Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2027 word(s) 2027 2022-01-17 08:48:57 |
2 format correct Meta information modification 2027 2022-01-26 09:42:08 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Tambaria, T. Adsorption Factors in Enhanced Coal Bed Methane Recovery. Encyclopedia. Available online: https://encyclopedia.pub/entry/18823 (accessed on 21 May 2024).
Tambaria T. Adsorption Factors in Enhanced Coal Bed Methane Recovery. Encyclopedia. Available at: https://encyclopedia.pub/entry/18823. Accessed May 21, 2024.
Tambaria, Theodora. "Adsorption Factors in Enhanced Coal Bed Methane Recovery" Encyclopedia, https://encyclopedia.pub/entry/18823 (accessed May 21, 2024).
Tambaria, T. (2022, January 26). Adsorption Factors in Enhanced Coal Bed Methane Recovery. In Encyclopedia. https://encyclopedia.pub/entry/18823
Tambaria, Theodora. "Adsorption Factors in Enhanced Coal Bed Methane Recovery." Encyclopedia. Web. 26 January, 2022.
Adsorption Factors in Enhanced Coal Bed Methane Recovery
Edit

Enhanced coal bed methane recovery using gas injection can provide increased methane extraction depending on the characteristics of the coal and the gas that is used. Accurate prediction of the extent of gas adsorption by coal are therefore important. Both experimental methods and modeling have been used to assess gas adsorption and its effects, including volumetric and gravimetric techniques, as well as the Ono–Kondo model and other numerical simulations. Thermodynamic parameters may be used to model adsorption on coal surfaces while adsorption isotherms can be used to predict adsorption on coal pores. In addition, density functional theory and grand canonical Monte Carlo methods may be employed. 

ECBM gas adsorption coal

1. Introduction

Coal bed methane (CBM) has been extracted from coal seams for many years [1]. This methane is held in micropores [2] and so numerous methods have been developed based on gas injection techniques to remove the maximum possible amount of methane from these micropores. CBM relied on the natural pressure in the coal bed, but these methods were unable to achieve complete extraction [3]. Since then, other technologies have been developed, including the use of polymers, water injection, and proppant injection [4][5][6][7][8][9][10]. Among the various method, gas injection has been found to provide maximum methane recovery of up to 90% [11]. The gases commonly used in enhanced CBM (ECBM) recovery are carbon dioxide (CO2) and nitrogen (N2) or and mixture of the two [12][13][14][15].
The extraction process comprises gas injection into the CBM reservoir, followed by the selective adsorption of the gas on the coal surfaces and in the coal pores, methane desorption from the coal matrix, and methane flow along fractures in the bed based on Darcy’s Law [16][17][18][19][20] (Figure 1). Both the gas injection and adsorption rates in the coal bed are critical because these factors affect the coal structure and thus the extent of methane recovery [18][21][22][23]. Based on accurate adsorption analyses, including factors such as bed swelling and permeability, the effectiveness of ECBM extraction can be predicted [18][24]. Accurate predictions of gas adsorption must also take into account the possibility of sequestration of the injected gas [17]. Despite that the publication of many research and review articles on the subject of gas adsorption on coal, the adsorption of gases by coal beds based on actual coal pore morphologies and chemical structures poorly understood [25][26].
Figure 1. The ECBM extraction process based on gas injection. Adapted from [16][17][18][19][20].

2. Methane in Coal

Methane is present in coal beds both as an adsorbed gas (accounting for 80–90% of the entire methane content in a coal seam) and a free gas [16]. The latter can be compressed in pore spaces, condensed as a solid or liquid, dissolved in the coal structure or adsorbed on surfaces [2]. Coal contains methane gas because of biogenic and thermogenic processes [27] that occur during coalification with resultant storage of the gas in the coal seams [19].
Biogenic methane produced by bacterial activity at shallow to moderate depths (<500 m) [2][28][29] begins with fragmentation of the coal macromolecules via two main processes; exfoliation and/or anaerobic oxidation [30]. The biogenic processes begin with oxygen consumption after which biologically-generated CO2 is converted to methane [29]. The anaerobic oxidation reactions are promoted by various bacterial species capable of oxidizing aromatic and aliphatic structures to CO2 [30]. The majority of the biogenic methane and CO2 generated in this manner are most likely dissolved in water and removed from the system during compaction and coalification [31].
The thermogenic formation of CBM results from kerogen or the cracking of heavier hydrocarbons and increases with depth [32]. Thermogenic processes that occur in deep coal [33] at higher pressures increase the coverage of the coal surface by the CBM and result in stronger interactions between adsorbate molecules [34]. Although the composition of coal bed gases does not have a strong relationship with either coal rank or depth, thermogenic generation usually begins in highly volatile bituminous rank coal and increases with rank [32]. The thermogenic processes cause coals with higher ranks to have greater holding capacities such that they retain more gas, and also yield micropores that act as methane reservoir [30][35][36][37].
CBM can be produced at almost any time during the coal life cycle based on methanogenic bacterial growth in response to heating if the coal is uplifted and favorable subsurface environmental conditions are restored [30]. Secondary biogenic gases are also generated through bacterial metabolic activity based in the introduction of bacteria by meteoric waters migrating through permeable coal beds [31]. The biological methane in coal be continuously produced, although thermogenic gases tends to result in higher total gas contents in coal beds compared with pure biogenic-derived gases [30].

3. Gas Adsorption Characteristic of Coal

Gas adsorption on coal is influenced by the specific characteristics of the coal. Research has confirmed the effects of the coal condition and the type of coal, as well as the moisture content, ash yield, maceral content and coal pore distribution on the efficiency of ECBM extraction.

3.1. Effects of Sample Condition

The sample aspect that has most frequently been shown to affect coal gas adsorption tests is particle size. Specifically bulk samples adsorb gases more slowly than crushed coal samples [38][39]. The crushed coal used for adsorption analyses is typically in the size range of 100–60 mesh [38][40][41][42] while bulk coal specimen are usually approximately 2 cm cubes [40]. The crushed coal has a higher diffusivity and requires a shorter measurement time to achieve equilibrium compared with coal blocks [39][43][44].
Even so, crushed coal has several detrimental effects on adsorption. As an example, this material will have a damaged pore network in which closed pores have been opened. Therefore, the sample surface area will have been increased so that the adsorption capacity is artificially improved compared with the original state [38]. The crushing of coal also decreases the moisture level and increases the amount of adsorbed gas [40]. When crushed coal is used, it is nearly impossible to observe coal shrinkage or swelling because of adsorption, in contrast to trials using solid coal [45].

3.2. Moisture Effects

Moisture is an important factor in adsorption because water molecules are highly polar [46], and so can modify, the gas adsorption kinetics, mechanisms and capacity [43][46][47][48][49][50]. A comparison of adsorption during ECBM extraction trial using moist and dry coals has shown that dry conditions provide the highest gas adsorption capacity and saturation values [51]. This occurs because the adsorption sites that were originally occupied by moisture become available for methane adsorption [49][52][53][54].
Dry coal has greater coal gas adsorption capacity but can yield a large correction factor because coal in the field contains natural moisture [47][55]. Natural (or inherent) moisture affects the methane adsorption capacity differently for each coal rank. Specifically, low-rank coals exhibit greater capacities for water retention medium-rank coals show pore-blockage and fewer micropores as a result of water adsorption which limits the gas adsorption capacities and high-rank coals contain numerous in micropores that provide sufficient pore space for the exchange of water and methane [49].

3.3. Ash Yield Effects

Ash yield is attributed to pore infilling, blockage cleats and fracture systems resulting from extraneous mineral matter (such as clays and carbonates) in coal [56]. The adsorption capacity of coal is decreased with increases in the mineral ash content [57] because this material reduces the storage capacity [53] and blocks gas migration [38]. The presence of mineral matter indicates that the increasing pore volume, especially in open pores and macropores, such that gas adsorption is inhibited at faster desorption rates [2][58][59]. For these reasons, coal having a high ash yield is generally not suitable for ECBM recovery with gas injection because it cannot absorb the injected gas or requires the application of high pressures and temperatures for adsorption [60].

3.4. Maceral Effects

In coal, organic matter is known as the maceral component and this material affects gas adsorption and absorption [41]. Generally, the feasibility of performing ECBM recovery is based on assessing the vitrinite content of coal [17].
Vitrinite is a type of maceral that affects the pore structure of the coals [61], especially the coal micropores and pore distribution [62]. A higher vitrinite content leads to a higher void volume [63], greater specific surface area (SSA) [45], increased adsorption capacity [56], and decreased desorption rate [2][53][64]. Coal that is rich in vitrinite also reacts more effectively to CO2 injection and undergoes swelling [65][66].
Liptinite is another type of maceral that affects the mesopores in coal [62]. By encouraging surface diffusion, liptinite can promote the adsorption of CO2 and also act as a medium for gas transport by adsorbing CO2 while acting as a catalyst [65]. Inertinite differs from vitrinite and liptinite that it contains more macropores and fewer micropores [67]. As a result of the dominance of macropores in this material, liptinite lower the apparent surface area of the coal [62], resulting in a shorter time being required to achieve equilibrium [68], and producing significant swelling upon CO2 injection [66].

3.5. Coal Pore Effects

Methane in coal is stored on the walls of micropore networks [2] an various methods are used to understand the manner in which gases can be extracted from these micropores. ECBM recovery research has demonstrated that these pores modify the adsorption and flow of gases that are injected into coal or other porous media [69][70].
The pore volume in coal is determined by its thermal maturity [71] such that increasing maturity increases the adsorption capacity [2]. Figure 2 presents pore size distribution curves for coals from low rank to high rank as obtained from nuclear magnetic resonance (NMR) analyses [72].
Figure 2. Pore distributions in different coal ranks as determined using an NMR method. Modified from [72].
Low-rank coal contains primary epigenetic pores having irregular shapes and poor connectivity. Although the dehydration of lignite to low-rank coal reduces the moisture and oxygen-to-carbon ratio of the material [36], low-rank coal exhibit a high degree of porosity and low pore compressibility [14].
Metamorphism changes pores into circular, oval or slit morphologies [73][74] and medium-rank coal contains pore sizes ranging from macropores to micropores [75]. Medium-rank coal with a high proportion of micropores is the most suitable for industrial methane production [73].
High-rank coals contain primarily micropores with limited pore connectivity as a result the coalification process [49][62]. Coalification also leads to smaller coal pore, larger surface areas, a greater number of micropores and a higher methane content [52][76][77]. Although increases in coal rank are associated with increases in the methane content [78], the pores gradually close and form flattened structure that make gas absorption impossible [35][37][62]. As a result of the small pore surfaces, gas injection into high-rank coal must be performed at high pressures [79].

4. Gas Injection for ECBM Recovery

As noted, CO2, N2 and their mixtures are commonly used for ECBM extraction, and the injection of pure or mixed gas will lead to different adsorption effect, as explained in this section.

4.1. CO2 Injection

CO2 is an acidic gas [80] that is widely for ECBM recovery because it can extract methane with significant efficiency [12]. Coal has a high adsorption affinity for CO2 and so this gas is adsorbed rapidly, whereupon it seeps into micropores [47][64][65][81][82][83][84]. The CO2 molecule also has a small kinetic diameter and so can replace methane originally present in the micropores [12][14][85]. However, the CO2 storage capacity is affected by temperature and pressure, both of which can change the coal structure and permeability [35][43][86][87][88][89][90].

4.2. N2 Injection

N2 is used for ECBM extraction because N2 promotes methane desorption from the coal matrix [12][14][91][92]. N2 reaches equilibrium quickly, leading to a more rapid response [91][93]. N2 adsorption increases with increases in pressure, although, N2 undergoes weak interactions with adsorbents [94][95]. N2 injection also alters the coal pore structure and increases the transition pore volume such that the pore volume, pore size distribution, and connectivity are all increased [96].

4.3. Mixed Gas (CO2-N2) Injection

ECBM recovery experiments using gas mixtures have been carried out, Such mixtures have been found to be applicable to low-permeability coal [13][97] because N2 prevents expansion of the coal matrix and increases the diffusion coefficient to provide faster methane extraction [14]. However, such mixtures are not suitable for carbon sequestration [98].
The desorption of methane is enhanced in the case that the mixed gas has a CO2 concentration of less than 10%. Increasing the proportion of CO2 increases the probability of adsorption while decreasing the desorption of methane [99]. The adsorption selectivity obtainable from a mixed gas injection may be calculated as [99]
(1)

where SCO2/N2 is the adsorption selectivity and, x and y represent the mole fractions of each gas in the adsorbed and bulk phase, respectively. An adsorption selectivity of 1 indicates that N2 is adsorbed more strongly than CO2 while a value greater than 1 indicates the opposite.

References

  1. Alexis, D.A.; Karpyn, Z.T.; Ertekin, T.; Crandall, D. Fracture permeability and relative permeability of coal and their dependence on stress conditions. J. Unconv. Oil Gas Resour. 2015, 10, 1–10.
  2. Crosdale, P.J.; Beamish, B.; Valix, M. Coalbed methane sorption related to coal composition. Int. J. Coal Geol. 1998, 35, 147–158.
  3. Wang, R.; Zhang, N.; Liu, X.; Wu, X.; Chen, J.; Ma, L. Characteristics of Pore Volume Distribution and Methane Adsorption on Shales. Adsorpt. Sci. Technol. 2015, 33, 915–938.
  4. Bae, J.-S.; Bhatia, S.K.; Rudolph, V.; Massarotto, P. Pore Accessibility of Methane and Carbon Dioxide in Coals. Energy Fuels 2009, 23, 3319–3327.
  5. Lu, T.; Yu, H.; Zhou, T.; Mao, J.; Guo, B. Improvement of methane drainage in high gassy coal seam using waterjet technique. Int. J. Coal Geol. 2009, 79, 40–48.
  6. Keshavarz, A.; Badalyan, A.; Carageorgos, T.; Bedrikovetsky, P.; Johnson, R. Stimulation of coal seam permeability by micro-sized graded proppant placement using selective fluid properties. Fuel 2015, 144, 228–236.
  7. Umezaki, T.; Kawamura, T.; Okamoto, K.; Hattori, A.; Kobayashi, Y. Swelling properties and coefficient of permeability of friction-reducing polymer for pull-out of temporary sheet piles. Soils Found. 2018, 58, 797–807.
  8. Ahamed, M.; Perera, M.; Dong-Yin, L.; Ranjith, P.; Matthai, S. Proppant damage mechanisms in coal seam reservoirs during the hydraulic fracturing process: A review. Fuel 2019, 253, 615–629.
  9. Lyu, S.; Wang, S.; Chen, X.; Shah, S.; Li, R.; Xiao, Y.; Dong, Q.; Gu, Y. Experimental study of a degradable polymer drilling fluid system for coalbed methane well. J. Pet. Sci. Eng. 2019, 178, 678–690.
  10. Li, Z.; Wei, G.; Liang, R.; Shi, P.; Wen, H.; Zhou, W. LCO2-ECBM technology for preventing coal and gas outburst: Integrated effect of permeability improvement and gas displacement. Fuel 2021, 285, 119219.
  11. Zarrouk, S.J.; Moore, T. Preliminary reservoir model of enhanced coalbed methane (ECBM) in a subbituminous coal seam, Huntly Coalfield, New Zealand. Int. J. Coal Geol. 2009, 77, 153–161.
  12. Shimada, S.; Li, H.; Oshima, Y.; Adachi, K. Displacement behavior of CH4 adsorbed on coals by injecting pure CO2, N2, and CO2–N2 mixture. Environ. Earth Sci. 2005, 49, 44–52.
  13. Seomoon, H.; Lee, M.; Sung, W. Analysis of methane recovery through CO 2 –N 2 mixed gas injection considering gas diffusion phenomenon in coal seam. Energy Explor. Exploit. 2016, 34, 661–675.
  14. Oudinot, A.Y.; Riestenberg, D.E.; Koperna, G.J. Enhanced Gas Recovery and CO2 Storage in Coal Bed Methane Reservoirs with N2 Co-Injection. Energy Procedia 2017, 114, 5356–5376.
  15. Cho, S.; Kim, S.; Kim, J. Life-cycle energy, cost, and CO2 emission of CO2-enhanced coalbed methane (ECBM) recovery framework. J. Nat. Gas Sci. Eng. 2019, 70, 102953.
  16. Qi, L.; Tang, X.; Wang, Z.; Peng, X. Pore characterization of different types of coal from coal and gas outburst disaster sites using low temperature nitrogen adsorption approach. Int. J. Min. Sci. Technol. 2017, 27, 371–377.
  17. Godec, M.; Koperna, G.; Gale, J. CO2-ECBM: A Review of its Status and Global Potential. Energy Procedia 2014, 63, 5858–5869.
  18. Mukherjee, M.; Misra, S. A review of experimental research on Enhanced Coal Bed Methane (ECBM) recovery via CO2 sequestration. Earth-Sci. Rev. 2018, 179, 392–410.
  19. Harpalani, S.; Ouyang, S. A New Laboratory Technique to Estimate Gas Diffusion Characteristics of Coals. In Proceedings of the International Coalbed Methane Symposium, Tuscaloosa, AL, USA, 11–17 June 1999; pp. 141–149.
  20. Vishal, V.; Mahanta, B.; Pradhan, S.; Singh, T.; Ranjith, P. Simulation of CO2 enhanced coalbed methane recovery in Jharia coalfields, India. Energy 2018, 159, 1185–1194.
  21. Pan, Z.; Connell, L.D. A theoretical model for gas adsorption-induced coal swelling. Int. J. Coal Geol. 2007, 69, 243–252.
  22. Kim, H.J.; Shi, Y.; He, J.; Lee, H.-H.; Lee, C.-H. Adsorption characteristics of CO2 and CH4 on dry and wet coal from subcritical to supercritical conditions. Chem. Eng. J. 2011, 171, 45–53.
  23. Hol, S.; Peach, C.J.; Spiers, C.J. Applied stress reduces the CO2 sorption capacity of coal. Int. J. Coal Geol. 2011, 85, 128–142.
  24. Pan, Z.; Connell, L.D. Modelling permeability for coal reservoirs: A review of analytical models and testing data. Int. J. Coal Geol. 2012, 92, 1–44.
  25. White, C.M.; Smith, D.H.; Jones, K.L.; Goodman, A.L.; Jikich, S.A.; LaCount, R.B.; DuBose, S.B.; Ozdemir, E.; Morsi, A.B.I.; Schroeder, K.T. Sequestration of Carbon Dioxide in Coal with Enhanced Coalbed Methane Recovery—A Review. Energy Fuels 2005, 19, 659–724.
  26. Busch, A.; Gensterblum, Y. CBM and CO2-ECBM related sorption processes in coal: A review. Int. J. Coal Geol. 2011, 87, 49–71.
  27. Susilawati, R.; Esterle, J.S.; Golding, S.D.; Mares, T.E. Microbial Methane Potential for the South Sumatra Basin Coal: Formation Water Screening and Coal Substrate Bioavailability. Energy Procedia 2015, 65, 282–291.
  28. Saghafi, A. Potential for ECBM and CO2 storage in mixed gas Australian coals. Int. J. Coal Geol. 2010, 82, 240–251.
  29. Ahmed, M.; Smith, J. Biogenic methane generation in the degradation of eastern Australian Permian coals. Org. Geochem. 2001, 32, 809–816.
  30. Moore, T.A. Coalbed methane: A review. Int. J. Coal Geol. 2012, 101, 36–81.
  31. Scott, A.R.; Kaiser, W.R.; Ayers, W.B. Thermogenic and secondary biogenic gases, San Juan Basin, Colorado and New Mexico—Implications for coalbed gas producibility. Am. Assoc. Pet. Geol. Bull. 1994, 78, 1186–1209.
  32. Rice, D.D. Composition and Origins of Coalbed Gas. AAPG Bull. 1993, 77.
  33. Al-Mahmoud, M.J.; Inan, S.; Al-Duaiji, A.A. Coal occurrence in the Jurassic Dhruma Formation in Saudi Arabia: Inferences on its gas and surface mining potential. Int. J. Coal Geol. 2014, 124, 5–10.
  34. Zheng, Y.; Li, Q.; Yuan, C.; Tao, Q.; Zhao, Y.; Zhang, G.; Liu, J. Influence of temperature on adsorption selectivity: Coal-based activated carbon for CH4 enrichment from coal mine methane. Powder Technol. 2019, 347, 42–49.
  35. Merkel, A.; Gensterblum, Y.; Krooss, B.; Amann-Hildenbrand, A. Competitive sorption of CH4, CO2 and H2O on natural coals of different rank. Int. J. Coal Geol. 2015, 150–151, 181–192.
  36. Levine, J.R. Coalification: The Evolution of Coal as a Source Rock and Reservoir Rock for Oil and Gas. In Hydrocarbon in Coal; Law, B.E., Rice, D.D., Eds.; American Association of Petroleum Geologists Studies in Geology: Tuscaloosa, AL, USA, 1993; Volume 38, pp. 39–77.
  37. Li, Y.; Zhang, C.; Tang, D.; Gan, Q.; Niu, X.; Wang, K.; Shen, R. Coal pore size distributions controlled by the coalification process: An experimental study of coals from the Junggar, Ordos and Qinshui basins in China. Fuel 2017, 206, 352–363.
  38. Olajossy, A. Some parameters of coal methane system that cause very slow release of methane from virgin coal beds (CBM). Int. J. Min. Sci. Technol. 2017, 27, 321–326.
  39. Pone, J.D.N.; Halleck, P.M.; Mathews, J.P. Sorption Capacity and Sorption Kinetic Measurements of CO2 and CH4 in Confined and Unconfined Bituminous Coal. Energy Fuels 2009, 23, 4688–4695.
  40. Kim, D.; Seo, Y.; Kim, J.; Han, J.; Lee, Y. Experimental and Simulation Studies on Adsorption and Diffusion Characteristics of Coalbed Methane. Energies 2019, 12, 3445.
  41. Busch, A.; Krooss, B.; Gensterblum, Y.; van Bergen, F.; Pagnier, H. High-pressure adsorption of methane, carbon dioxideand their mixtures on coals with a special focus on the preferential sorption behaviour. J. Geochem. Explor. 2003, 78–79, 671–674.
  42. Mastalerz, M.; Gluskoter, H.; Rupp, J. Carbon dioxide and methane sorption in high volatile bituminous coals from Indiana, USA. Int. J. Coal Geol. 2004, 60, 43–55.
  43. Battistutta, E.; van Hemert, P.; Lutynski, M.; Bruining, H.; Wolf, K.-H. Swelling and sorption experiments on methane, nitrogen and carbon dioxide on dry Selar Cornish coal. Int. J. Coal Geol. 2010, 84, 39–48.
  44. Ozdemir, E.; I Morsi, B.; Schroeder, K. CO2 adsorption capacity of argonne premium coals. Fuel 2004, 83, 1085–1094.
  45. Skoczylas, N.; Pajdak, A.; Młynarczuk, M. CO2 Adsorption–Desorption Kinetics from the Plane Sheet of Hard Coal and Associated Shrinkage of the Material. Energies 2019, 12, 4013.
  46. Švábová, M.; Weishauptová, Z.; Přibyl, O. The effect of moisture on the sorption process of CO2 on coal. Fuel 2012, 92, 187–196.
  47. Busch, A.; Gensterblum, Y.; Krooss, B.; Littke, R. Methane and carbon dioxide adsorption–diffusion experiments on coal: Upscaling and modeling. Int. J. Coal Geol. 2004, 60, 151–168.
  48. Crosdale, P.J.; Moore, T.; Mares, T. Influence of moisture content and temperature on methane adsorption isotherm analysis for coals from a low-rank, biogenically-sourced gas reservoir. Int. J. Coal Geol. 2008, 76, 166–174.
  49. Chen, M.-Y.; Cheng, Y.-P.; Li, H.-R.; Wang, L.; Jin, K.; Dong, J. Impact of inherent moisture on the methane adsorption characteristics of coals with various degrees of metamorphism. J. Nat. Gas Sci. Eng. 2018, 55, 312–320.
  50. Guo, H.; Cheng, Y.; Wang, L.; Lu, S.; Jin, K. Experimental study on the effect of moisture on low-rank coal adsorption characteristics. J. Nat. Gas Sci. Eng. 2015, 24, 245–251.
  51. Hao, D.; Zhang, L.; Li, M.; Tu, S.; Zhang, C.; Bai, Q.; Wang, C. Experimental study of the moisture content influence on CH4 adsorption and deformation characteristics of cylindrical bituminous coal core. Adsorpt. Sci. Technol. 2018, 36, 1512–1537.
  52. 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.
  53. Laxminarayana, C.; Crosdale, P.J. Role of coal type and rank on methane sorption characteristics of Bowen Basin, Australia coals. Int. J. Coal Geol. 1999, 40, 309–325.
  54. Li, J.; Li, B. Evolution features of coal matrix porosity with the variation in temperature and stress. IOP Conf. Ser. Mater. Sci. Eng. 2017, 191, 12050.
  55. Siemons, N.; Busch, A. Measurement and interpretation of supercritical CO2 sorption on various coals. Int. J. Coal Geol. 2007, 69, 229–242.
  56. Kumar, H.; Mishra, M.K.; Mishra, S. Sorption capacity of Indian coal and its variation with rank parameters. J. Pet. Explor. Prod. Technol. 2019, 9, 2175–2184.
  57. Yalçin, E.; Durucan, Ş. Methane capacities of Zonguldak coals and the factors affecting methane adsorption. Min. Sci. Technol. 1991, 13, 215–222.
  58. Faiz, M.M.; Aziz, N.I.; Hutton, A.C.; Jones, B.G. Porosity and gas sorption capacity of some eastern Australian coals in relation to coal rank and composition. Coalbed Methane Symp. 1992, 19, 9–13.
  59. Karayiğit, A.I.; Mastalerz, M.; Oskay, R.G.; Buzkan, I. Bituminous coal seams from underground mines in the Zonguldak Basin (NW Turkey): Insights from mineralogy, coal petrography, Rock-Eval pyrolysis, and meso-and microporosity. Int. J. Coal Geol. 2018, 199, 91–112.
  60. Fitzgerald, J.; Pan, Z.; Sudibandriyo, M.; Robinson, J.R.; Gasem, K.; Reeves, S. Adsorption of methane, nitrogen, carbon dioxide and their mixtures on wet Tiffany coal. Fuel 2005, 84, 2351–2363.
  61. Bustin, R.; Clarkson, C. Geological controls on coalbed methane reservoir capacity and gas content. Int. J. Coal Geol. 1998, 38, 3–26.
  62. Shen, J.; Qin, Y.; Zhao, J. Maceral Contribution to Pore Size Distribution in Anthracite in the South Qinshui Basin. Energy Fuels 2019, 33, 7234–7243.
  63. Rodrigues, C.F.A.; de Sousa, M.J.L. The measurement of coal porosity with different gases. Int. J. Coal Geol. 2002, 48, 245–251.
  64. Beamish, B.; Crosdale, P.J. Instantaneous outbursts in underground coal mines: An overview and association with coal type. Int. J. Coal Geol. 1998, 35, 27–55.
  65. Karacan, C.; Mitchell, G.D. Behavior and effect of different coal microlithotypes during gas transport for carbon dioxide sequestration into coal seams. Int. J. Coal Geol. 2003, 53, 201–217.
  66. Larsen, J.W. The effects of dissolved CO2 on coal structure and properties. Int. J. Coal Geol. 2004, 57, 63–70.
  67. Unsworth, J.F.; Fowler, C.S.; Jones, L.F. Moisture in coal: 2. Maceral effects on pore structure. Fuel 1989, 68, 18–26.
  68. Keshavarz, A.; Sakurovs, R.; Grigore, M.; Sayyafzadeh, M. Effect of maceral composition and coal rank on gas diffusion in Australian coals. Int. J. Coal Geol. 2017, 173, 65–75.
  69. Brandani, S.; Mangano, E.; Sarkisov, L. Net, excess and absolute adsorption and adsorption of helium. Adsorption 2016, 22, 261–276.
  70. Zhou, Y.; Zhang, R.; Huang, J.; Li, Z.; Zhao, Z.; Zeng, Z. Effects of pore structure and methane adsorption in coal with alkaline treatment. Fuel 2019, 254, 115600.
  71. Bakshi, T.; Prusty, B.; Pathak, K.; Nayak, B.; Mani, D.; Pal, S. Source rock characteristics and pore characterization of Indian shale. J. Nat. Gas Sci. Eng. 2017, 45, 761–770.
  72. Qin, L.; Li, S.; Zhai, C.; Lin, H.; Zhao, P.; Yan, M.; Ding, Y.; Shi, Y. Joint analysis of pores in low, intermediate, and high rank coals using mercury intrusion, nitrogen adsorption, and nuclear magnetic resonance. Powder Technol. 2020, 362, 615–627.
  73. Li, S.; Tang, D.; Xu, H.; Yang, Z. The pore-fracture system properties of coalbed methane reservoirs in the Panguan Syncline, Guizhou, China. Geosci. Front. 2012, 3, 853–862.
  74. Li, Z.; Liu, D.; Cai, Y.; Wang, Y.; Teng, J. Adsorption pore structure and its fractal characteristics of coals by N2 adsorption/desorption and FESEM image analyses. Fuel 2019, 257, 116031.
  75. Clarkson, C.; Bustin, R. 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.
  76. Parkash, S.; Chakrabartty, S. Microporosity in Alberta Plains coals. Int. J. Coal Geol. 1986, 6, 55–70.
  77. Sun, W.; Feng, Y.; Jiang, C.; Chu, W. Fractal characterization and methane adsorption features of coal particles taken from shallow and deep coalmine layers. Fuel 2015, 155, 7–13.
  78. Yao, Y.; Liu, D. Effects of igneous intrusions on coal petrology, pore-fracture and coalbed methane characteristics in Hongyang, Handan and Huaibei coalfields, North China. Int. J. Coal Geol. 2012, 96–97, 72–81.
  79. Staib, G.; Sakurovs, R.; Gray, E.M.A. A pressure and concentration dependence of CO2 diffusion in two Australian bituminous coals. Int. J. Coal Geol. 2013, 116–117, 106–116.
  80. Zhou, Y.; Li, Z.; Zhang, R.; Wang, G.; Yu, H.; Sun, G.; Chen, L. CO2 injection in coal: Advantages and influences of temperature and pressure. Fuel 2018, 236, 493–500.
  81. Yamazaki, T.; Aso, K.; Chinju, J. Japanese potential of CO2 sequestration in coal seams. Appl. Energy 2006, 83, 911–920.
  82. Busch, A.; Gensterblum, Y.; Krooss, B.M. High-Pressure Sorption of Nitrogen, Carbon Dioxide, and their Mixtures on Argonne Premium Coals. Energy Fuels 2007, 21, 1640–1645.
  83. Cui, X.; Bustin, R.; Dipple, G. Selective transport of CO2, CH4, and N2 in coals: Insights from modeling of experimental gas adsorption data. Fuel 2004, 83, 293–303.
  84. Zheng, G.; Pan, Z.; Tang, S.; Ling, B.; Lv, D.; Connell, L.D. Laboratory and Modeling Study on Gas Diffusion with Pore Structures in Different-Rank Chinese Coals. Energy Explor. Exploit. 2013, 31, 859–877.
  85. Bhowmik, S.; Dutta, P. Adsorption rate characteristics of methane and CO2 in coal samples from Raniganj and Jharia coalfields of India. Int. J. Coal Geol. 2013, 113, 50–59.
  86. Charrière, D.; Pokryszka, Z.; Behra, P. 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.
  87. Zhao, J.; Tang, D.; Qin, Y.; Xu, H.; Liu, Y.; Wu, H. Characteristics of Methane (CH4) Diffusion in Coal and Its Influencing Factors in the Qinshui and Ordos Basins. Energy Fuels 2018, 32, 1196–1205.
  88. Li, X.; Yan, X.; Kang, Y. Effect of temperature on the permeability of gas adsorbed coal under triaxial stress conditions. J. Geophys. Eng. 2017, 15, 386–396.
  89. Fang, H.; Sang, S.; Liu, S. The coupling mechanism of the thermal-hydraulic-mechanical fields in CH4-bearing coal and its application in the CO2-enhanced coalbed methane recovery. J. Pet. Sci. Eng. 2019, 181, 106177.
  90. Wang, X.; Zhang, D.; Su, E.; Jiang, Z.; Wang, C.; Chu, Y.; Ye, C. Pore structure and diffusion characteristics of intact and tectonic coals: Implications for selection of CO2 geological sequestration site. J. Nat. Gas Sci. Eng. 2020, 81, 103388.
  91. Dutta, A. Multicomponent Gas Diffusion and Adsorption in Coals for Enhanced Methane Recovery. Master’s Thesis, Stanford University, Stanford, USA, June 2009.
  92. Abunowara, M.; Sufian, S.; Bustam, M.A.; Eldemerdash, U.; Suleman, H.; Bencini, R.; Assiri, M.A.; Ullah, S.; Al-Sehemi, A.G. Experimental measurements of carbon dioxide, methane and nitrogen high-pressure adsorption properties onto Malaysian coals under various conditions. Energy 2020, 210, 118575.
  93. George, J.S.; Barakat, M. The change in effective stress associated with shrinkage from gas desorption in coal. Int. J. Coal Geol. 2001, 45, 105–113.
  94. Joewondo, N. Pore Sturucture of Micro and Mesoporous Mudrocks Based on Nitrogen and Carbon Sorption. Master’s Thesis, Colorado School of Mines, Golden, Colorado, 2018.
  95. Aleghafouri, A.; Mohsen-Nia, M.; Mohajeri, A.; Mahdyarfar, M.; Asghari, M. Micropore Size Analysis of Activated Carbons Using Nitrogen, Carbon Dioxide and Methane Adsorption Isotherms: Experimental and Theoretical Studies. Adsorpt. Sci. Technol. 2012, 30, 307–316.
  96. Wang, H.; Fu, X.; Jian, K.; Li, T.; Luo, P. Changes in coal pore structure and permeability during N 2 injection. J. Nat. Gas Sci. Eng. 2015, 27, 1234–1241.
  97. Fang, Z.; Li, X.; Hu, H. Gas mixture enhance coalbed methane recovery technology: Pilot tests. Energy Procedia 2011, 4, 2144–2149.
  98. Reeves, S. The Coal-Seq Project: Key Results from Field, Laboratory, and Modeling Studies. Greenh. Gas Control Technol. 2005, II, 1399–1403.
  99. Zhu, H.; He, X.; Xie, Y.; Guo, S.; Huo, Y.; Wang, W. A Study on the Effect of Coal Metamorphism on the Adsorption Characteristics of a Binary Component System: CO2 and N2. ACS Omega 2020, 6, 523–532.
More
Information
Subjects: Energy & Fuels
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 409
Revisions: 2 times (View History)
Update Date: 26 Jan 2022
1000/1000
Video Production Service