, the uptake of CO2 on Youyang shale is 0.17 mmol/g under 298 K and 1.80 MPa, while the uptake of CO2 on kaolinite clay under a lower temperature of 288 K and 1.80 MPa is only 0.10 mmol/g. In addition, KH values of CH4 on kaolinite clay are slightly bigger than those on shale. Therefore, the uptakes of CH4 on kaolinite clay and shale are almost equivalent.

5.5

5.5 Thermodynamics analysis of CH4, N2 and CO2 in kaolinite clay

5.5.1

5.5.1 Surface potentials of CH4, N2 and CO2 in kaolinite clay

Surface potential ΔΩ corresponds to the needed minimum work to load adsorption material to reach a specific level (Mofarahi and Bakhtyari, 2015). Fig. 4 presents the gained ΔΩ of three gases in kaolinite clay. The ΔΩ values of three gases under experimental temperature and pressure are all negative. The absolute values of ΔΩ versus equilibrium pressure show a rapid increase upon further filling of adsorbate molecules into pore space. Under higher pressure (loading) condition, the more isothermal work is needed to load additional adsorbate molecules into the pore structure than at the preliminary stage (Ridha and Webley, 2010). The enhancement of isothermal work for adsorbate molecules with pressure is owing to the screened active sorption sites of porous material. As depicted in Fig. 4, ΔΩ for three gases are close to zero when the equilibrium pressure approaches zero. On the basis of the definition of ΔΩ proposed by Van Ness, the specific chemical potential of porous media in contact with molecules is becoming equal to the specific potential of porous media without contact with molecules at lower loading condition (Van, 1969). Thereby, reducing pressure causes ΔΩ to approach zero.

Surface potentials at different temperatures: (a) CO2, (b) N2 and (c) CH4
Fig. 4
Surface potentials at different temperatures: (a) CO2, (b) N2 and (c) CH4

It is clear that the absolute values of ΔΩ for three gases display a noticeable reduction upon the temperature improvement from 288 K to 328 K. Previous study has shown that the less adsorption quantity brings about the lower (less negative) surface potential (Du et al., 2020b). Thus, adsorption amount and surface potential both reduce as the temperature increases. This is because the less adsorbate molecules are required to fill into cavity pore under higher temperature.

At the same condition, CO2 has the biggest absolute value of ΔΩ, while N2 possesses the smallest absolute value of ΔΩ. The smallest uptake of N2 on kaolinite clay results in the least isothermal work and the lowest ΔΩ. The increase rates of ΔΩ for CO2 and CH4 are larger at the early stage, and then they reduce gradually with pressure. However, the increase rate of ΔΩ for N2 keeps the same over the whole pressure range. The variation trends of ΔΩ with pressure are accordant with the change trends of the uptakes of three gases with pressure.

5.5.2

5.5.2 Changes of Gibbs free energy of CH4, N2 and CO2 in kaolinite clay

When temperature and pressure keep constant, the increment of free energy per unit surface area is called Gibbs free energy (Zhou et al., 2011). Gibbs free energy is caused by the asymmetric force field of atoms (molecules or ions) on the material surface (Nie et al., 2000). When the surface of porous material is formed, the force acting on the surface atoms is unbalanced, and the surface atoms are affected by the attraction which points vertically to the interior of solid. Therefore, the surface atoms tend to move towards the interior of solid, which makes the surface atoms to obtain an additional energy, namely, Gibbs free energy. Accordingly, the change of Gibbs free energy ΔG is a critical parameter to measure the spontaneity of adsorption system. Fig. 5 describes ΔG of three gases in kaolinite clay. ΔG values of three gases are all negative, which indicates that gas adsorption in kaolinite clay is spontaneous process.

Changes of Gibbs free energy at different temperatures: (a) CO2, (b) N2 and (c) CH4.
Fig. 5
Changes of Gibbs free energy at different temperatures: (a) CO2, (b) N2 and (c) CH4.

As the pressure rises, the absolute value of ΔG enhances, showing that increasing pressure improves the spontaneity of sorption. As a result, there is a monotonous enhancement of the uptake with pressure. Injection of N2 into shale reservoir can cause CH4 desorption by reducing CH4 partial pressure and lowering CH4 adsorption spontaneity. The technology of N2-ESGR is also feasible to drive CH4 exploitation. Moreover, the influence of pressure on ΔG of CH4 and CO2 is more evident than that of N2. For N2 adsorption on kaolinite clay, improving temperature decreases the spontaneity of adsorption. For the sorption of CH4 and CO2, the relation of ΔG with temperature is complicated. At the pressure larger than 0.10 MPa, low temperature helps to gain higher ΔG of CO2. When pressure is higher than 0.50 MPa, low temperature corresponds to the higher ΔG of CH4. Therefore, only under the condition of high pressure, low temperature facilitates the spontaneity of adsorption of CO2 and CH4 in kaolinite clay.

As given in Fig. 5, the absolute value of ΔG of CO2 is the biggest, while N2 has the smallest absolute value of ΔG. Consequently, CO2 adsorption on kaolinite clay has the highest spontaneity, and the spontaneity degree of N2 adsorption on kaolinite clay is the lowest. According to the minimum energy principle, the lower the system energy is, the more stable the system is (Zhou et al., 2011). The matrix surface always tries to adsorb other substances around to reduce the surface free energy. The larger change of surface free energy gives the adsorption material a greater motivation to reduce energy by adsorbing gas. Therefore, CO2 with biggest ΔG has the largest uptake on kaolinite clay, while the adsorption of N2 on kaolinite clay is the most difficult.

The change of surface free energy is not only related to the spontaneity of adsorption, but also to the matrix deformation. As pointed out by Nie et al. (2000), the larger change of surface free energy can lead to the greater deformation and the easier fracture of coal body. For the selection of site for CO2 geological sequestration, one of most significant factors is the long-term stability of CO2 storage (Huang et al., 2018). To increase the safety of CO2 sequestration into deep subsurface and reduce the risk in CO2 leakage into surrounding strata, more efforts should be focused on the deformation and the change of strength of shale after CO2 adsorption.

5.5.3

5.5.3 Isosteric heats of adsorption of CH4, N2 and CO2 in kaolinite clay

Isosteric heat of adsorption Qst indicates adsorption heat effect and is one of the critical variables for the design of adsorptive separation process. Fig. 6 depicts the acquired Qst of three gases versus loading in kaolinite clay. The enthalpy change ΔH (-Qst) values for three gases are negative, which confirms that the adsorption of three gases in kaolinite clay is exothermic process. All of Qst values are smaller than 40 kJ/mol, validating that the adsorption of three gases in kaolinite clay is physisorption.

Isosteric heats of adsorption for three gases in kaolinite clay.
Fig. 6
Isosteric heats of adsorption for three gases in kaolinite clay.

With the loading increases, Qst for three gases all show a monotonous downward trend. If the adsorption material surface is energetically homogeneous, Qst is not relevant to the loading, while Qst changes with loading when the adsorption material surface is energetically heterogeneous (Chakraborty et al., 2006; Sircar, 1991). The decrease in Qst with loading proves the heterogeneous surface of kaolinite clay. During the adsorption process, Qst value is affected by the interaction energy of adsorbate-adsorbate and adsorbate-adsorbent pairs. At the preliminary stage of adsorption, lots of active adsorption sites are available, and the adsorbates are directly adsorbed on the preferential adsorption sites, causing the stronger force between adsorbate and adsorbent and the weaker lateral interaction between adsorbate molecules. As the loading increases, more and more pore space is occupied, and the interaction energy of adsorbate-adsorbent is lower than adsorbate-adsorbate interaction energy. Thus, Qst declines with improving loading.

Clearly, CO2 has the largest Qst, and Qst value of N2 is the smallest. The biggest Qst indicates that the interaction of CO2-kaolinite is the strongest. Accordingly, N2 with the lowest Qst has the weakest interaction with kaolinite. The quadrupole moment strengthens the interaction of CO2 with kaolinite, and hence CO2 adsorption gives off more heat. Meanwhile, Qst enhances constantly until the diameter ratio of the pore to the molecular approaches one, and increasing pore curvature has a positive effect on Qst (Keffer et al., 1996). Because CO2 can migrate into smaller pores, more heat is released for CO2 adsorption. The reduction rate of Qst with loading for N2 is the fastest. The rapid decrease in Qst illustrates the weaker intermolecular force between N2 molecules at the late stage of adsorption.

5.5.4

5.5.4 Entropy losses of CH4, N2 and CO2 in kaolinite clay

Entropy loss ΔS is defined as the change from the entropy of a clean porous material and adsorbates in the perfect-gas state to that of the adsorption system containing a porous material with adsorbed molecules inside its pore (Myers, 2004). Information about ΔS during the adsorption process is conducive to comprehending the packing patter and the confined mobility of adsorbed adsorbates.

The results of ΔS of three gases are detailed in Fig. 7. From Fig. 7, it can be seen that ΔS values are all negative, indicating that gas adsorption in kaolinite clay is from an unrestricted stage to a restricted stage. As reported by Zhou et al. (2012), adsorption always implies the formation of a more orderly rearrangement on matrix surface and the loss of the degree of freedom of molecules. Within the micropore, the value of translational entropy is bigger than that of rotational entropy, and the majority of entropy loss is contributed by the loss of translational degree during sorption process (Myers, 2004). Therefore, the translation of molecules is more limited in the micropore.

Entropy losses for three gases at different temperatures.
Fig. 7
Entropy losses for three gases at different temperatures.

As the loading increases, the absolute values of ΔS for three gases decline. Larger adsorption loading causes the lower (less negative) ΔS. When the available pore is gradually occupied by adsorbate molecules, the disorderliness degree of investigated adsorption system increases. Therefore, higher adsorption loading condition is unfavorable for the efficient packing. At the early phase, molecules are adsorbed in the smaller pores, and the adsorbed molecules are more confined. With the increase in loading, the majority of injected molecules are retained in the bigger pores. In these large pores, the movement of adsorbed molecule is not effectively limited, bringing about the lower ΔS. In addition, under different temperatures, the curves of ΔS basically coincide. Therefore, temperature has a negligible influence on ΔS, which is in accordance with the report of Duan et al. (2016) about CH4 and CO2 adsorption on shale.

It should be noted that ΔS of N2 is the lowest, while CO2 has the highest ΔS. Myers (2004) discovered that the adsorbent with smaller pores has bigger absolute values of entropy change. Due to the more restricted pore for CO2 adsorption, the movement of adsorbed CO2 is more difficult. Meanwhile, the higher quadrupole moment and polarizability enhance the interaction of CO2-kaolinite, giving rise to the more efficient adsorbate packing by producing the structured fluid. Therefore, ΔS values of CO2 are higher than those of N2 and CH4. The highest ΔS reveals that adsorbed CO2 molecules form a most ordered rearrangement. The smaller ΔS for N2 than CH4 is because of the lower polarizability of N2, which induces the weaker interaction between electrostatic-field gradients and N2 molecules. Furthermore, the decrease of ΔS for N2 with loading is more rapid than those for CO2 and CH4. The quick reduction of ΔS for N2 elucidates the weaker interaction strength of N2-kaolinite and the lower orderliness of adsorbed N2 on kaolinite surface.

6

6 Conclusions

In this paper, the adsorption experiments of CO2, N2 and CH4 at 288 K, 308 K and 328 K on kaolinite clay were conducted. The thermodynamics analysis of CO2, N2 and CH4 adsorption was performed to explain the adsorption behavior and understand the adsorption mechanism. The main conclusions are as following:

  • The values of αCO2/CH4 are greater than one, confirming the feasibility of replacement of adsorbed CH4 by CO2 on kaolinite clay. The values of αCO2/CH4, αCO2/N2 and αCH4/N2 all exhibit a decreasing trend with temperature. The shallow gas-bearing reservoir is beneficial for the enhanced CH4 recovery by CO2 injection.

  • The adsorption affinity for three gases on kaolinite clay is as follows: CO2 > CH4 > N2. CO2 has a strongest interaction with kaolinite clay, while the interaction of N2-kaolinite is the lowest.

  • The spontaneity degree of CO2 adsorption on kaolinite clay is the highest, while N2 has the lowest adsorption spontaneity on kaolinite clay. Injecting N2 into shale reservoir can cause CH4 desorption by lowering the spontaneity of CH4 adsorption. Adsorbed CO2 molecules form a most ordered rearrangement on kaolinite surface.

7

7 Data availability

The data applied to support the results in this study are available from the corresponding author upon request.

CRediT authorship contribution statement

Xidong Du: Writing – original draft. Dongdong Pang: Conceptualization, Supervision. Yuan Zhao: Data curation, Resources. Zhenkun Hou: Formal analysis, Project administration. Hanglong Wang: Investigation. Yugang Cheng: Methodology.

Acknowledgments

This study was supported by Yunnan Fundamental Research Projects (Grant No. 202101BE070001-039), Yunnan Department of Education Science Research Fund Project (Grant No. 2022J0055), National Natural Science Foundation of China (Grant No. 51904049), China Postdoctoral Science Foundation (Grant No. 2021M693750), and Open Research Fund of Key Laboratory of Safety and High-efficiency Coal Mining, Ministry of Education, Anhui University of Science and Technology (Grant No. JYBSYS2020207).

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. , , , , . Differential retention and release of CO2 and CH4 in kerogen nanopores: implications for gas extraction and carbon sequestration. Fuel. 2018;220:1-7.
    [Google Scholar]
  2. , , . Gas adsorption by nanoporous materials: future applications and experimental challenges. Mrs Bull.. 2013;38(5):412-421.
    [Google Scholar]
  3. , , , . On the thermodynamic modeling of the isosteric heat of adsorption and comparison with experiments. Appl. Phys. Lett.. 2006;89:171901.
    [Google Scholar]
  4. , , , , . High-pressure adsorption of gases on shales: measurements and modeling. Int. J. Coal Geol.. 2012;95:34-46.
    [Google Scholar]
  5. , , , . Effect of pressure and temperature on diffusion of CO2 and CH4 into coal from the Lorraine basin (France) Int. J. Coal Geol.. 2010;81:373-380.
    [Google Scholar]
  6. , , , , , , . Effects of supercritical CO2 treatment temperatures on mineral composition, pore structure and functional groups of shale: implications for CO2 sequestration. Sustainability. 2020;12:3927.
    [Google Scholar]
  7. , , , , , , . Adsorption equilibrium for sulfur dioxide, nitric oxide, carbon dioxide, nitrogen on 13X and 5A zeolites. Chem. Eng. J.. 2012;188:77-85.
    [Google Scholar]
  8. , , , , , , , . Study on the adsorption of CH4, CO2 and various CH4/CO2 mixture gases on shale. Alex. Eng. J.. 2020;59:5165-5178.
    [Google Scholar]
  9. , , , , , , , . CO2 and CH4 adsorption on different rank coals: A thermodynamics study of surface potential, Gibbs free energy change and entropy loss. Fuel. 2021;283:118886
    [Google Scholar]
  10. , , , , . Investigation of CO2-CH4 displacement and transport in shale for enhanced shale gas recovery and CO2 sequestration. ASME J. Energy Resour. Technol.. 2017;139:012909.
    [Google Scholar]
  11. , , , , . The influence of CO2 injection pressure on CO2 dispersion and the mechanism of CO2-CH4 displacement in shale. ASME J. Energy Resour. Technol.. 2018;140:012907.
    [Google Scholar]
  12. , , , , , . Experimental study on the kinetics of adsorption of CO2 and CH4 in gas-bearing shale reservoirs. Energ. Fuel.. 2019;33:12587-12600.
    [CrossRef] [Google Scholar]
  13. , , , , , , . Enhanced shale gas recovery by the injections of CO2, N2, and CO2/N2 mixture gases. Energ. Fuel.. 2019;33:5091-5101.
    [Google Scholar]
  14. , , , , , , , , , . Thermodynamics analysis of the adsorption of CH4 and CO2 on montmorillonite. Appl. Clay Sci.. 2020;192:105631.
    [Google Scholar]
  15. , , , , , , , , . Adsorption of CH4, N2, CO2, and their mixture on montmorillonite with implications for enhanced hydrocarbon extraction by gas injection. Appl. Clay Sci.. 2021;210:106160
    [Google Scholar]
  16. , , . Investigation into adsorption equilibrium and thermodynamics for water vapor on montmorillonite clay. AIChE J. 2021e17550
    [CrossRef] [Google Scholar]
  17. , . A thermodynamics study of CO2 and CH4 adsorption on Sichuan basin shales. Chongqing, China: Chongqing University; .
  18. , , , , . Adsorption equilibrium of CO2 and CH4 and their mixture on Sichuan basin shale. Energ. Fuel.. 2016;30:2248-2256.
    [Google Scholar]
  19. , , , , . Natural factors influencing coal sorption properties. I. Petrography and sorption properties of coals. Fuel. 1966;45:267-275.
    [Google Scholar]
  20. , , , , . Influences of the composition and pore structure of a shale on its selective adsorption of CO2 over CH4. J. Nat. Gas Sci. Eng.. 2017;46:296-306.
    [Google Scholar]
  21. , , , . Adsorption of chromotrope dye onto activated carbons obtained from the seeds of various plants: equilibrium and kinetic studies. Arab. J. Chem.. 2017;10:S2225-S2233.
    [Google Scholar]
  22. , , . Adsorption of methane and carbon dioxide on gas shale and pure mineral samples. J. Unconventional Oil Gas Resources. 2014;8:14-24.
    [Google Scholar]
  23. , , , , , , , . Effect of organic type and moisture on CO2/CH4 competitive adsorption in kerogen with implications for CO2 sequestration and enhanced shale gas recovery. Appl. Energ.. 2018;210:28-43.
    [Google Scholar]
  24. , , , , , , . CO2 geological sequestration: displacement behavior of shale gas methane by carbon dioxide injection. Int. J. Greenh. Gas Con.. 2017;66:48-59.
    [Google Scholar]
  25. , , . Supercritical CO2 and CH4 uptake by illite-smectite clay minerals. Environ. Sci. Technol.. 2019;53:11588-11596.
    [Google Scholar]
  26. , , , . Laboratory and simulation investigation of enhanced coalbed methane recovery by gas injection. Transp. Porous Med.. 2008;73:141-159.
    [Google Scholar]
  27. , , . Effect of water on methane and carbon dioxide sorption in clay minerals by Monte Carlo simulations. Fluid Phase Equilibr.. 2014;382:10-20.
    [Google Scholar]
  28. , , , , , . Adsorption of direct black-38 azo dye on p-tert-butylcalix[6]arene immobilized material. Arab. J. Chem.. 2014;7:125-131.
    [Google Scholar]
  29. , , , , , . Effect of pressure and temperature on CO2/CH4 competitive adsorption on kaolinite by Monte Carlo simulations. Materials. 2020;13:2851.
    [Google Scholar]
  30. , , , . The effect of nanopore shape on the structure and isotherms of adsorbed fluids. Adsorption. 1996;2:9-21.
    [Google Scholar]
  31. , , , , . Assessing the feasibility of CO2 storage in the New Albany shale (Devonian-Mississippian) with potential enhanced gas recovery using reservoir simulation. Int. J. Greenh. Gas Con.. 2013;17:111-126.
    [Google Scholar]
  32. , , , , , , , . Competitive adsorption and diffusion of CH4/CO2 binary mixture within shale organic nanochannels. J. Nat. Gas Sci. Eng.. 2018;53:329-336.
    [Google Scholar]
  33. , , , , . CO2/CH4 competitive adsorption in shale: implications for enhancement in gas production and reduction in carbon emissions. Environ. Sci. Technol.. 2019;53:9328-9336.
    [Google Scholar]
  34. , , , , . Competitive sorption of CH4, CO2 and H2O on natural coals of different rank. Int. J. Coal Geol.. 2015;150–151:181-192.
    [Google Scholar]
  35. , , . Experimental investigation and thermodynamic modeling of CH4/N2 adsorption on zeolite 13X. J. Chem. Eng. Data. 2015;60(3):683-696.
    [Google Scholar]
  36. , . Characterization of nanopores by standard enthalpy and entropy of adsorption of probe molecules. Colloid. Surface. A. 2004;241:9-14.
    [Google Scholar]
  37. , , , . Surface free energy of coal and its calculation. J. Taiyuan Univ. Technol.. 2000;31(4):346-348.
    [Google Scholar]
  38. , , , , , , , , . Adsorption equilibrium of methane and carbon dioxide on microwave-activated carbon. Sep. Purf. Technol.. 2012;98:321-326.
    [Google Scholar]
  39. , , , , . Shale gas reservoir treatment by a CO2-based technology. J. Nat. Gas Sci. Eng.. 2015;26:1595-1606.
    [Google Scholar]
  40. , , , , , , , , . Influence of supercritical CO2 exposure on water wettability of shale: Implications for CO2 sequestration and shale gas recovery. Energy 2021
    [CrossRef] [Google Scholar]
  41. , , , , , , . Investigation of adsorption kinetics of CH4 and CO2 on Yanchang shale exposure to supercritical CO2. Energy. 2021;236:121410
    [Google Scholar]
  42. , , . Entropic effects and isosteric heats of nitrogen and carbon dioxide adsorption on chabazite zeolites. Micropor. Mesopor. Mat.. 2010;132:22-30.
    [Google Scholar]
  43. , . Isosteric heats of multicomponent gas adsorption on heterogeneous adsorbents. Langmuir. 1991;7:3065-3069.
    [Google Scholar]
  44. , . Estimation of isosteric heats of adsorption of single gas and multicomponent gas mixtures. Ind. Eng. Chem. Res.. 1992;31:1813-1819.
    [Google Scholar]
  45. , , , , . Adsorption equilibrium and thermodynamics of CO2 and CH4 on carbon molecular sieves. Appl. Surf. Sci.. 2017;396:870-878.
    [Google Scholar]
  46. , , , . Molecular simulation of CO2/CH4 competitive adsorption in kaolinite based on adsorption potential theory. Bull. Mineral., Petrol. Geochem.. 2018;37(4):724-730.
    [Google Scholar]
  47. , , , , . Adsorption models for methane in shale: review, comparison, and application. Energ. Fuel.. 2017;31(10):10787-10801.
    [Google Scholar]
  48. , , , , , . Adsorption affinity of different types of coal: mean isosteric heat of adsorption. Energ. Fuel.. 2015;29:3609-3615.
    [Google Scholar]
  49. , . Adsorption of gases on solids. Review of role of thermodynamics. Ind. Eng. Chem. Fundamen.. 1969;8(3):464-473.
    [Google Scholar]
  50. , , . Molecular insight into competitive adsorption of methane and carbon dioxide in montmorillonite: effect of clay structure and water content. Fuel. 2019;239:32-43.
    [Google Scholar]
  51. , , , , , , , . Molecular simulation of CO2/CH4 competitive adsorption on shale kerogen for CO2 sequestration and enhanced gas recovery. J. Phys. Chem. C. 2018;122:17009-17018.
    [Google Scholar]
  52. , , , , , , , . Influence of CO2 exposure on high-pressure methane and CO2 adsorption on various rank coals: implications for CO2 sequestration in coal seams. Energ. Fuel.. 2015;29:3785-3795.
    [Google Scholar]
  53. , , , , . Adsorption behavior of methane on kaolinite. Ind. Eng. Chem. Res.. 2017;56(21):6229-6238.
    [Google Scholar]
  54. , , , , , , . Adsorption thermodynamic property of Yangchang formation shale in Ordos basin. Coal Geol. Exploration. 2018;46(4):22-27.
    [Google Scholar]
  55. , , , . Effect of water on methane adsorption on the kaolinite (001) surface based on molecular simulations. Appl. Surf. Sci.. 2018;439:792-800.
    [Google Scholar]
  56. , , , , . Strategy optimization on industrial CO2 sequestration in the depleted Wufeng-Longmaxi formation shale in the Northeastern Sichuan basin, SW China: from the perspective of environment and energy. ACS Sustain. Chem. Eng.. 2020;8(30):11435-11445.
    [Google Scholar]
  57. , , , . Thermodynamic analysis of competitive adsorption of CO2 and CH4 on coal matrix. J. China Coal Soc.. 2011;36(8):1307-1311.
    [Google Scholar]
  58. , , , , . Adsorption mechanism of CO2/CH4 in kaolinite clay: insight from molecular simulation. Energ. Fuel.. 2019;33:6542-6551.
    [Google Scholar]
  59. , , , , , . Thermodynamics for the adsorption of SO2, NO and CO2 from flue gas on activated carbon fiber. Chem. Eng. J.. 2012;200–202:399-404.
    [Google Scholar]
Show Sections