Next Article in Journal
Influence of Partial Arc on Electric Field Distribution of Insulator Strings for Electrified Railway Catenary
Previous Article in Journal
The Half-Sine Method: A New Accurate Location Method Based on Wavelet Transform for Transmission-Line Protection from Single-Ended Measurements
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Supercritical CO2 Exposure-Induced Surface Property, Pore Structure, and Adsorption Capacity Alterations in Various Rank Coals

1
College of Civil Engineering, Yancheng Institute of Technology, Yancheng 221051, China
2
State Key Laboratory of Coal Mine Disaster Dynamics and Control, Chongqing University, Chongqing 400044, China
3
Geofluids, Geomechanics and Geoenergy (3G) Research Group, Chongqing University, Chongqing 400044, China
4
China Coal Technology and Engineering Group Shenyang Research Institute, Fushun 113122, China
5
State Key Laboratory of Coal Mine Safety Technology, Fushun 113122, China
6
School of Earth Sciences, East China University of Technology, Nanchang 330013, China
*
Author to whom correspondence should be addressed.
Energies 2019, 12(17), 3294; https://doi.org/10.3390/en12173294
Submission received: 10 August 2019 / Revised: 21 August 2019 / Accepted: 25 August 2019 / Published: 27 August 2019
(This article belongs to the Section H: Geo-Energy)

Abstract

:
Carbon dioxide (CO2) has been used to replace coal seam gas for recovery enhancement and carbon sequestration. To better understand the alternations of coal seam in response to CO2 sequestration, the properties of four different coals before and after supercritical CO2 (ScCO2) exposure at 40 °C and 16 MPa were analyzed with Fourier Transform infrared spectroscopy (FTIR), low-pressure nitrogen, and CO2 adsorption methods. Further, high-pressure CO2 adsorption isotherms were performed at 40 °C using a gravimetric method. The results indicate that the density of functional groups and mineral matters on coal surface decreased after ScCO2 exposure, especially for low-rank coal. With ScCO2 exposure, only minimal changes in pore shape were observed for various rank coals. However, the micropore specific surface area (SSA) and pore volume increased while the values for mesopore decreased as determined by low-pressure N2 and CO2 adsorption. The combined effects of surface property and pore structure alterations lead to a higher CO2 adsorption capacity at lower pressures but lower CO2 adsorption capacity at higher pressures. Langmuir model fitting shows a decreasing trend in monolayer capacity after ScCO2 exposure, indicating an elimination of the adsorption sites. The results provide new insights for the long-term safety for the evaluation of CO2-enhanced coal seam gas recovery.

1. Introduction

Recently, concerns about increasing carbon dioxide (CO2) concentration in the atmosphere have driven research into the technical reduction of the emission of CO2 from fossil fuel use [1]. Carbon capture and storage (CCS) is an effective approach for CO2 mitigation currently under consideration [2,3]. Deep, un-minable coal seams have been targeted as safer sites for CO2 sequestration because the sequestrated CO2 is predominantly stored as a relatively stable adsorption phase in coal seams. Meanwhile, the replacement of coal seam gas (CSG) offsets some proportion of sequestration costs [4,5,6]. Accordingly, an understanding of the physical and chemical alterations of coal reservoirs in response to CO2 sequestration would offer a scientific foundation on which to base long-term storage predictions.
Recently, several studies have investigated the chemical and structural properties of various rank coals [7,8,9,10]. These studies have demonstrated that coal mass is an organic-rich porous rock, consisting of micropores (<2 nm in diameter) and mesopores (2–50 nm in diameter) in coal matrix and cleats (>50 nm in diameter). The structural properties of coal vary with the degree of maturity as determined by reflectance or volatile matter [11]. In the CO2-enhanced coal bed methane recovery (CO2-ECBM) process, CO2 fluids first flow into the cleat system of coal mass and then slowly diffuse into mesopores and micropores. In mesopores and micropores, CO2 can displace the pre-adsorbed CSG through the competitive adsorption mechanism, thus achieving CO2 storage and CSG enhancements [12].
The swelling effect of coal matrix induced by CO2 adsorption has been reported extensively [13,14,15,16]. As reviewed by Perera et al. [17], the swelling of the coal matrix is accompanied by a series of irreversible effects on coal seam, including the variation of coal seam permeability and mechanical properties. The matrix swelling effect is closely related to coal rank, where higher-ranked coal exhibits a lower matrix swelling rate [18]. Normally, the potential coal seams for CO2 sequestration are at depths of 800–1000 m, where temperature and pressure are beyond the critical point of CO2 (Tc = 31.8 °C, Pc = 7.38 MPa) and CO2 exists in its supercritical phase. Supercritical CO2 (ScCO2) exhibits liquid-like density and dissolution power as well as gas-like viscosity and transport behavior [19]. Laboratory experiments have demonstrated that ScCO2 adsorption causes a significantly higher coal matrix swelling effect in comparison with subcritical CO2 adsorption due to the higher uptake capacity of ScCO2 [6,20]. Further, Hol et al. [21] revealed that the adsorption-induced heterogeneous swelling at the maceral scale, accompanied by differential accessibility of the coal microstructure, can form microfractures in the coal mass. In this respect, the sequestrated CO2 may alter the pore morphology of the coal mass during long-term storage.
In addition to the adsorption-induced coal matrix swelling effect, CO2 molecules can dissolve into the coal matrix, especially when CO2 exists in a supercritical state [22,23]. According to Goodman et al. [24], CO2 is a more chemically potential adsorbate than CSG and can act as plasticizer of coal mass, leading to a potential structural rearrangement that affects the molecular structure or intermolecular bonding of coal matrix. Moreover, ScCO2 can act as a solvent capable of extracting organic hydrocarbons from the coal matrix as demonstrated by Kolak and Burruss [25]. Their experiments demonstrated that the extracted amount and type of hydrocarbons varied for different rank coals, with the highest hydrocarbon concentrations mobilized from high volatile bituminous coal.
In consideration of the effects of the fluid-solid interaction between ScCO2 and coal matrix, it is essential to evaluate the chemical and structural alterations in coal with ScCO2 exposure. The alterations in coal properties may further change the flow behavior and adsorption capacities of CO2 on the coal mass, thus influencing the substantial CO2 injection rate and CSG recovery efficiency. Previous studies only considered the effects of ScCO2 on the chemical and structural properties of the coal mass. Systematic knowledge of the adsorption capacity alteration of various rank coals with ScCO2 exposure is still needed to better estimate the feasibility of CO2 sequestration in coal seams.
The objective of this study is to investigate the variation of chemical and structural properties as well as the corresponding alteration of the adsorption capacity in response to ScCO2 exposure for various rank coals. To address this issue, Fourier transform infrared spectroscopy (FTIR) analysis, low-pressure N2 and CO2 sorption method, and high-pressure CO2 adsorption experiments were conducted on four Chinese coals of different rank.

2. Materials and Methods

2.1. Sample Collection and Preparation

Four coal samples collected from coal mines in China were used in this study: lignite from No. 1–2 coal seam in Daliuta coal mine, Shendong coal field (DL coal); middle volatile bituminous coal from No. 2 coal seam in Zhaolou coal mine, Heze coalfield (ZL coal); low volatile bituminous coal from No. C-9 coal seam in Guiyuan coal mine, Guizhou coalfield (GY coal); anthracite from No. 6 coal seam in Datong coal mine, Sichuan coalfield (DT coal). The information of the sampling coal mines is shown in Figure 1.
Coal blocks were first drilled from freshly exposed mining faces and then immediately sealed in a plastic bag and transported to the laboratory. To accelerate the coal–ScCO2 interaction, all the coal lumps were crushed and sieved to particles of 18–20 mesh. Proximate and ultimate analyzes were carried out to characterize coal samples. The analysis results are presented in Table 1.

2.2. Exposure of Coal to ScCO2

In this study, the raw coal samples were firstly saturated to ScCO2 and then further tests were conducted to compare coal properties alteration before and after ScCO2 exposure. A specific high-pressure sealed vessel was fabricated and used to saturate the coal samples with ScCO2 as schematically shown in Figure 2. The heart of this apparatus are pressure cells, an ISCO (260D) syringe pump (Teledyne ISCO, Lincoln, NE, USA) and the data system. The sample was first inserted into the pressure cell with a temperature-controlled thermostatic water bath with an accuracy of ± 0.1 °C. Then, CO2 was injected from the gas cylinder into the sample cell by the ISCO syringe pump. Prior to interaction, the system was degassed under vacuum at 60 °C for 12 h. The reaction pressure was set as 16 MPa, with a constant temperature value of 40 °C, simulating CO2 injection into deep (~800 m) coal seams [17]. The exposure time was set at 30 days to ensure a sufficient coal–ScCO2 interaction. After the saturation period, the sample cell was gradually de-pressurized by 0.1 MPa/min to avoid possible damage to the coal samples. The coal samples were then covered with plastic wrap to avoid any possible changes to the saturation state. All the tests were repeated three times to increase the confidence level and reliability. The average data of three tests were used in this study.

2.3. Fourier Transform Infrared Spectroscopy (FTIR) Analysis

The surface functional groups of both the raw and ScCO2-treated coal samples were determined using a FTIR Nicelet 6700 (Thermo Fisher Scientific, Waltham, MA, USA). The coal samples for FTIR analysis were prepared following the potassium bromide (KBr) pellet method. The coal powders (0.5–1 mg in weight) and the dried KBr were ground at a mass ratio of 1:160. The spectral was obtained in the 400 cm-1 and 4000 cm−1 wavenumber regions with a resolution of 4 cm−1.

2.4. Low-Pressure N2 and CO2 Isotherm Measurements

Low-pressure gas adsorption measurements were conducted on an Accelerated Surface Area and Porosimeter System ASAP 2020M (Micrometrics Instruments Corporation, Norcross, GA, USA). Using standard adsorbates N2 and CO2, information about the mesopore and micropore structure of the coal samples was obtained. Before analysis with either N2 or CO2, the samples (weighing 1–2 g) were degassed under vacuum by heating at 105 °C for 12 h to remove water and other volatile matter. The adsorption-desorption isotherms of N2 were measured at liquid nitrogen (−196.15 °C) under the relative pressure ranging from 0.005 to 0.99. The isotherms of CO2 adsorption were performed at 0 °C under the relative pressure of 0.005–0.032.
At each pressure set point, the sorption equilibrium was established automatically when the system pressure remained stable for 30 s. The absolute pressure tolerance was set as 5 mmHg (6.66 mbar). After each test, the specific surface area (SSA), pore volume, and pore size distribution (PSD) of the coal samples were calculated based on multiple models.

2.5. High-Pressure CO2 Isotherm Measurements

High-pressure CO2 adsorption isotherms were performed using a gravimetric analyzer IGA (Hiden Isochema Limited, Warrington, UK). A detailed description of the IGA system and the general adsorption isotherm test procedure can be found in reference [26]. All isotherms were measured to a pressure of 1.8 MPa at 40 °C. Prior to tests, the samples were degassed at 105 °C under vacuum (<10−6 Pa) for 12 h to remove the water content from the samples. During the experiment, equilibrium was established when the rate of sample weight change was less than 1% or a time limit was reached. Then, the gas pressure was regulated to the next step. The buoyancy effect was corrected automatically in the measurement.

3. Results and Discussion

3.1. Effect of ScCO2 Exposure on Coal Surface Property

The effect of ScCO2 exposure on surface property of various rank coals was examined by FTIR spectroscopy, a technique well suited to study the surface functionalities. The FTIR spectra for the raw and ScCO2-treated coal samples are shown in Figure 3. As presented in Figure 3, the FTIR spectra of the coal samples were typical of complex, heterogeneous organic materials containing multiple type of surface functional groups. With ScCO2 exposure, both the -OH groups (at a wave number of around 3450 cm−1) and the amide carbonyl groups (at a wave number of around 1620 cm−1) show an obvious decrease for DL coal, ZL coal, and GY coal, with only a slight decrease for DT coal. The prominent bands at the wave number around 1000 cm−1 possibly contributed to mineral matter and clay, which show an obvious decrease for all the coals after ScCO2 exposure, indicating an elimination of mineral matter and clay. The bands ranging 900–400 cm−1 show an obvious decrease for DL coal and DT coal after ScCO2 exposure, which is attributed to the Si-O and Si-O-Al bands.
In a comparison of the spectra of the raw and ScCO2-treated coal samples, obvious variations were found in functional group distribution, especially for the low-rank coal. According to Wang et al. [27], the decrease in surface functional group density may be attributable to the extracting effect or reactivity of ScCO2 fluid. The decrease of -OH groups may be related to the dissolution of ScCO2 to the inherent moisture in the coal samples; however, the elimination of the amide carbonyl groups may be related to the extraction of the organic volatile corresponds from coal with ScCO2 exposure, as demonstrated by Li et al. [28]. In addition, a decrease in the distribution of mineral matter content in the coal samples was observed after ScCO2 exposure. However, Mastalerz et al. [29] detected no change in the functional groups of two different coal samples after saturation with CO2. This is because the saturation conditions they conducted were at 20 °C and 4.14 MPa where CO2 was in a sub-critical state in which it cannot extract corresponds from coal matrix. The loss of the organic and inorganic corresponds in the coal matrix with ScCO2 exposure may influence its adsorption capacity for CO2, which will be discussed later in this study.

3.2. Effect of ScCO2 Exposure on Coal Pore Structure

3.2.1. Low-Pressure N2 and CO2 Sorption Isotherms

The adsorption-desorption isotherms of N2 on the raw and ScCO2-treated coal samples at −196 °C are shown in Figure 4. All the isotherms correspond to the type IV isotherm of IUPAC classification [30]. This type of isotherm is considered to be associated with the distribution of mesopores in solids [31]. Distinct adsorption-desorption hysteresis loop is observed in each isotherm at a high relative pressure (>0.45). The presence of the hysteresis loop suggests that capillary condensation occurred within the mesopores. The shape of the hysteresis loop is a reflection of pore morphology within the coal matrix [32]. As shown in Figure 4, there is only a slight difference in the shape of the hysteresis loop in the isotherms between the raw and ScCO2-treated coal samples. It can be deduced that interaction with ScCO2 has a minimal influence on pore shape in various rank coals. This is consistent with previous studies [33,34].
Figure 4 also shows that the adsorbed amount of N2 on the ScCO2-treated coal samples is lower than that on the raw samples, suggesting that the adsorption capacity of N2 on various rank coals uniformly decreased after ScCO2 exposure. At relative pressure less than 0.2, there is no distinct difference in N2 adsorption capacity between the raw and ScCO2-treated coal samples. With the elevated relative pressure, the adsorption amount for the raw coal samples increases more rapidly than the ScCO2-treated coal samples. This implies that exposure to ScCO2 mainly influences the multilayer capacity of N2 adsorption. In this respect, exposure to ScCO2 may have a more distinct effect on the broader mesopores than on the narrow mesopores.
The N2 adsorption-desorption method can only provide mesopore information of the coal samples because N2 molecules cannot penetrate the fine micropores and pore shrinkage of coal at the extremely low temperature of −196.15 °C. Low pressure CO2 adsorption at 0 °C can overcome this drawback due to the shorter equilibrium time and disappearance of the pore shrinkage effect [35].
Figure 5 shows the adsorption isotherms of CO2 on the raw and ScCO2-treated coal samples at 0 °C under a low relative pressure (≤0.033). A comparison of Figure 4 and Figure 5 reveals that, for various rank coals, the adsorption capacity of CO2 is larger than N2 at the same relative pressure due to the accessibility of CO2 molecules to the narrow micropores of coal matrix in comparison with N2 molecules. Figure 5 shows that the low-pressure adsorption capacity of CO2 on the ScCO2-treated coal samples is higher than that on the raw coal samples. This observation suggests an increase of accessibility in the micropores after ScCO2 exposure, potentially attributable to the micro-fracturing of the coal matrix with ScCO2 exposure, as demonstrated by Hol et al. [21].

3.2.2. Pore Properties Determined by Low-Pressure N2 and CO2 Sorption

The mesopore specific surface area (SSA) and total pore volume (Vt) of the coal samples with N2 adsorbate were evaluated by the Brunauer-Emmett-Teller (BET) and Barrett-Joyner-Halenda (BJH) models, respectively. The SSA and micropore volume (Vmic) with CO2 adsorbate were evaluated by Dubinin-Radushkevich (D-R) and Dubinin-Astakhov (D-A) models [27]. These data can be calculated automatically by the analytical software in ASAP2020.
The estimated SSA and pore volume of the raw and ScCO2-treated coal samples from N2 and CO2 sorption data are summarized in Table 2. The results show that, with ScCO2 exposure, the mesopore SSA and Vt of the coal samples are slightly decreased, as seen in Figure 6. As shown in Table 2, reductions of about 0.366 to 0.065 m2/g of mesopore SSA and 9 × 10−3 to 2 × 10−3 cm3/g of Vt values can be observed after ScCO2 exposure. These results indicate that the accessibility of mesopore in various rank coals decreased after ScCO2 exposure. This may have been induced by the irreversible swelling effect of the coal matrix, which may block the pathways of gas molecules and reduce the pore volume [36]. With regard to the average pore width (D), the average pore diameter of GY coal reduced sharply after ScCO2 exposure while the values for DL coal, ZL coal and NM coal increased slightly. This may be due to the larger matrix swelling of GY coal by ScCO2 exposure than the other coal samples.
Unlike mesopore, the values of micropore SSA and Vmic for the coal samples increased notably with ScCO2 exposure, as seen in Figure 7. According to Table 2, the D-R SSA (SD-R) and D-A SSA (SD-A) determined by CO2 adsorption exhibit a similar trend with a 15.8 to 137.0 m2/g and 19.3 to 44.2 m2/g increase for various rank coals, respectively. Simultaneously, increases of 0.005 to 0.014 cm3/g of Vmic values can be observed after ScCO2 exposure. The increased amount of micropore SSA and Vmic in GY coal is the largest, followed by DT coal, DL coal, and ZL coal, respectively.
In combination with the N2 and CO2 adsorption results, the alterations of the pore structure in various rank coals were mainly controlled by two distinct mechanisms: (1) the extraction effect of ScCO2 removes some volatile matters in the pore medium; (2) SCCO2 exposure induces irreversible swelling which may partially block the pore throat in the coal medium. The two mechanisms account for the formation of micropore at the expense of mesopore in the coals after SCCO2 exposure. An increased amount of the micropore is larger than the decreased amount of the mesopore, implying that the combined influence of the two mechanisms results in an increased amount of the adsorptive pore.

3.2.3. Pore Size Distribution Determined by Low-Pressure N2 and CO2 Sorption

The mesopore volume distribution of the raw and ScCO2-treated coal samples derived from BJH model are presented in Figure 8. The results indicate that the PSD of raw and ScCO2-treated coal samples show multimodal distribution in the mesopore range. According to Figure 8, contact with ScCO2 does not change the profile of PSD significantly in the mesopore range for various rank coals. However, the major peaks for ZL, GY, and DT coal were weakened after ScCO2 exposure. This observation is consistent with Gathitu et al. [37], who revealed that the mesopores in coal matrix may collapse during the interaction with ScCO2, resulting in a decrease of the mesopore amount in the coal mass. According to Figure 8, pore size distributions are very similar for GY and DT coal at >5 nm, a potential effect of the similar ranks of GY and DT coal.
The micropore size distribution of the raw and ScCO2-treated coal samples was determined by nonlocal density functional theory (NLDFT), an effective method to calculate pore size 0.5–1 nm; results are shown in Figure 9. In contrast with a similar mesopore size distribution before and after ScCO2 exposure, the micropore size distribution differs between the raw and ScCO2-treated coal samples, indicating a change in micropore morphology after ScCO2 exposure. The pore width of the raw coal samples shows unimodal distribution with major peaks at 0.80–0.88 nm; in contrast, the pore width of ScCO2-treated coal samples shows multimodal distribution with major peaks at around 0.59–0.65 nm, 0.75–0.8 nm, and 0.82–0.85 nm. This implies that a new micropore formed in various rank coals with ScCO2 exposure. As mentioned above, this can be interpreted by micro-fracturing of the coal matrix with ScCO2 exposure. Pan et al. [38] further indicated that many closed pores in the raw coals opened and transmitted to adsorption pore with ScCO2 exposure, resulting in an increase of micropore connectivity.

3.3. Effect of Sc-CO2 Exposure on High-Pressure CO2 Adsorption

The high-pressure adsorption isotherms of CO2 on the raw and ScCO2-treated coal samples are shown in Figure 10. All the isotherms belong to type I, indicating an occurrence of monolayer adsorption [30]. Figure 10 shows that, although ScCO2 exposure does not change the shape of the adsorption isotherms for CO2 adsorption, it significantly affects the adsorption capacities of various rank coals. Notably, ScCO2 exposure can enhance CO2 adsorption at a low pressure; however, the CO2 adsorption capacity for all the ScCO2-treated samples is lower than the raw coal samples at higher pressure. This may be the combined effects of surface property and pore structure alterations for the coal samples with ScCO2 exposure. At a low pressure, the adsorption capacity is mainly dominated by the micropore; the increase of micropore volume leads to higher adsorption capacity after ScCO2 exposure. The adsorption transferred to the mesopores at a higher pressure whereas a decrease in mesopore volume resulted in lower adsorption capacity. In addition, the density of surface functional groups on coal surface can promote CO2 adsorption [39,40]. Thus, the extraction of functional groups on the coal surface by ScCO2 may be another reason for the decrease of CO2 adsorption capacity at a higher pressure.
To describe the high-pressure adsorption isotherms of CO2 on various rank coals before and after ScCO2 exposure, the Langmuir model, which is based on monolayer adsorption, is employed, which was given as:
n = n m b p 1 + b p
where n is adsorbed amount, p is the equilibrium pressure, nm is monolayer adsorption capacity, also known as Langmuir volume, and b is a constant known as Langmuir constant [41].
The fitting parameters for the isotherms were determined by Origin Lab software and the best-fit results are summarized in Table 3. The adsorption isotherms were well fitted by Langmuir with R2 greater than 0.99 for all cases. The values of nm for the raw coals are larger than the ScCO2 treated coals, as seen in Figure 11a, indicating a decrease of monolayer ability after ScCO2 exposure. As mentioned above, this may be due to the reduction of the surface functional group density with ScCO2 exposure. Unlike nm, the values of b for the coal samples were increased, as seen in Figure 11b, suggesting that the monolayer ability can be attained at a lower pressure after ScCO2 exposure.

3.4. Implications for CO2 Sequestration in Coal Seams for Coal Seam Gas (CSG) Enhancement

To date, several projects of ECBM have been carried out in USA, Canada, China, Japan, and Australia. A comprehensive understanding of the effects of ScCO2 exposure on coal seams during the storage is essential to predict the feasibility of these projects [42]. On the basis of the results of this study, ScCO2 will significantly influence the surface property pore structure as well as high-pressure CO2 adsorption capacity in the coal seam during long-term storage. These effects may be aggravated during in-situs conditions, due to the combined effect of ScCO2 and the seam water. Thus, field-scale reservoir tests are still needed to better understand the response of coal seams to CO2 sequestration.

4. Conclusions

In this study, the alterations of surface property and pore structure as well as the corresponding high-pressure CO2 adsorption capacity in various rank coals were experimentally studied on raw and ScCO2-treated coal samples (T = 40 °C, P = 16 MPa). The major conclusions are summarized as follows:
(1)
For various rank coals, the density of the surface functional groups and inorganic corresponds decreased after ScCO2 exposure due to the dissolution and extraction effect of ScCO2;
(2)
The interaction with ScCO2 has a minimal influence on pore shape of various rank coals. The micropore SSA and pore volume were increased while values for mesopore decreased;
(3)
The combined effects of surface property and pore structure alterations lead to a higher CO2 adsorption capacity at a low pressure but a lower CO2 adsorption capacity at a high pressure;
(4)
The adsorption isotherms of the raw and ScCO2-treated coals can be well described by Langmuir model. The monolayer capacity of the various rank coals decreased after exposure to ScCO2, indicating an elimination of the adsorption sites with ScCO2 exposure.
Finally, it can be concluded from the results of this study that ScCO2 exposure has a distinct influence on the surface property, pore structure, and high-pressure CO2 adsorption capacity of the coals, irrespective of its rank.

Author Contributions

Z.Z. and Z.L. conceived and designed the experiments; Z.L. performed the experiments; Z.L. and T.W. analyzed the data; Z.Z. and X.L. contributed materials and analysis tools; Z.L. and X.D. wrote the paper.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 51674047 and 51611140122, National Science Fund for Distinguished Young Scholars, grant number 51625401 and Special Youth Project of Science and Technology Innovation Enterprise Capital of China Coal Technology Engineering Group Co., Ltd., grant number 2018-2-QN016.

Acknowledgments

This study is supported by the National Natural Science Foundation of China (Grant No. 51674047 and 51611140122), National Science Fund for Distinguished Young Scholars (Grant No. 51625401), and Special Youth Project of Science and Technology Innovation Enterprise Capital of China Coal Technology Engineering Group Co., Ltd. (Grant No. 2018-2-QN016).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Leung, D.Y.C.; Caramanna, G.; Maroto-Valer, M.M. An overview of current status of carbon dioxide capture and storage technologies. Renew. Sustain. Energy Rev. 2014, 39, 426–443. [Google Scholar] [CrossRef] [Green Version]
  2. Peter, M.; Li, Z.; Maximilian, R.; Gkiokchan, M.; Yuan, W.; Christian, S.; Martin, R.; Detlef, S. Carbon capture for CO2 emission reduction in the cement industry in Germany. Energies 2019, 12, 2432. [Google Scholar]
  3. Du, X.; Gu, M.; Liu, Z.; Zhao, Y.; Sun, F.; Wu, T. Enhanced shale gas recovery by the injection of CO2, N2 and CO2/N2 mixture gases. Energy Fuels 2019, 33, 5091–5111. [Google Scholar] [CrossRef]
  4. White, C.M.; Smith, D.H.; Jones, K.L.; Goodman, A.L.; Jikich, S.A.; LaCount, R.B.; DuBose, S.B.; Ozdemir, E.; Morsi, B.I.; Schroeder, K.T. Sequestration of carbon dioxide in coal with enhanced coalbed methane recovery—A review. Energy Fuels 2005, 19, 659–724. [Google Scholar] [CrossRef]
  5. Busch, A.; Gensterblum, Y. CBM and CO2-ECBM related sorption processes in coal: A review. Int. J. Coal. Geol. 2011, 87, 49–71. [Google Scholar] [CrossRef]
  6. Perera, M.S.A. A comprehensive overview of CO2 flow behaviour in deep coal seams. Energies 2018, 11, 906. [Google Scholar] [CrossRef]
  7. Gan, H.; Nandi, S.P.; Walker, P.L., Jr. Nature of the porosity in American coals. Fuel 1972, 51, 272–277. [Google Scholar] [CrossRef]
  8. Li, Q.; Lin, B.; Wang, K.; Zhao, M.; Ruan, M. Surface properties of pulverized coal and its effects on coal mine methane adsorption behaviors under ambient conditions. Powder Technol. 2015, 270, 278–286. [Google Scholar] [CrossRef]
  9. Wang, F.; Cheng, Y.; Lu, S.; Jin, K.; Zhao, W. Influence of coalification on the pore characteristics of middle-high rank coal. Energy Fuels 2014, 28, 5729–5736. [Google Scholar] [CrossRef]
  10. Jian, K.; Fu, X.; Ding, Y.; Wang, H.; Li, T. Characteristics of pores and methane adsorption of low-rank coal in China. J. Nat. Gas Sci. Eng. 2015, 27, 207–218. [Google Scholar] [CrossRef]
  11. Okolo, G.N.; Everson, R.C.; Neomagus, H.W.J.P.; Roberts, M.J.; Sakurovs, R. Comparing the porosity and surface areas of coal as measured by gas adsorption, mercury intrusion and SAXS techniques. Fuel 2015, 141, 293–304. [Google Scholar] [CrossRef]
  12. Pini, R.; Ottiger, S.; Storti, G.; Mazzotti, M. Pure and competitive adsorption of CO2, CH4 and N2 on coal for ECBM. Energy Procedia 2009, 1, 1705–1710. [Google Scholar] [CrossRef]
  13. Jasinge, D.; Ranjith, P.G.; Choi, X.; Fernando, J. Investigation of the influence of coal swelling on permeability characteristics using natural brown coal and reconstituted brown coal specimens. Energy 2012, 39, 303–309. [Google Scholar] [CrossRef]
  14. Day, S.; Fry, R.; Sakurovs, R. Swelling of Australian coals in supercritical CO2. Int. J. Coal Geol. 2008, 74, 41–52. [Google Scholar] [CrossRef]
  15. Vandamme, M.; Brochard, L.; Lecampion, B.; Coussy, O. Adsorption and strain: The CO2-induced swelling of coal. J. Mech. Phys. Solids 2010, 58, 1489–1505. [Google Scholar] [CrossRef]
  16. Pluymakers, A.; Liu, J.; Kohler, F.; Renard, F.; Dysthe, D. A high resolution interferometric method to measure local swelling due to CO2 exposure in coal and shale. Int. J. Coal Geol. 2018, 187, 131–142. [Google Scholar] [CrossRef]
  17. Perera, M.S.A.; Ranjith, P.G.; Choi, S.K.; Bouazza, A.; Kodikara, J.; Airey, D. A review of coal properties pertinent to carbon dioxide sequestration in coal seams: With special reference to Victorian Brown coals. Environ. Earth Sci. 2011, 64, 223–235. [Google Scholar] [CrossRef]
  18. Ranathunga, A.S.; Perera, M.S.A.; Ranjith, P.G.; Rathnaweera, T.D.; Zhang, X.G. Effect of coal rank on CO2 adsorption induced coal matrix swelling with different CO2 properties and reservoir depths. Energy Fuels 2017, 31, 5297–5305. [Google Scholar] [CrossRef]
  19. Vishal, V.; Singh, T.N. A laboratory investigation of permeability of coal to supercritical CO2. Geotech. Geol. Eng. 2015, 33, 1009–1016. [Google Scholar] [CrossRef]
  20. Perera, M.S.A.; Ranjith, P.G.; Choi, S.K.; Airey, D. The effects of sub-critical and super-critical carbon dioxide adsorption-induced coal matrix swelling on the permeability of naturally fractured black coal. Energy 2011, 36, 6442–6450. [Google Scholar] [CrossRef]
  21. Hol, S.; Spiers, C.J.; Peach, C.J. Microfracturing of coal due to interaction with CO2 under unconfined conditions. Fuel 2012, 97, 569–584. [Google Scholar] [CrossRef]
  22. Clarksona, C.R.; Bustin, R.M. The Effect of pore structure and gas pressure upon the transport properties of coal: A laboratory and modeling study.1: Isotherms and pore volume distributions. Fuel 1999, 78, 1333–1344. [Google Scholar] [CrossRef]
  23. Larsen, J.W. The effects of dissolved CO2 on coal structure and properties. Int. J. Coal Geol. 2004, 57, 63–70. [Google Scholar] [CrossRef]
  24. Goodman, A.L.; Favors, R.N.; Larsen, J.W. Argonne coal structure rearrangement caused by sorption of CO2. Energy Fuel 2006, 20, 2537–2543. [Google Scholar] [CrossRef]
  25. Kolak, J.J.; Burruss, R.C. Geochemical investigation of the potential for mobilizing non-methane hydrocarbons during carbon dioxide storage in deep coal beds. Energy Fuels 2006, 20, 566–574. [Google Scholar] [CrossRef]
  26. Liu, Z.; Zhang, Z.; Lu, Y.; Choi, S.K. Surface properties and pore structure of anthracite, bituminous coal and lignite. Energies 2018, 11, 1502. [Google Scholar] [CrossRef]
  27. Wang, Q.; Zhang, D.; Wang, H.; Jiang, W.; Wu, X.; Yang, J.; Huo, P. Influence of CO2 exposure on high-pressure methane and CO2 adsorption on various rank coals: Implications for CO2 sequestration in coal seams. Energy Fuels 2015, 29, 3785–3795. [Google Scholar] [CrossRef]
  28. Li, H.; Chang, Q.; Dai, Z.; Chen, X.; Wang, F. Upgrading effects of supercritical carbon dioxide extraction on physicochemical characteristics of Chinese low-rank coals. Energy Fuels 2017, 31, 13305–13316. [Google Scholar] [CrossRef]
  29. Emmett, P.H.; Dewitt, T.W. The low temperature adsorption of Nitrogen, Oxygen, Argon, Hydrogen, N–Butane and carbon dioxide on porous glass and on Partially Dehydrated Chabazite. J. Am. Chem. Soc. 1943, 65, 617–621. [Google Scholar] [CrossRef]
  30. Sing, K.S.W.; Everett, D.H.; Haul, R.A.W.; Mouscou, L.; Pierotti, R.A.; Rouquerol, J.; Siemieniewska, T. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity. Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  31. Mastalerz, M.; Drobniak, A.; Rupp, J. Meso- and micropore characteristics of coal lithotypes: Implications for CO2 adsorption. Energy Fuels 2008, 22, 4049–4061. [Google Scholar] [CrossRef]
  32. Boer, J.H.D.; Lippens, B.C. Studies on pore systems in catalysts II. The shapes of pores in aluminum oxide systems. J. Catal. 1964, 3, 38–43. [Google Scholar]
  33. Zhang, D.; Gu, L.; Li, S.; Lian, P.; Tao, J. Interactions of supercritical CO2 with coal. Energy Fuels 2013, 27, 387–393. [Google Scholar] [CrossRef]
  34. Kutchko, B.G.; Goodman, A.L.; Rosenbaum, E.; Natesakhawat, S.; Wagner, K. Characterization of coal before and after supercritical CO2 exposure via feature relocation using field-emission scanning electron microscopy. Fuel 2013, 107, 777–786. [Google Scholar] [CrossRef]
  35. Zhao, J.; Xu, H.; Tang, D.; Mathews, J.; Li, S.; Tao, S. A comparative evaluation of coal specific surface area by CO2 and N2 adsorption and its influence on CH4 adsorption capacity at different pore sizes. Fuel 2016, 183, 420–431. [Google Scholar] [CrossRef]
  36. Do, D.D.; Do, H.D. Appropriate volumes for adsorption isotherm studies: The absolute void volume, accessible pore volume and enclosing particle volume. J. Colloid Interface Sci. 2007, 316, 317–330. [Google Scholar] [CrossRef]
  37. Gathitu, B.B.; Chen, W.; Mcclure, M. Effects of coal interaction with supercritical CO2: Physical structure. Ind. Eng. Chem. Res. 2009, 48, 5024–5034. [Google Scholar] [CrossRef]
  38. Pan, Z.; Niu, Q.; Wang, K.; Shi, X.; Li, M. The closed pores of tectonically deformed coal studied by small-angle X-ray scattering and liquid nitrogen adsorption. Micropor. Mesopor. Mater. 2016, 224, 245–252. [Google Scholar] [CrossRef]
  39. Chen, Y.; Mastalerz, M.; Schimmelmann, A. Characterization of chemical functional groups in macerals across different coal ranks via micro-FTIR spectroscopy. Int. J. Coal Geol. 2012, 104, 22–33. [Google Scholar] [CrossRef]
  40. Huang, X.; Chu, W.; Sun, W.; Jiang, C.; Feng, Y.; Xue, Y. Investigation of oxygen-containing group promotion effect on CO2–coal interaction by density functional theory. Appl. Surf. Sci. 2014, 299, 162–169. [Google Scholar] [CrossRef]
  41. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef]
  42. Zhang, K.; Cheng, Y.; Jin, K.; Guo, H.; Li, W. Effects of supercritical CO2 fluids on pore morphology of coal: Implications for CO2 geological sequestration. Energy Fuels 2017, 31, 4731–4741. [Google Scholar] [CrossRef]
Figure 1. The information of the sampling coal mines: (a) location, and (b–e) stratigraphic column.
Figure 1. The information of the sampling coal mines: (a) location, and (b–e) stratigraphic column.
Energies 12 03294 g001
Figure 2. Schematic diagram of coal–ScCO2 interaction system.
Figure 2. Schematic diagram of coal–ScCO2 interaction system.
Energies 12 03294 g002
Figure 3. Fourier transform infrared spectroscopy (FTIR) spectra for the raw and ScCO2-treated coal samples: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Figure 3. Fourier transform infrared spectroscopy (FTIR) spectra for the raw and ScCO2-treated coal samples: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Energies 12 03294 g003
Figure 4. N2 adsorption-desorption isotherms on the raw and ScCO2-treated coal samples at −196 °C: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Figure 4. N2 adsorption-desorption isotherms on the raw and ScCO2-treated coal samples at −196 °C: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Energies 12 03294 g004
Figure 5. Low-pressure CO2 adsorption isotherms on the raw and ScCO2-treated coal samples at 0 °C: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Figure 5. Low-pressure CO2 adsorption isotherms on the raw and ScCO2-treated coal samples at 0 °C: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Energies 12 03294 g005
Figure 6. Mesopore specific surface area (SSA) and volume of the raw and ScCO2-treated coal samples: (a) mesopore SSA; (b) mesopore volume.
Figure 6. Mesopore specific surface area (SSA) and volume of the raw and ScCO2-treated coal samples: (a) mesopore SSA; (b) mesopore volume.
Energies 12 03294 g006
Figure 7. D-R micropore SSA and volume of the raw and ScCO2-treated coal samples: (a) micropore SSA; (b) micropore volume.
Figure 7. D-R micropore SSA and volume of the raw and ScCO2-treated coal samples: (a) micropore SSA; (b) micropore volume.
Energies 12 03294 g007
Figure 8. Pore size distribution (PSD) for the raw and ScCO2-treated coal samples determined by N2 adsorption: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Figure 8. Pore size distribution (PSD) for the raw and ScCO2-treated coal samples determined by N2 adsorption: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Energies 12 03294 g008
Figure 9. PSD for the raw and ScCO2-treated coal samples determined by CO2 adsorption: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Figure 9. PSD for the raw and ScCO2-treated coal samples determined by CO2 adsorption: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Energies 12 03294 g009
Figure 10. Adsorption isotherms of CO2 on the raw and ScCO2-treated coal samples at 40 °C and the fitting curves by Langmuir model: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Figure 10. Adsorption isotherms of CO2 on the raw and ScCO2-treated coal samples at 40 °C and the fitting curves by Langmuir model: (a) DL coal; (b) ZL coal; (c) GY coal; (d) DT coal.
Energies 12 03294 g010
Figure 11. Fitting parameters for Langmuir model: (a) nm; (b) b.
Figure 11. Fitting parameters for Langmuir model: (a) nm; (b) b.
Energies 12 03294 g011
Table 1. Proximate and ultimate analyzes of the coal samples.
Table 1. Proximate and ultimate analyzes of the coal samples.
SampleRo max (%)Proximate Analysis (wt %)Ultimate Analysis (wt % daf)
CfixVdafAadMCHNO
DL coal0.4253.5536.9619.422.4672.714.951.1920.50
ZL coal0.8165.1028.7116.241.5082.544.571.0811.03
GY coal1.1478.346.1410.522.0886.924.011.206.03
DT coal1.8673.8512.8113.341.9690.113.790.962.05
Notes: Cfix, fixed carbon; Vdaf, volatile matters; Aad, ash; M, moisture.
Table 2. Structural properties of the raw and ScCO2-treated coal samples determined by N2 and CO2 sorption.
Table 2. Structural properties of the raw and ScCO2-treated coal samples determined by N2 and CO2 sorption.
SampleStateN2 AdsorptionCO2 Adsorption
SBET/m2·g−1Vt/cm3·g−1D/nmSD-R/m2·g−1SD-A/m2·g−1Vmic/cm3·g−1
DL coalraw coal1.4980.006818.1698.7161.050.0216
ScCO2-treated 1.1320.006021.20118.584.180.0327
ZL coalraw coal0.7440.004323.1254.5445.770.0219
ScCO2-treated 0.6110.004126.8492.6565.100.0268
GY coalraw coal0.6160.002818.19222.4160.90.0667
ScCO2-treated 0.5510.001913.79359.4205.10.0810
DT coalraw coal0.5860.001610.92120.474.340.0264
ScCO2-treated 0.3870.001111.37176.6100.60.0353
Table 3. Fitting parameters of the Langmuir model for CO2 adsorption at 40 °C.
Table 3. Fitting parameters of the Langmuir model for CO2 adsorption at 40 °C.
SampleStatenm/mmol·g−1BR2
DL coalraw coal0.1022.9230.9993
ScCO2-treated 0.0924.0560.9954
ZL coalraw coal0.1810.4160.9991
ScCO2-treated 0.1100.9060.9975
GY coalraw coal0.2432.5380.9990
ScCO2-treated 0.2114.0500.9964
DT coalraw coal0.2021.9940.9981
ScCO2-treated 0.1635.0290.9960

Share and Cite

MDPI and ACS Style

Liu, Z.; Zhang, Z.; Liu, X.; Wu, T.; Du, X. Supercritical CO2 Exposure-Induced Surface Property, Pore Structure, and Adsorption Capacity Alterations in Various Rank Coals. Energies 2019, 12, 3294. https://doi.org/10.3390/en12173294

AMA Style

Liu Z, Zhang Z, Liu X, Wu T, Du X. Supercritical CO2 Exposure-Induced Surface Property, Pore Structure, and Adsorption Capacity Alterations in Various Rank Coals. Energies. 2019; 12(17):3294. https://doi.org/10.3390/en12173294

Chicago/Turabian Style

Liu, Zhenjian, Zhenyu Zhang, Xiaoqian Liu, Tengfei Wu, and Xidong Du. 2019. "Supercritical CO2 Exposure-Induced Surface Property, Pore Structure, and Adsorption Capacity Alterations in Various Rank Coals" Energies 12, no. 17: 3294. https://doi.org/10.3390/en12173294

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