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 -- 1421 2023-09-21 07:20:50 |
2 format change -91 word(s) 1330 2023-09-21 08:46:43 |

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.
Adsul, T.; Ghosh, S.; Kumar, S.; Tiwari, B.; Dutta, S.; Varma, A.K. Methanogenesis. Encyclopedia. Available online: https://encyclopedia.pub/entry/49449 (accessed on 14 May 2024).
Adsul T, Ghosh S, Kumar S, Tiwari B, Dutta S, Varma AK. Methanogenesis. Encyclopedia. Available at: https://encyclopedia.pub/entry/49449. Accessed May 14, 2024.
Adsul, Tushar, Santanu Ghosh, Susheel Kumar, Balram Tiwari, Subir Dutta, Atul Kumar Varma. "Methanogenesis" Encyclopedia, https://encyclopedia.pub/entry/49449 (accessed May 14, 2024).
Adsul, T., Ghosh, S., Kumar, S., Tiwari, B., Dutta, S., & Varma, A.K. (2023, September 21). Methanogenesis. In Encyclopedia. https://encyclopedia.pub/entry/49449
Adsul, Tushar, et al. "Methanogenesis." Encyclopedia. Web. 21 September, 2023.
Methanogenesis
Edit

Coal bed methane (CBM) extraction has astounding effects on the global energy budget. Since the earliest discoveries of CBM, this natural gas form has witnessed ever-increasing demands from the core sectors of the economy. CBM is an unconventional source of energy occurring naturally within coal beds. The multiphase CBM generation during coal evolution commences with microbial diagenesis of the sedimentary organic matter during peatification, followed by early to mature thermogenic kerogen decomposition and post-coalification occurrences. Indeed, the origin of the CBM and, moreover, its economically valuable retention within coal seams is a function of various parameters. Several noticeable knowledge gaps include the controls of coal make-up and its physicochemical position on the CBM generation and genetic link through fossil molecular and stable isotopic integration with the parent coal during its evolution.

methanogenesis microbially enhanced–coalbed methane molecular signals

1. Origin of Methane in Coal Beds

The origin of coal bed methane (CBM) can be distinguished broadly into (a) primary microbial (PM), (b) thermogenic (T), and (c) mixed sources [1][2][3][4][5][6][7]. Meanwhile, microbial methane is present primarily in low-rank coals, while high-rank coals often comprise secondary late mature microbial gas [2][6][7][8][9][10][11][12]. Thermogenic gas can also be divided into early mature thermogenic gas (EMT) and late mature thermogenic gas (LMT) [9] based on its mode of origin. The origin of CBM can be evidenced by (a) the gas molecular ratio (methane/(ethane + propane): CH4/(C2H6 + C3H8)); (b) the stable carbon and deuterium isotopes of methane (δ13C–CH4 and δD–CH4); (c) the stable carbon isotope of carbon dioxide (δ13C–CO2); (d) the stable deuterium and oxygen isotopes of coal seam water (δD–H2O and δ18O– H2O); (e) the stable carbon isotopic discrimination factor (αCO2− CH4); and (f) the carbon isotopic difference between CO2 and CH413CCO2−H4), [2][3][4][7][9][11][13][14][15][16].

2. Microbial Methanogenesis: An Overview

Microbial methanogenesis is the ultimate phase of organic matter biodegradation, which yields methane and carbon dioxide [5]. The microbial decomposition of organic matter involves geopolymer conversion to low-molecular-weight hydrocarbons (CLMW) within the coal beds. Microbially mediated transformations of the CLMW compounds yield methane precursors, i.e., methanol, formate, H2, acetate, carbon monoxide, etc. [12][17]. The obligate anaerobes, such as methanogenic archaea, facilitate the final stage of biodegradation and yield methane. Carbon dioxide is a subsidiary product of microbial methanogenesis that is required by the electron balance of the CLMW compounds [5][9]. The methane precursor compounds donate an electron to the methanogenic archaea in a syntrophic association in the final stage of biodegradation. However, direct electron transfer from microbes to methanogens may also occur without requiring the intermediary H2 precursor [18][19]. Methanogens, including Methanospirillum, Methanococcus, and Methanobacterium, can utilize the H2 and CO2 precursors, whereas only Methanosaeta and Methanosarcina archaea can utilize the acetate. Methanosarcina utilizes either acetate or CO2 + H2 [17]. Assessments based on the DNA of the microbial communities suggest that their archaeal diversity is less than that of the bacteria in unconventional reservoirs [20][21][22]. Proteobacteria, Bacteroidetesm, Firmicutes, and Actinobacteria are abundant in coal beds [22][23][24][25][26], and are capable of metabolizing and decomposing kerogen and hydrocarbons. Meanwhile, Methanobacteriales, Methanomicrobiales, and Methanosarcinales reflect the archaeal diversity in CBM basins [27][28].

3. Pathways of Primary Microbial Methanogenesis

Methanogenic archaea proliferate in oxygen-poor anoxic redox conditions. In addition, thermogenic constraints influence microbial methanogenesis. Competitive substrates, such as formate, acetate, and H2 [7], are scavenged by heterotrophic microbes that intercede in the iron, nitrate, and sulfate reduction pathways. These non-methanogenic pathways produce more free energy per mole of the substrate than the methanogenic route [5]. Microbial methanogenesis takes place primarily through two ways in water free of alternative electron acceptors (such as sulfate): (a) the hydrogenotrophic or CO2-reduction pathway (Equation (1)), which involves H2 as an electron donor and CO2 as an electron acceptor; and (b) acetate fermentation, or the acetoclastic route (Equation (2)), which is characterized by the reduction of a methyl group to methane and the oxidation of a carboxylic group to carbon dioxide [2][5][16][17][29].
CO2 + 4H2 = CH4 + 2H2O     (1)
CH3COOH = CH4 + CO2       (2)
The hydrogenotrophic and acetoclastic methanogenesis involve terminal enzymatic activities, and these pathways engage the methyl–coenzyme M reductase. Meanwhile, during hydrogenotrophic methanogenesis, primary activation and transfer of the C1 unit from the substrate uses the methanofuran coenzyme [17]. Further, in the acetoclastic pathway, methane production from acetate is activated by acetyl–coenzyme A. Subsequent enzymatic procedures form methyl–coenzyme M, which is further reduced to methane. The presence of transition metals, such as cobalt and nickel, also influences the enzymatic activities during methanogenesis [30][31]. These two transition metals are present as trace elements in coal beds, and, thus, may have a considerable effect on CBM production. Both of these methanogenic pathways are exogenic and yield free energy, and the gain in free energy is used to synthesize adenosine triphosphate (ATP). Oremland et al. [32] reported that the free energy yield during hydrogenotrophic methanogenesis (−135 kJ mol−1) is much higher compared to acetoclastic methanogenesis (−31 kJ mol−1). This free energy gain during hydrogenotrophic methanogenesis is sufficient to synthesize at least one molecule of ATP. The terminal enzymatic activities involving the methyl–coenzyme M reductase help to conserve energy. These enzymatic activities are crucial for preserving the methanogenic society, and also favor energetically efficient metabolisms in energetically inadequate environments [17].
Methanogenesis using methylamines and methanol by Methanolobus has also been reported in addition to the acetoclastic and hydrogenotrophic pathways. Demethoxylation produces ranges of methylated compounds, which can be utilized as non-competitive substrates [12][21][33][34]. Meanwhile, in the presence of microbially-reducible iron oxides or sulfates, methylated substrates are used by methylotrophs when bacteria surpass the methanogens in utilizing the competitive substrates. Sulfate abundance may adversely affect methanogenesis by restricting the methanogens from utilizing the competitive substrates. In the presence of sulfur, sulfate-reducing microbes efficiently scavenge the acetate, making hydrogenotrophic methanogenesis the dominant pathway in marine sediments. On the other hand, acetoclastic methanogenesis is dominant in freshwater sediments that lack sulfur [7][16][17][35]. A lack of sulfate-reducing microbes promotes the formation of short-chain-length volatile fatty acids and acetate, which offer a proper substrate for acetoclastic methanogenesis in a freshwater environment [6][7]. In addition, the methylotrophic pathway is also considered to be dominant in marine sediment due to the abundance of sulfate in seawater [36][37]. Furthermore, salinity and temperature control the growth, proliferation, and extent of biodegradation of organic matter. Methanogenic activities are hindered at temperatures > ~80 °C [38]. Further, freshwater recharge can induce methanogenesis in coal beds [10][39][40][41]. Salinity may also influence the archaea, as the energy needed to remove salts from bacterial cells alleviates the methanogenic potential [5]. Hydrogenotrophic methanogenesis can tolerate higher salinity, while acetate precursor is scavenged at lower salinity conditions by methanogens [42][43]. Meanwhile, methylotrophic methanogenesis that utilizes a methylated substrate persists at a greater level of salinity than acetoclastic and even hydrogenotrophic methanogenesis [44]. Moreover, it is also observed that hydrogenotrophic methanogenesis tolerates more salt compared to the acetoclastic pathway at low temperatures (≤30 °C), while at 60 °C, both methanogenic routes become less tolerant to salt [38].

4. Secondary Microbial Methanogenesis

Secondary microbial gases may occur in the medium- to high-rank coals and may alter the primary thermogenic methane isotopic fingerprints if mixing occurs [2][11]. The secondary microbial methanogenesis may occur either by the hydrogenotrophic or acetoclastic pathway. Weathering of coal seams, atmospheric exposure, and mixing between meteoric water and coal bed formation water may introduce microbes into the coal beds, which metabolize residual kerogen and have previously produced wet gases to generate methane. In hydrogen-rich coals or petroliferous coal reservoirs, the biodegradation of liquid hydrocarbons may also generate secondary microbial gases. The carbon dioxide formed along with these gases shows 13C enrichment and is, thus, isotopically heavier than primary microbial gases [45]. Low-rank coals often possess higher gas content than high-rank coals, which may further indicate a plausible influence of secondary microbial methanogenesis. The microbial communities introduced into coal beds through the aforementioned pathways may utilize the hydrocarbons and residual labile kerogen to yield methane. If this secondary microbial methane is present in considerable abundance, it may re–saturate the coal to adsorption isotherm [46]. The stable isotopic differences between the primary and secondary microbial methane (discussed in the following sections) were documented by Milkov and Etiope [47] in their proposed gas genetic diagrams.

References

  1. Gao, L.; Mastalerz, M.; Schimmelmann, A. The Origin of Coalbed Methane. In Coal Bed Methane: From Prospect to Pipeline; Elsevier Inc.: Amsterdam, The Netherlands, 2014; pp. 7–29.
  2. Golding, S.D.; Boreham, C.J.; Esterle, J.S. Stable isotope geochemistry of coal bed and shale gas and related production waters: A review. Int. J. Coal Geol. 2013, 120, 24–40.
  3. Hamilton, S.; Golding, S.; Baublys, K.; Esterle, J. Stable isotopic and molecular composition of desorbed coal seam gases from the Walloon Subgroup, eastern Surat Basin, Australia. Int. J. Coal Geol. 2014, 122, 21–36.
  4. Kinnon, E.; Golding, S.; Boreham, C.; Baublys, K.; Esterle, J. Stable isotope and water quality analysis of coal bed methane production waters and gases from the Bowen Basin, Australia. Int. J. Coal Geol. 2010, 82, 219–231.
  5. Vinson, D.S.; Blair, N.E.; Martini, A.M.; Larter, S.; Orem, W.H.; McIntosh, J.C. Microbial methane from in situ biodegradation of coal and shale: A review and reevaluation of hydrogen and carbon isotope signatures. Chem. Geol. 2017, 453, 128–145.
  6. Whiticar, M. Stable Isotope Geochemistry of Coals, Humic Kerogens and Related Natural Gases. Int. J. Coal. Geol. 1996, 32, 191–215.
  7. Whiticar, M.J. Carbon and hydrogen isotope systematics of bacterial formation and oxidation of methane. Chem. Geol. 1999, 161, 291–314.
  8. Clayton, J. Geochemistry of coalbed gas—A review. Int. J. Coal Geol. 1998, 35, 159–173.
  9. Dutta, S.; Ghosh, S.; Varma, A.K. Methanogenesis in the Eocene Tharad lignite deposits of Sanchor Sub-Basin, Gujarat, India: Insights from gas molecular ratio and stable carbon isotopic compositions. J. Nat. Gas Sci. Eng. 2021, 91, 103970.
  10. Faiz, M.; Hendry, P. Significance of microbial activity in Australian coal bed methane reservoirs—A review. Bull. Can. Pet. Geol. 2006, 54, 261–272.
  11. Ghosh, S.; Golding, S.D.; Varma, A.K.; Baublys, K.A. Stable isotopic composition of coal bed gas and associated formation water samples from Raniganj Basin, West Bengal, India. Int. J. Coal Geol. 2018, 191, 1–6.
  12. Strąpoć, D.; Mastalerz, M.; Dawson, K.; Macalady, J.; Callaghan, A.V.; Wawrik, B.; Turich, C.; Ashby, M. Biogeochemistry of Microbial Coal-Bed Methane. Annu. Rev. Earth Planet. Sci. 2011, 39, 617–656.
  13. Bernard, B.; Brooks, J.; Sackett, W. Natural gas seepage in the Gulf of Mexico. Earth Planet. Sci. Lett. 1976, 31, 48–54.
  14. Conrad, R. Quantification of methanogenic pathways using stable carbon isotopic signatures: A review and a proposal. Org. Geochem. 2005, 36, 739–752.
  15. Schoell, M. Multiple origins of methane in the Earth. Chem. Geol. 1988, 71, 1–10.
  16. Whiticar, M.; Faber, E.; Schoell, M. Biogenic methane formation in marine and freshwater environments: CO2 reduction vs. acetate fermentation—Isotope evidence. Geochim. Cosmochim. Acta 1986, 50, 693–709.
  17. Formolo, M. The Microbial Production of Methane and Other Volatile Hydrocarbons. In Handbook of Hydrocarbon and Lipid Microbiology; Timmis, K.N., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 113–126.
  18. Rotaru, A.-E.; Shrestha, P.M.; Liu, F.; Shrestha, M.; Shrestha, D.; Embree, M.; Zengler, K.; Wardman, C.; Nevin, K.P.; Lovley, D.R. A new model for electron flow during anaerobic digestion: Direct interspecies electron transfer to Methanosaeta for the reduction of carbon dioxide to methane. Energy Environ. Sci. 2013, 7, 408–415.
  19. Rotaru, A.-E.; Shrestha, P.M.; Liu, F.; Markovaite, B.; Chen, S.; Nevin, K.P.; Lovley, D.R. Direct Interspecies Electron Transfer between Geobacter metallireducens and Methanosarcina barkeri. Appl. Environ. Microbiol. 2014, 80, 4599–4605.
  20. Colosimo, F.; Thomas, R.; Lloyd, J.R.; Taylor, K.G.; Boothman, C.; Smith, A.D.; Lord, R.; Kalin, R.M. Biogenic methane in shale gas and coal bed methane: A review of current knowledge and gaps. Int. J. Coal Geol. 2016, 165, 106–120.
  21. Barnhart, E.P.; De León, K.B.; Ramsay, B.D.; Cunningham, A.B.; Fields, M.W. Investigation of coal-associated bacterial and archaeal populations from a diffusive microbial sampler (DMS). Int. J. Coal Geol. 2013, 115, 64–70.
  22. Green, M.S.; Flanegan, K.C.; Gilcrease, P.C. Characterization of a methanogenic consortium enriched from a coalbed methane well in the Powder River Basin, U.S.A. Int. J. Coal Geol. 2008, 76, 34–45.
  23. Jones, E.J.; Voytek, M.A.; Warwick, P.D.; Corum, M.D.; Cohn, A.; Bunnell, J.E.; Clark, A.C.; Orem, W.H. Bioassay for estimating the biogenic methane-generating potential of coal samples. Int. J. Coal Geol. 2008, 76, 138–150.
  24. Jones, E.J.P.; Voytek, M.A.; Corum, M.D.; Orem, W.H. Stimulation of Methane Generation from Nonproductive Coal by Addition of Nutrients or a Microbial Consortium. Appl. Environ. Microbiol. 2010, 76, 7013–7022.
  25. Li, D.; Hendry, P.; Faiz, M. A survey of the microbial populations in some Australian coalbed methane reservoirs. Int. J. Coal Geol. 2008, 76, 14–24.
  26. Strąpoć, D.; Mastalerz, M.; Schimmelmann, A.; Drobniak, A.; Hedges, S. Variability of geochemical properties in a microbially dominated coalbed gas system from the eastern margin of the Illinois Basin, USA. Int. J. Coal Geol. 2008, 76, 98–110.
  27. An, D.; Caffrey, S.M.; Soh, J.; Agrawal, A.; Brown, D.; Budwill, K.; Dong, X.; Dunfield, P.F.; Foght, J.; Gieg, L.M.; et al. Metagenomics of Hydrocarbon Resource Environments Indicates Aerobic Taxa and Genes to be Unexpectedly Common. Environ. Sci. Technol. 2013, 47, 10708–10717.
  28. Meslé, M.; Dromart, G.; Oger, P. Microbial methanogenesis in subsurface oil and coal. Res. Microbiol. 2013, 164, 959–972.
  29. Conrad, R.; Bak, F.; Seitz, H.J.; Thebrath, B.; Mayer, H.P.; Schütz, H. Hydrogen turnover by psychrotrophic homoacetogenic and mesophilic methanogenic bacteria in anoxic paddy soil and lake sediment. FEMS Microbiol. Lett. 1989, 62, 285–293.
  30. Hausrath, E.M.; Liermann, L.J.; House, C.H.; Ferry, J.G.; Brantley, S.L. The effect of methanogen growth on mineral substrates: Will Ni markers of methanogen-based communities be detectable in the rock record? Geobiology 2007, 5, 49–61.
  31. Kida, K.; Shigematsu, T.; Kijima, J.; Numaguchi, M.; Mochinaga, Y.; Abe, N.; Morimura, S. Influence of Ni2+ and Co2+ on methanogenic activity and the amounts of coenzymes involved in methanogenesis. J. Biosci. Bioeng. 2001, 91, 590–595.
  32. Oremland, R.S.; Whiticar, M.J.; Strohmaier, F.E.; Kiene, R.P. Bacterial ethane formation from reduced, ethylated sulfur compounds in anoxic sediments. Geochim. Cosmochim. Acta 1988, 52, 1895–1904.
  33. Doerfert, S.N.; Reichlen, M.; Iyer, P.; Wang, M.; Ferry, J.G. Methanolobus zinderi sp. nov., a methylotrophic methanogen isolated from a deep subsurface coal seam. Int. J. Syst. Evol. Microbiol. 2009, 59, 1064–1069.
  34. Wawrik, B.; Mendivelso, M.; Parisi, V.A.; Suflita, J.M.; Davidova, I.A.; Marks, C.R.; Van Nostrand, J.D.; Liang, Y.; Zhou, J.; Huizinga, B.J.; et al. Field and laboratory studies on the bioconversion of coal to methane in the San Juan Basin. FEMS Microbiol. Ecol. 2012, 81, 26–42.
  35. Hornibrook, E.R.; Longstaffe, F.J.; Fyfe, W.S. Evolution of stable carbon isotope compositions for methane and carbon dioxide in freshwater wetlands and other anaerobic environments. Geochim. Cosmochim. Acta 2000, 64, 1013–1027.
  36. King, G.M. Utilization of hydrogen, acetate, and “noncompetitive”; substrates by methanogenic bacteria in marine sediments. Geomicrobiol. J. 1984, 3, 275–306.
  37. Oremland, R.S.; Polcin, S. Methanogenesis and Sulfate Reduction: Competitive and Noncompetitive Substrates in Estuarine Sediments. Appl. Environ. Microbiol. 1982, 44, 1270–1276.
  38. Head, I.M.; Gray, N.; Larter, S. Life in the slow lane; biogeochemistry of biodegraded petroleum containing reservoirs and implications for energy recovery and carbon management. Front. Microbiol. 2014, 5, 566.
  39. Martini, A.M.; Budai, J.M.; Walter, L.M.; Schoell, M. Microbial generation of economic accumulations of methane within a shallow organic-rich shale. Nature 1996, 383, 155–158.
  40. Martini, A.; Walter, L.; Budai, J.; Ku, T.; Kaiser, C.; Schoell, M. Genetic and temporal relations between formation waters and biogenic methane: Upper Devonian Antrim Shale, Michigan Basin, USA. Geochim. Cosmochim. Acta 1998, 62, 1699–1720.
  41. McIntosh, J.; Walter, L.; Martini, A. Pleistocene recharge to midcontinent basins: Effects on salinity structure and microbial gas generation. Geochim. Cosmochim. Acta 2002, 66, 1681–1700.
  42. McIntosh, J.C.; Warwick, P.D.; Martini, A.M.; Osborn, S.G. Coupled hydrology and biogeochemistry of Paleocene–Eocene coal beds, northern Gulf of Mexico. GSA Bull. 2010, 122, 1248–1264.
  43. Schlegel, M.E.; McIntosh, J.C.; Petsch, S.T.; Orem, W.H.; Jones, E.J.; Martini, A.M. Extent and limits of biodegradation by in situ methanogenic consortia in shale and formation fluids. Appl. Geochem. 2013, 28, 172–184.
  44. Oren, A. Thermodynamic limits to microbial life at high salt concentrations. Environ. Microbiol. 2010, 13, 1908–1923.
  45. Milkov, A.V. Worldwide distribution and significance of secondary microbial methane formed during petroleum biodegradation in conventional reservoirs. Org. Geochem. 2011, 42, 184–207.
  46. Scott, A.R.; Kaiser, W.R.; Ayers, W.B., Jr. 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.
  47. Milkov, A.; Etiope, G. Geochemistry of Shale Gases from around the World. In Proceedings of the 80th EAGE Conference and Exhibition 2018, Copenhagen, Denmark, 11–14 June 2018; EAGE Publications BV: Kosterijland, Bunnik, 2018; pp. 1–5.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , ,
View Times: 309
Revisions: 2 times (View History)
Update Date: 21 Sep 2023
1000/1000