Next Article in Journal
The Impact of Behavioral and Structural Remedies on Electricity Prices: The Case of the England and Wales Electricity Market
Next Article in Special Issue
An Integrally Embedded Discrete Fracture Model with a Semi-Analytic Transmissibility Calculation Method
Previous Article in Journal
Efficient Economic and Resilience-Based Optimization for Disaster Recovery Management of Critical Infrastructures
Previous Article in Special Issue
Retention of Hydraulic Fracturing Water in Shale: The Influence of Anionic Surfactant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Laboratory Study on Changes in the Pore Structures and Gas Desorption Properties of Intact and Tectonic Coals after Supercritical CO2 Treatment: Implications for Coalbed Methane Recovery

1
State Key Laboratory of Coal Mine Disaster Dynamics and Control, Chongqing University, Chongqing 400044, China
2
State Key Laboratory of Gas Disaster Monitoring and Emergency Technology, Chongqing 400037, China
3
China Coal Technology and Engineering Group Chongqing Research Institute, Chongqing 400037, China
4
Zhengzhou Engineering Co., Ltd. of China Railway Seventh Group, Zhengzhou 450052, China
*
Authors to whom correspondence should be addressed.
Energies 2018, 11(12), 3419; https://doi.org/10.3390/en11123419
Submission received: 31 October 2018 / Revised: 26 November 2018 / Accepted: 3 December 2018 / Published: 6 December 2018
(This article belongs to the Special Issue Sustainability of Fossil Fuels)

Abstract

:
Tectonic coals in coal seams may affect the process of enhanced coalbed methane recovery with CO2 sequestration (CO2-ECBM). The main objective of this study was to investigate the differences between supercritical CO2 (ScCO2) and intact and tectonic coals to determine how the ScCO2 changes the coal’s properties. More specifically, the changes in the tectonic coal’s pore structures and its gas desorption behavior were of particular interest. In this work, mercury intrusion porosimetry, N2 (77 K) adsorption, and methane desorption experiments were used to identify the difference in pore structures and gas desorption properties between and intact and tectonic coals after ScCO2 treatment. The experimental results indicate that the total pore volume, specific surface area, and pore connectivity of tectonic coal increased more than intact coal after ScCO2 treatment, indicating that ScCO2 had the greatest influence on the pore structure of the tectonic coal. Additionally, ScCO2 treatment enhanced the diffusivity of tectonic coal more than that of intact coal. This verified the pore structure experimental results. A simplified illustration of the methane migration before and after ScCO2 treatment was proposed to analyze the influence of ScCO2 on the tectonic coal reservoir’s CBM. Hence, the results of this study may provide new insights into CO2-ECBM in tectonic coal reservoirs.

1. Introduction

As a major greenhouse gas, CO2 causes global warming and initiated a series of negative effects on the balance of the natural ecosystem and the sustainable development of human society [1,2,3,4,5,6,7]. To mitigate CO2 emissions into the atmosphere, enhanced coalbed methane recovery with CO2 sequestration (CO2-ECBM) is considered to be a promising technology, and it receives widespread attention [8,9,10,11,12]. This technology can not only provide CO2 storage, but can also enhance the production of coalbed methane (CBM). To date, many scholars proved the feasibility of injecting CO2 into methane-bearing coal seams using various experimental methods and field tests. Fulton et al. observed that the recovery of CH4 after CO2 injection was 57% higher than that of natural emissions [13]. Reznik et al., Jessen et al., and Dutka et al. also verified that CO2 injection could contribute to the improvement of CH4 recovery [14,15,16]. In 1996, the San Juan Basin was selected as the first demonstration site for CO2-ECBM. After CO2 injection, CBM production was increased by five times, and the recovery of CH4 reached 77–95%, lasting for over six years [17].
The coal seams most suitable for CO2 sequestration are commonly more than 800 m deep, and the temperature and pressure in coal seams at the most suitable depths are, in many cases, above the critical point for CO2 (critical temperature and pressure: 31.05 °C and 7.39 MPa) [18]. Numerous studies were undertaken to better understand the interaction between supercritical CO2 (ScCO2) and intact coal. When coal is exposed to ScCO2, the ScCO2 is incorporated into the coal and it rearranges the coal’s structure [19]. It is generally agreed that ScCO2 can change the pore morphology of coal irreversibly, including pore volume, pore size distribution, surface area, and pore connectivity [20,21]. Additionally, ScCO2 is an organic solvent and can mobilize some of the organic matter in the coal. If some of the CO2 is dissolved in the reservoir water, the pH decreases and the CO2-enriched solution can dissolve some of the minerals present [22,23,24]. Studies also showed that CO2 adsorption can cause the coal’s matrix to swell [25]. However, the above studies were all focused on the effects of ScCO2 on intact coal, and there was little research on tectonic coal–ScCO2 interactions.
Tectonic coal is deformed, sheared coal. The deformation and shearing was caused by tectonism that occurred long after coal formation and diagenesis of the coal seams [26]. Tectonic coal is present all over the world, and there is substantial tectonized coal in China [27,28]. The existence of tectonic coal in some coalfields complicates safe and efficient CO2 sequestration. This is because the physical and chemical properties of tectonic coal differ from those of intact coal. Additionally, the permeability of tectonic coal is much lower than that of intact coal. This leads to the formation of CBM-rich coals; thus, tectonic coals have huge CBM exploitation potential. However, a CBM extraction well drilled in an area hosting significant tectonic coal may seriously restrict CBM production because the tectonic coal’s low permeability restricts gas desorption and migration [29,30]. Therefore, in such an area, CO2-ECBM is likely to be the technique implemented to solve this problem.
The coal’s pore structure affects gas adsorption, diffusion, and seepage. Previous studies [31,32] showed that the pore structure in intact coal and tectonic coal are different and affect how gas desorbs from these coals. These differences are important when the subject of long-term CO2 storage in coal seams is considered. However, limited research focused on this topic, especially the comprehensive and systematic analysis of the differences of pore structures and gas desorption properties between intact coal and tectonic coal during CO2-ECBM. Therefore, a better understanding of the pore morphologies and desorption in intact and tectonic coals after ScCO2 treatment is of significance for CO2-ECBM in tectonic coal reservoirs.
In this study, the effects of ScCO2 on the pore structures of intact and tectonic coals was studied using mercury intrusion porosimetry (MIP) and N2 (77 K) adsorption experiments. Additionally, gas desorption of intact and tectonic coals after ScCO2 treatment was analyzed using methane desorption experiments. The main focus of this study was the analysis of the pore structures and gas desorption properties of intact and tectonic coals before and after ScCO2 treatment, and results were compared to analyze the influence of tectonic coal on the CO2-ECBM process.

2. Materials and Methods

2.1. Coal Samples

Both the intact coal and tectonic coal samples used in the experiments were collected from the Shanxi Formation #3 coal seam in the Sihe coal mine. This mine is in the southern Qinshui basin in Shanxi Province, China (Figure 1). The southern Qinshui basin is not only a major CBM production area, but also the area in which China first carried out a pilot CO2-ECBM project [33]. The coal samples for this study were collected from fresh working faces in the mine. The samples were then immediately sealed, packed, and sent to the laboratory with minimal delay to prevent oxidation. At the laboratory, standard crushing and screening equipment was used to crush and screen the coal samples to different particle sizes for the different experiments. For each experiment, the crushed and sieved intact coal and the tectonic coal sub-samples were divided into two fractions for comparative analyses: one fraction was treated with ScCO2, and the other was not treated. Selected properties of the intact coal and the tectonic coal are listed in Table 1. Compared with intact coal, the mineral content and Langmuir volume of the tectonic coal is greater, indicating that tectonic coal contains more minerals and can absorb more methane. The moisture, ash, and Ro, max values of the tectonic coal are slightly higher than those of the intact coal.

2.2. Experimental Methods

For this study, MIP, N2 (77 K) adsorption, and methane desorption experiments on both intact and tectonic coal specimens with and without ScCO2 treatment were conducted.
A schematic of the set-up used for the ScCO2 treatment is shown in Figure 2. According to the pressure gradient and temperature gradient in the southern Qinshui basin, the temperature of the reservoir is 35 °C and the pressure is 8 MPa at the depth of 800 m. Therefore, 35 °C and 8 MPa were chosen as the treatment conditions to replicate the in situ conditions. Before the ScCO2 treatment, coal samples were first degassed in a vacuum chamber at 50 °C and 4 Pa for 24 h. Then, the constant temperature bath was set to 35 °C and the sample tank was filled with CO2 using an ISCO pump. Subsequently, the pressure was maintained at 8 MPa for 72 h, and the constant temperature bath was held at 35 °C.
The pore structure of both ScCO2-treated and untreated coal samples was analyzed using a PM33-GT-12 mercury porosimeter. This instrument can measure pore diameters between 0.007 and 1000 μm over a pressure range of 1.5–231,000 kPa. The porosimetry data were modeled using the Washburn Equation [34]:
r = 2 σ cos θ p c ,
where r is the pore radius of the porous material (nm), σ is the surface tension of mercury (dyn/cm2), θ is the contact angle between mercury and the porous material’s surface (°), and pc is the capillary pressure (MPa).
The pore size distribution (PSD) was determined using the N2 (77 K) adsorption method employing a Quadrasorb SI instrument. The N2 adsorption–desorption isotherms of ScCO2-treated and untreated coal samples were obtained at a temperature of −196 °C with a relative pressure (p/p0) range of 0.01–0.99. The Barrett-Joyner-Halenda (BJH) method was used to calculate total pore volume (TPV), density functional theory (DFT) was used to calculate PSD, and the Brunauer-Emmett-Teller (BET) method was used to calculate the specific surface area (SSA) [35,36,37].
Methane desorption data from both treated and untreated intact and tectonic coals were collected during the gas desorption experiments. The experiments were run according to China National Standards AQ/T 1065-2008 and GB 474-2008; the experimental set-up is shown in Figure 3. Coal particles from 1 to 3 mm in diameter were placed in a container and degassed at 50 °C and 4 Pa for 24 h. After degassing, the container, in a 30 °C water bath, was quickly filled with 99.9% pure methane to achieve the methane adsorption equilibrium (2 MPa and 4 MPa). When adsorption equilibrium was reached, valve 4 was opened to vent all the free methane from the container. The container was then connected to the gas desorption measuring cylinder for the gas desorption experiment. During the experiment, the volume of methane desorbed was recorded at specific times; the desorption segment of each experiment lasted two hours.
The whole experimental process is shown in Figure 4.

3. Results and Discussion

3.1. Pore Structure Analysis

3.1.1. Pore Size Distribution

To analyze the MIP data from treated and untreated intact and tectonic coal samples quantitatively, the Hodot classification for coal pore sizes was used. This classification scheme divides the pores into macropores (>1000 nm), mesopores (100–1000 nm), transition pores (10–100 nm), and micropores (<10 nm) [38].
Figure 5 shows the PSDs for both ScCO2-treated and untreated coal samples. It is clear from the figure that the macropores and mesopores in the tectonic coal were more developed. This effect was reported by a number of previous studies [39,40]. It is interesting to note that the macropores and mesopores in the tectonic coal were obviously larger after ScCO2 treatment, and the effect of the ScCO2 treatment on the tectonic coal’s pores was greater than its effect on the intact coal’s pores. According to the data in Table 2, the proportion of macropores and mesopores in the tectonic coal was 33.74% before ScCO2 treatment, but 36.02% after treatment, greater than the pore percentages in the intact coal (15.91% before and 21.37% after). It can be deduced that SCCO2 treatment may observably promote the development of mesopores and macropores of tectonic coal. The data also show that the porosity of the treated intact and tectonic coal samples increased by 18.90% (from 4.18% to 4.97%) and 23.14% (from 5.23% to 6.44%), respectively. Previous studies [22,23,41] showed that ScCO2 can mobilize some of the polycyclic aromatic hydrocarbons and aliphatic hydrocarbons in the coal. Additionally, CO2 will form carbonic acid when it is dissolved in the water in the coal seams, and this acid can dissolve some of the inorganic minerals in the coal, such as calcites, dolomites, and magnesites [42]. This may result in increased pore sizes. Additionally, Cao et al. found that the mean extraction yield of the tectonic coal was 1.45% using organic solvent, whereas that of the intact coal was 0.44% [43]. Apparently, the mobilizing effect in tectonic coal is more pronounced than its effect in intact coal. These MIP data show that the changes in PSD for the transition pores and micropores were less significant. To investigate the PSDs of the smaller pores in more detail, additional analyses were performed using N2 (77 K) adsorption.
Using the N2 (77 K) manometric adsorption technique can avoid the destruction of pores under high pressure; thus, it is commonly used for PSD measurements of smaller pores [26]. As shown in Figure 6, after ScCO2 treatment, the PSDs for small pores were not significantly affected, although a slight tendency for pore sizes to increase could be discerned. This is consistent with the MIP results. The PSDs of both the treated and the untreated intact and tectonic coals had two peaks in the portion of the isotherm, representing pores with diameters smaller than 10 nm, indicating that the micropore structure of these coal samples was relatively developed.

3.1.2. Pore Volume and Surface Area

Table 3 summarizes the changes in the TPVs and SSAs for untreated and ScCO2-treated intact and tectonic coal samples. Overall, the BJH-TPVs and the BET-SSAs (the values determined by N2 (77 K) adsorption) showed a slight increase after ScCO2 treatment; the TPVs and SSAs determined by MIP showed a more significant increase. These increases are consistent with the PSD results. Additionally, the TPVs and SSAs of both the treated and untreated tectonic coal samples were higher than those for the intact coal samples. A study by Qu et al. [44] explained this point very well, and suggested that the tectonism that the coal underwent greatly promoted the fracture of macromolecular chains and aromatic layers, which was conducive to the development of molecular structural disorder in the coal. This led to increased pore volume and, hence, increased surface area.

3.1.3. Pore Connectivity

The pores in coal can be divided into four types according to their shapes: cross-linked, passing, dead end, and closed pores. The first three types are called open pores [45]. Some useful information about the pores can be extracted from the MIP injection/ejection curves and the hysteresis loops. As shown in Figure 7, after ScCO2 treatment, the mercury injection volumes for the intact coal and the tectonic coal increased by 19.09% (from 0.0220 to 0.0262 mL/g) and 29.45% (from 0.0326 to 0.0422 mL/g), respectively. The mercury withdrawal volume for these two coals showed the same trend, increasing by 9.14% (from 0.0197 to 0.0215 mL/g) and 17.42% (from 0.0264 to 0.0310 mL/g), indicating that the open pore volume increased after ScCO2 treatment [46]. Moreover, the hysteresis loops of all treated coals increased, indicating that the pore network was more complex with more bottleneck pores, and there was an increase in the number of open pores [47]. Thus, the connectivity of the treated tectonic coal sample was excellent, which may affect the methane desorption, diffusion, and seepage.

3.2. Desorption Analysis

Carbon dioxide sequestration in coal seams enhances CBM production. Simplistically, CO2 sequestration can be thought of as just gas in/gas out: carbon dioxide seepage, diffusion, and adsorption in, coupled with methane desorption, diffusion, and seepage out. The preceding pore structure analysis showed that ScCO2 can change the pore structure in intact and tectonic coals, and this will undoubtedly affect the desorption, diffusion, and seepage of the methane. Previous studies showed that the diffusion of methane through the coal’s matrix has an important influence on CBM production [48,49]. However, research on how ScCO2 affects methane diffusion in tectonic coal is still limited. Therefore, the methane desorption from ScCO2-treated intact and tectonic coals was analyzed.

3.2.1. Gas Desorption Curves

Methane desorption curves for untreated and treated coal samples are shown in Figure 8. The cumulative desorption volume increased as the desorption time increased. The slopes of the desorption curves were steepest in the early stages of the desorption experiments, and the slopes gradually decreased with time. Compared with intact coal, the slopes of the desorption curves of tectonic coal were greater. This is because the mesopores and macropores of tectonic coal were more developed. When the adsorption equilibrium pressure was 4 MPa, the volumes of gas desorbed from the coal samples were higher than the volumes of gas desorbed when the absorption equilibrium pressure was 2 MPa. This is because, during adsorption, the coal matrix could absorb more methane at higher equilibrium pressures [50].
In general, ScCO2 treatment has a considerable effect on the coal sample’s gas desorption. As shown in Figure 7, when the equilibrium pressure was 4 MPa and 2 MPa, the final desorption volume of treated tectonic coal increased by 2.8118 mL/g and 3.8073 mL/g, respectively. However, under the same conditions, the final volume of gas desorbed from treated intact coal was slightly greater than volume of gas desorbed from untreated tectonic coal (2.1950 mL/g and 3.1542 mL/g). The final desorption volume for treated tectonic coal was 1.33 to 1.40 times that of the desorption volume for intact coal. These results indicate that more CBM can be extracted from a tectonic coal CBM reservoir after the coal in the reservoir is injected with ScCO2.
To further analyze the changes in gas desorption from ScCO2-treated intact and tectonic coal, their gas diffusion coefficients were compared in the subsequent section.

3.2.2. Gas Desorption Diffusion Coefficients

Based on Fick’s diffusion laws, an analytic solution for a diffusion equation was presented by Crank [51] under Dirichlet boundary conditions (Equation (2)). However, this analytical solution is difficult to apply to practical engineering problems because it is in the form of an infinite series. Yang [52] simplified Equation (1) to a more practical form (Equation (3)), and this equation is widely used in engineering and can represent coal’s gas desorption very well. The two equations are as follows:
Q t Q = 1 6 π 2 n = 1 1 n 2 e D n 2 π 2 t r c 2 ,
Q t Q = 1 e B 1 t ,
where D is a diffusion coefficient, t is the desorption time, rc is the average particle diameter, Q t is the cumulative desorption amount at time t, Q is the ultimate desorption amount, and B1 is a fitting parameter related to the diffusion coefficient D and the average diameter rc, which can be calculated from B1 = K((4π2D)/rc2). K is a correction parameter commonly taken to be equal to 1.
The variable Q is, in most cases, calculated using Equation (4) [53].
Q = ( V L P e q P L + P e q V L P a P L + P a ) ( 1 M a d A a d ) ,
where P e q is the definite equilibrium pressure, P a is the atmospheric pressure, M a d is the moisture content, and A a d is the ash content of the coal samples.
Equations (3) and (4) can be used to calculate the diffusion coefficients for the coals used in this study using the data from the methane desorption curves in Figure 7 as input parameters. The results are shown in Table 4.
According to the results listed in Table 4, it is clear that ScCO2 treatment considerably increased the tectonic coal’s diffusion coefficient, and the treatment increased the tectonic coal’s diffusion coefficient more than it increased the coefficient for intact coal. For example, when the equilibrium pressure was 2 MPa, ScCO2 treatment increased the diffusion coefficient for treated tectonic coal by 87.59%, but the treatment only increased intact coal’s diffusion coefficient by 45.98%. It can also be seen in Table 4 that, under the same conditions, tectonic coal’s diffusion coefficients were higher than those for intact coal. This is because desorption largely depends on the properties of the pores and the pore structure in the coal. As indicated by the MIP test results, the proportion of macropores and mesopores increased after ScCO2 treatment, especially for tectonic coal, and larger pores can provide better channels for gas migration. In short, tectonic coal has stronger diffusivity capacity.

3.2.3. Implication for CO2-ECBM in Tectonic Coal Reservoirs

Liu et al. surveyed the distribution of tectonic coal reservoirs in China and indicated that the CBM resource in these reservoirs was as much as 5.60 trillion cubic meters. This amounts to 39.20% of China’s CBM resources [54]. Although the tectonic coal reservoirs have great potential for CBM development, the permeability of this type of gas reservoir is very low. This means that developing the resources in these CMB reservoirs would be very difficult.
A simplified illustration of the migration of methane in a coal seam is show in Figure 9. In Figure 9, methane migration is divided into three stages: desorption, diffusion, and seepage. Red circles represent adsorbed methane in the coal matrix, green circles represent free methane, yellow squares represent the coal skeleton, and purple squares represent pores in the matrix. In tectonic coal reservoirs, gas migration to the fractures is slow because of the tectonic coal’s low permeability. Therefore, only a small amount of methane can flow to CBM wells and the production of CBM is low (Figure 9, untreated). The results of the methane desorption experiments described in previous sections show that both the desorption capacities and diffusion coefficients of treated coals were higher than in untreated coals. Therefore, more adsorbed methane was desorbed from the coal’s matrix and diffused to the fractures after CO2 was injected into the tectonic coal reservoir (Figure 9, treated). The pressure gradient must increase when more free methane is present in the fractures, as shown in Figure 9, treated (more green circles exist in the fracture). According to Darcy’s law, an increasing pressure gradient must lead to an increase in gas flow per unit time; thus, more methane will flow to CBM wells. Additionally, the pore structure analysis described in Section 3.1 indicated that the number of the macropores and mesopores of the coal increased after ScCO2 treatment, which provides more space for gas migration and improves the absolute permeability of coal seams to some extent. Furthermore, Liu et al. (2018) demonstrated that CBM production was mainly controlled by diffusion after six days of extraction when the diffusion coefficient was 1 × 10–12 m2/s [49]. Therefore, a higher diffusion coefficient has a positive effect on CBM production. In short, CO2-ECBM could overcome the negative effects of low permeability in tectonic coal reservoirs to some extent, thereby promoting CBM development.

4. Conclusions

This study investigated changes in the pore structures of intact coal and tectonic coal after ScCO2 treatment using mercury intrusion porosimetry and N2 (77 K) adsorption, and determined the effects of these changes on the coal’s gas desorption and diffusion properties. The changes in desorption and diffusion were used to study the implications for enhanced coalbed methane recovery with CO2 sequestration in tectonic coal reservoirs. The experimental results suggest the following conclusions:
(1)
Compared with intact coal, the macropores and mesopores in tectonic coal were obviously larger after ScCO2 treatment. Additional, the TPV, SSA, and pore connectivity of treated tectonic coal were significantly improved. Pore structure analysis showed that tectonic coal was significantly affected by ScCO2 treatment. This was because tectonic coal contained more minerals and the mobilizing effect in tectonic coal was more pronounced.
(2)
The results of the methane desorption experiment showed that the desorption capacity of intact coal and tectonic coal was improved to a certain extent by ScCO2 treatment; however, the diffusion coefficient of the treated tectonic coal increased twice as much as that of intact coal. This change was consistent with the pore structure experimental results. The enhancement of the tectonic coal’s diffusion capacity after ScCO2 treatment can partially overcome the limitation imposed on tectonic coal reservoir CBM development by the coal’s inherent low permeability. The results of this study may provide new insights into CO2-ECBM in tectonic coal reservoirs.

Author Contributions

E.S. and Y.L. conceived and designed the experiments. E.S. and L.L. performed the experiments. E.S. and Q.Z. analyzed the data. E.S. and F.N. wrote the paper.

Funding

This work was financially supported by the National Science and Technology Major Project of China (Grant No. 2016ZX05043005 and 2016ZX05045004), the State Key Research Development Program of China (Grant No. 2016YFC0801404 and 2016YFC0801402), and the National Natural Science Foundation of China (51674050 and 51704046), which are gratefully acknowledged.

Acknowledgments

The authors thank the editor and anonymous reviewers for their valuable advice. The authors thank David Frishman, PhD, from Liwen Bianji, Edanz Group China (www.liwenbianji.cn/ac), for editing the English text of a draft of this manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zou, Q.; Lin, B.; Zheng, C.; Hao, Z.; Zhai, C.; Liu, T.; Liang, J.; Yan, F.; Yang, W.; Zhu, C. Novel integrated techniques of drilling-slotting-separation-sealing for enhanced coal bed methane recovery in underground coal mines. J. Nat. Gas Sci. Eng. 2015, 26, 960–973. [Google Scholar] [CrossRef]
  2. Gale, W. A review of energy associated with coal bursts. Int. J. Min. Sci. Technol. 2018, 28, 755–761. [Google Scholar] [CrossRef]
  3. Yu, G.; Zhai, C.; Qin, L.; Tang, Z.; Wu, S.; Xu, J. Changes to coal pores by ultrasonic wave excitation of different powers. J. China Univ. Min. Technol. 2018, 47, 264–270. [Google Scholar]
  4. Liu, C.; Li, S.; Yang, S. Gas emission quantity prediction and drainage technology of steeply inclined and extremely thick coal seams. Int. J. Min. Sci. Technol. 2018, 28, 415–422. [Google Scholar]
  5. Liang, B.; Jia, L.; Sun, W.; Jiang, Y. Experimental on the law of coal deformation and permeability under desorption and seepage. J. China Univ. Min. Technol. 2018, 47, 935–941. [Google Scholar]
  6. Chang, K.; Tian, H. Technical scheme and application of pressure-relief gas extraction in multi-coal seam mining region. Int. J. Min. Sci. Technol. 2018, 28, 483–489. [Google Scholar]
  7. Zou, Q.; Lin, B. Fluid−solid coupling characteristics of gas-bearing coal subjected to hydraulic slotting: An experimental investigation. Energy Fuels 2018, 32, 1047–1060. [Google Scholar] [CrossRef]
  8. 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]
  9. Haszeldine, R.S. Carbon capture and storage: How green can black be? Science 2009, 325, 1647–1652. [Google Scholar] [CrossRef]
  10. Zhang, X.; Ranjith, P.G.; Perera, M.S.A. Gas transportation and enhanced coalbed methane recovery processes in deep coal seams: A review. Energy Fuels 2016, 30, 8832–8849. [Google Scholar] [CrossRef]
  11. Gale, J. Geological storage of CO2: What do we know, where are the gaps and what more needs to be done? Energy 2004, 29, 1329–1338. [Google Scholar] [CrossRef]
  12. Perera, M. A Comprehensive Overview of CO2 Flow Behaviour in Deep Coal Seams. Energies 2018, 11, 906. [Google Scholar] [CrossRef]
  13. Fulton, P.F.; Parente, C.A.; Rogers, B.A.; Shah, N.; Reznik, A. A laboratory investigation of enhanced recovery of methane from coal by carbon dioxide injection. In SPE Unconventional Gas Recovery Symposium; Society of Petroleum Engineers: Pittsburgh, PA, USA, 1980. [Google Scholar]
  14. Reznik, A.A.; Singh, P.K.; Foley, W.L. An analysis of the effect of CO2 injection on the recovery of in-situ methane from bituminous coal: An experimental simulation. Soc. Petrol. Eng. J. 1984, 24, 521–528. [Google Scholar] [CrossRef]
  15. Jessen, K.; Tang, G.-Q.; Kovscek, A.R. Laboratory and simulation investigation of enhanced coalbed methane recovery by gas injection. Transp. Porous Media 2008, 73, 141–159. [Google Scholar] [CrossRef]
  16. Dutka, B.; Kudasik, M.; Pokryszka, Z.; Skoczylas, N.; Topolnicki, J.; Wierzbicki, M. Balance of CO2/CH4 exchange sorption in a coal briquette. Fuel Process. Technol. 2013, 106, 95–101. [Google Scholar] [CrossRef]
  17. Stevens, S.H.; Kuuskraa, V.A.; Spector, D.; Riemer, P. CO2 sequestration in deep coal seams: Pilot results and worldwide potential. In Proceedings of the Greenhouse Gas Control Technology, Saskatoon, SK, Canada, 4–6 October 1999; pp. 175–180. [Google Scholar]
  18. Span, R.; Wagner, W. A new equation of state for carbon dioxide covering the fluid region from the triple-point temperature to 1100 K at pressures up to 800 MPa. J. Phys. Chem. Ref. Data 1996, 25, 1509–1596. [Google Scholar] [CrossRef]
  19. Larsen, J.W. The effects of dissolved CO2 on coal structure and properties. Int. J. Coal Geol. 2004, 57, 63–70. [Google Scholar] [CrossRef]
  20. Zhang, K.; Cheng, Y.; Jin, K.; Guo, H.; Liu, Q.; Dong, J.; 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]
  21. Chen, R.; Qin, Y.; Wei, C.; Wang, L.; Wang, Y.; Zhang, P. Changes in pore structure of coal associated with Sc-CO2 extraction during CO2-ECBM. Appl. Sci. 2017, 7, 931. [Google Scholar] [CrossRef]
  22. Liu, C.J.; Wang, G.X.; Sang, S.X.; Rudolph, V. Changes in pore structure of anthracite coal associated with CO2 sequestration process. Fuel 2010, 89, 2665–2672. [Google Scholar] [CrossRef]
  23. Kolak, J.J.; Hackley, P.C.; Ruppert, L.F.; Warwick, P.D.; Burruss, R.C. Using ground and intact coal samples to evaluate hydrocarbon fate during supercritical CO2 injection into coal beds: Effects of particle size and coal moisture. Energy Fuels 2015, 29, 5187–5203. [Google Scholar] [CrossRef]
  24. 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]
  25. 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]
  26. Cai, Y.; Liu, D.; Pan, Z.; Yao, Y.; Li, J.; Qiu, Y. Pore structure and its impact on CH4 adsorption capacity and flow capability of bituminous and subbituminous coals from Northeast China. Fuel 2013, 103, 258–268. [Google Scholar] [CrossRef]
  27. Lu, S.Q.; Cheng, Y.P.; Li, W.; Wang, L. Pore structure and its impact on CH4 adsorption capability and diffusion characteristics of normal and deformed coals from Qinshui Basin. Int. J. Oil Gas Coal Technol. 2015, 10, 76–78. [Google Scholar] [CrossRef]
  28. Ju, Y.W.; Jiang, B.; Hou, Q.L.; Wang, G.L. The new structure-genetic classification system in tectonically deformed coals and its geological significance. J. China Coal Soc. 2004, 29, 513–517. [Google Scholar]
  29. Shi, J.Q.; Durucan, S. A bidisperse pore diffusion model for methane displacement desorption in coal by CO injection. Fuel 2003, 82, 1219–1229. [Google Scholar] [CrossRef]
  30. An, F.H.; Cheng, Y.P. An explanation of large-scale coal and gas outbursts in underground coal mines: The effect of low-permeability zones on abnormally abundant gas. Nat. Hazard Earth Syst. 2014, 14, 4751–4775. [Google Scholar] [CrossRef]
  31. Clarkson, 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]
  32. Pan, J.; Zhao, Y.; Hou, Q.; Jin, Y. Nanoscale pores in coal related to coal rank and deformation structures. Transp. Porous Media 2015, 107, 543–554. [Google Scholar] [CrossRef]
  33. Wong, S.; Law, D.; Deng, X.; Robinson, J.; Kadatz, B.; Gunter, W.D.; Ye, J.; Feng, S.; Fan, Z. Enhanced coalbed methane and CO2 storage in anthracitic coals—Micro-pilot test at South Qinshui, Shanxi, China. Int. J. Greenhouse Gas Control 2007, 1, 215–2228. [Google Scholar] [CrossRef]
  34. Washburn, E.W. The dynamics of capillary flow. Phys. Rev. Ser. 1921, 17, 273–283. [Google Scholar] [CrossRef]
  35. Barrett, E.P.; Joyner, L.G.; Halenda, P.P. The determination of pore volume and area distributions in porous substances. I. Computations from nitrogen isotherms. J. Am. Chem. Soc. 1951, 73, 373–380. [Google Scholar] [CrossRef]
  36. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  37. Landers, J.; Gor, G.Y.; Neimark, A.V. Density functional theory methods for characterization of porous materials. Colloids Surf. A 2013, 437, 3–32. [Google Scholar] [CrossRef]
  38. Hodot, B.B. Outburst of Coal and Coalbed Gas; China Industry Press: Beijing, China, 1966. [Google Scholar]
  39. Hou, Q.; Li, H.; Fan, J.; Ju, Y.; Wang, T.; Li, X.; Wu, Y. Structure and coalbed methane occurrence in tectonically deformed coals. Sci. China Earth Sci. 2012, 55, 1755–1763. [Google Scholar] [CrossRef]
  40. Li, M.; Jiang, B.; Lin, S.; Lan, F.; Wang, J. Structural controls on coalbed methane reservoirs in faer coal mine, southwest china. J. Earth Sci.-China 2013, 24, 437–448. [Google Scholar] [CrossRef]
  41. Perera, M.S.A. Influences of CO2 injection into deep coal seams: A review. Energy Fuels 2017, 31, 10324–10334. [Google Scholar] [CrossRef]
  42. Wellman, T.P.; Grigg, R.B.; Mcpherson, B.J.; Svec, R.K.; Lichtner, P.C. Evaluation of CO2–brine–reservoir rock interaction with laboratory flow tests and reactive transport modeling. In International Symposium on Oilfield Chemistry; Society of Petroleum Engineers: Pittsburgh, PA, USA, 2003. [Google Scholar]
  43. Cao, Y.; Davis, A.; Liu, R.; Liu, X.; Zhang, Y. The influence of tectonic deformation on some geochemical properties of coals—A possible indicator of outburst potential. Int. J. Coal Geol. 2003, 53, 69–79. [Google Scholar] [CrossRef]
  44. Qu, Z.; Wang, G.G.X.; Jiang, B.; Rudolph, V.; Dou, X.; Li, M. Experimental study on the porous structure and compressibility of tectonized coals. Energy Fuels 2010, 24, 2964–2973. [Google Scholar] [CrossRef]
  45. Rouquerol, F.; Rouquerol, J.; Sing, K. Adsorption by Powders and Porous Solids; Academic Press: Cambridge, MA, USA, 2014. [Google Scholar]
  46. Wang, H.; Fu, X.; Jian, K.; Li, T.; Luo, P. Changes in coal pore structure and permeability during N2 injection. J. Nat. Gas Sci. Eng. 2015, 27, 1234–1241. [Google Scholar] [CrossRef]
  47. Guo, H.; Cheng, Y.; Yuan, L.; Wang, L.; Zhou, H. Unsteady-state diffusion of gas in coals and its relationship with coal pore structure. Energy Fuels 2016, 30, 7014–7024. [Google Scholar] [CrossRef]
  48. Pan, Z.; Connell, L.D.; Camilleri, M.; Connelly, L. Effects of matrix moisture on gas diffusion and flow in coal. Fuel 2010, 89, 3207–3217. [Google Scholar] [CrossRef]
  49. Liu, Z.; Cheng, Y.; Dong, J.; Jiang, J.; Wang, L.; Li, W. Master role conversion between diffusion and seepage on coalbed methane production: Implications for adjusting suction pressure on extraction borehole. Fuel 2018, 223, 373–384. [Google Scholar] [CrossRef]
  50. Jiang, H.; Cheng, Y.; Yuan, L. A Langmuir-like desorption model for reflecting the inhomogeneous pore structure of coal and its experimental verification. RSC. Adv. 2014, 5, 2434–2440. [Google Scholar] [CrossRef]
  51. Crank, J. Mathematics of Diffusion. In Handbook of Chemistry Physic; Oxford University Press: Oxford, UK, 1975. [Google Scholar]
  52. Yang, Q.; Wang, Y. Theory of methane diffusion from coal cuttings and its application. J. China Coal Soc. 1986, 11, 87–94. [Google Scholar]
  53. Shouqing, L.; Yuanping, C.; Hongxing, Z.; Haijun, G. Gas desorption characteristics of the high-rank intact coal and fractured coal. Int. J. Min. Sci. Technol. 2015, 25, 819–825. [Google Scholar]
  54. Liu, C.; Zhou, F.; Yang, K.; Xiao, X.; Liu, Y. Failure analysis of borehole liners in soft coal seam for gas drainage. Eng. Failure Anal. 2014, 42, 274–283. [Google Scholar] [CrossRef]
Figure 1. Map showing the location of the Sihe coal mine, the mine from which the samples were collected.
Figure 1. Map showing the location of the Sihe coal mine, the mine from which the samples were collected.
Energies 11 03419 g001
Figure 2. Schematic diagram showing the set-up used for the supercritical CO2 treatments.
Figure 2. Schematic diagram showing the set-up used for the supercritical CO2 treatments.
Energies 11 03419 g002
Figure 3. Schematic diagram showing the set-up used for the gas desorption experiments.
Figure 3. Schematic diagram showing the set-up used for the gas desorption experiments.
Energies 11 03419 g003
Figure 4. The whole experimental process.
Figure 4. The whole experimental process.
Energies 11 03419 g004
Figure 5. Pore size distributions determined by mercury intrusion porosimetry (MIP) for samples of (a) intact coal and (b) tectonic coal with and without supercritical CO2 (ScCO2) treatment.
Figure 5. Pore size distributions determined by mercury intrusion porosimetry (MIP) for samples of (a) intact coal and (b) tectonic coal with and without supercritical CO2 (ScCO2) treatment.
Energies 11 03419 g005
Figure 6. Pore size distributions of the untreated and ScCO2-treated coal samples as determined by the density functional theory (DFT) analyses of N2 (77 K) adsorption isotherms for (a) intact coal and (b) tectonic coal.
Figure 6. Pore size distributions of the untreated and ScCO2-treated coal samples as determined by the density functional theory (DFT) analyses of N2 (77 K) adsorption isotherms for (a) intact coal and (b) tectonic coal.
Energies 11 03419 g006
Figure 7. Mercury intrusion porosimetry injection and ejection curves for untreated and treated coal samples: (a) intact coal; (b) tectonic coal.
Figure 7. Mercury intrusion porosimetry injection and ejection curves for untreated and treated coal samples: (a) intact coal; (b) tectonic coal.
Energies 11 03419 g007
Figure 8. Methane desorption curves for untreated and ScCO2-treated coal samples: (a) 2 MPa; (b) 4 MPa.
Figure 8. Methane desorption curves for untreated and ScCO2-treated coal samples: (a) 2 MPa; (b) 4 MPa.
Energies 11 03419 g008
Figure 9. Sketch showing methane migration in untreated and treated coal.
Figure 9. Sketch showing methane migration in untreated and treated coal.
Energies 11 03419 g009
Table 1. Maximum vitrinite reflectances, Langmuir volume and pressure, macerals, and partial proximate analyses for the coal samples used in this study.
Table 1. Maximum vitrinite reflectances, Langmuir volume and pressure, macerals, and partial proximate analyses for the coal samples used in this study.
SamplesTypeR0, max (%)Langmuir VolumeLangmuir PressureProximate Analysis (wt.%)Macerals (vol.%)
Mad (%)Aad (%)Vdaf (%)V (%)I (%)M (%)
Intact coalAnthracite3.1345.830.811.3514.268.4679.7017.223.08
Tectonic coal3.2648.770.891.5817.098.2476.4216.726.86
R0, max, maximum vitrinite reflectance; Mad, moisture content; Aad, ash yield; Vdaf, volatile matter. The subscript “ad” stands for air-dried basis. V, vitrinite; I, inertinite; M, mineral.
Table 2. Mercury intrusion porosimetry porosities and pore volume distributions for supercritical CO2 (ScCO2)-treated and untreated coal samples.
Table 2. Mercury intrusion porosimetry porosities and pore volume distributions for supercritical CO2 (ScCO2)-treated and untreated coal samples.
SamplePorosity (%)Pore Volume Distribution (%)
V1/VtV2/VtV3/VtV4/Vt
Intact coal, untreated4.1827.7356.3612.273.64
Intact coal, treated4.9720.2358.4016.035.34
Tectonic coal, untreated5.2319.6346.6324.858.90
Tectonic coal, treated6.4419.4344.5524.8811.14
V1 = pore volume of micropores; V2 = pore volume of transition pores; V3 = pore volume of mesopores; V4 = pore volume of macropores; Vt = total pore volume.
Table 3. Total pore volumes and specific surface areas for untreated and ScCO2-treated intact and tectonic coal samples.
Table 3. Total pore volumes and specific surface areas for untreated and ScCO2-treated intact and tectonic coal samples.
SampleBJH-TPV (mL/g)BET-SSA (m2/g)MIP-TPV (mL/g)MIP-SSA (m2/g)
Intact coal, untreated0.0153.4210.02200.165
Intact coal, treated0.0163.8620.02620.190
Tectonic coal, untreated0.0193.9640.03260.214
Tectonic coal, treated0.0214.4020.04220.256
BET, Brunauer-Emmett-Teller; BJH, Barrett-Joyner-Halenda; SSA, specific surface area; TPV, total pore volume. The BJH-TPV and BET-SSA values are from N2 (77 K) adsorption determinations.
Table 4. Gas diffusion coefficients for supercritical CO2-treated and untreated tectonic and intact coals.
Table 4. Gas diffusion coefficients for supercritical CO2-treated and untreated tectonic and intact coals.
SamplesIntact CoalTectonic CoalPressure
UntreatedTreatedChangeUntreatedTreatedChange
D (×10−12 m2/s)2.63603.848045.98%3.99917.502087.59%2 MPa
4.21556.009142.55%5.29129.884886.82%4 MPa

Share and Cite

MDPI and ACS Style

Su, E.; Liang, Y.; Li, L.; Zou, Q.; Niu, F. Laboratory Study on Changes in the Pore Structures and Gas Desorption Properties of Intact and Tectonic Coals after Supercritical CO2 Treatment: Implications for Coalbed Methane Recovery. Energies 2018, 11, 3419. https://doi.org/10.3390/en11123419

AMA Style

Su E, Liang Y, Li L, Zou Q, Niu F. Laboratory Study on Changes in the Pore Structures and Gas Desorption Properties of Intact and Tectonic Coals after Supercritical CO2 Treatment: Implications for Coalbed Methane Recovery. Energies. 2018; 11(12):3419. https://doi.org/10.3390/en11123419

Chicago/Turabian Style

Su, Erlei, Yunpei Liang, Lei Li, Quanle Zou, and Fanfan Niu. 2018. "Laboratory Study on Changes in the Pore Structures and Gas Desorption Properties of Intact and Tectonic Coals after Supercritical CO2 Treatment: Implications for Coalbed Methane Recovery" Energies 11, no. 12: 3419. https://doi.org/10.3390/en11123419

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