Methanogen

Not to be confused with methanotrophs.

Methanogens are microorganisms that produce methane as a metabolic byproduct in anoxic conditions. They uniquely belong to the domain of archaea. They are common in wetlands, where they are responsible for marsh gas, and in the digestive tracts of animals such as ruminants and humans, where they are responsible for the methane content of belching in ruminants and flatulence in humans.[1] In marine sediments the biological production of methane, also termed methanogenesis, is generally confined to where sulfates are depleted, below the top layers.[2] Moreover, the methanogenic archaea populations play an indispensable role in anaerobic wastewater treatments.[3] Others are extremophiles, found in environments such as hot springs and submarine hydrothermal vents as well as in the "solid" rock of the Earth's crust, kilometers below the surface.

Physical description

Methanogens are coccoid (spherical shaped) or bacilli (rod shaped). There are over 50 described species of methanogens, which do not form a monophyletic group, although all methanogens belong to Archaea. They are anaerobic organisms and cannot function under aerobic conditions. They are very sensitive to the presence of oxygen even at trace level. Usually, they cannot sustain oxygen stress for a prolonged time. However, Methanosarcina barkeri is exceptional in possessing a superoxide dismutase (SOD) enzyme, and may survive longer than the others in the presence of O2.[4][5] Some methanogens, called hydrogenotrophic, use carbon dioxide (CO2) as a source of carbon, and hydrogen as a reducing agent.

The reduction of carbon dioxide into methane in the presence of hydrogen can be expressed as follows:

CO2 + 4 H2 → CH4 + 2H2O

Some of the CO2 is reacted with the hydrogen to produce methane, which creates an electrochemical gradient across cell membrane, used to generate ATP through chemiosmosis. In contrast, plants and algae use water as their reducing agent.

Methanogens lack peptidoglycan, a polymer that is found in the cell walls of the Bacteria but not in those of Archaea. Some methanogens have a cell wall that is composed of pseudopeptidoglycan. Other methanogens do not, but have at least one paracrystalline array (S-layer) made up of proteins that fit together like a jigsaw puzzle.

Methanogens and extreme environments

Methanogens play the vital ecological role in anaerobic environments of removing excess hydrogen and fermentation products that have been produced by other forms of anaerobic respiration. Methanogens typically thrive in environments in which all electron acceptors other than CO2 (such as oxygen, nitrate, ferriciron (Fe(III)), and sulfate) have been depleted. In deep basaltic rocks near the mid ocean ridges, they can obtain their hydrogen from the serpentinisation reaction of olivine as observed in the hydrothermal field of Lost City.

The thermal breakdown of water and water radiolysis are other possible sources of hydrogen.

Methanogens are key agents of remineralization of organic carbon in continental margin sediments and other aquatic sediments with high rates of sedimentation and high sediment organic matter. Under the correct conditions of pressure and temperature, biogenic methane can accumulate in massive deposits of methane clathrates,[6] which account for a significant fraction of organic carbon in continental margin sediments and represent a key reservoir of a potent greenhouse gas.[7]

Methanogens have been found in several extreme environments on Earth - buried under kilometres of ice in Greenland and living in hot, dry desert soil. They are known to be the most common archaebacteria in deep subterranean habitats. Live microbes making methane were found in a glacial ice core sample retrieved from about three kilometres under Greenland by researchers from the University of California, Berkeley. They also found a constant metabolism able to repair macromolecular damage, at temperatures of 145 to –40°C.[8]

Another study[9] has also discovered methanogens in a harsh environment on Earth. Researchers studied dozens of soil and vapour samples from five different desert environments in Utah, Idaho and California in the United States, and in Canada and Chile. Of these, five soil samples and three vapour samples from the vicinity of the Mars Desert Research Station in Utah were found to have signs of viable methanogens.[10]

Some scientists have proposed that the presence of methane in the Martian atmosphere may be indicative of native methanogens on that planet.[11]

Closely related to the methanogens are the anaerobic methane oxidizers, which utilize methane as a substrate in conjunction with the reduction of sulfate and nitrate.[12] Most methanogens are autotrophic producers, but those that oxidize CH3COO are classed as chemotroph instead.

Comparative genomics and molecular signatures

Comparative genomic analysis has led to the identification of 31 signature proteins which are specific for the methanogens (also known as Methanoarchaeota). Most of these proteins are related to methanogenesis, and they could serve as potential molecular markers for the methanogens.[13] Additionally, 10 proteins found in all methanogens which are shared by Archaeoglobus, suggest that these two groups are related.[13] In phylogenetic trees, the methanogens are not monophyletic and they are generally split into three clades.[13][14] Hence, the unique shared presence of large numbers of proteins by all methanogens could be due to lateral gene transfers.[13]

Fermentative metabolism

Although most marine biogenic methane is the result of carbon dioxide (CO2) reduction, a small amount is derived from acetate (CH3COO) fermentation.[15]

In the fermentation pathway, acetic acid undergoes a dismutation reaction to produce methane and carbon dioxide:[16][17]

CH3COO + H+ CH4 + CO2       ΔG° = -36 kJ/reaction

This disproportionation reaction is enzymatically catalysed. One electron is transferred from the carbonyl function (e donor) of the carboxylic group to the methyl group (e acceptor) of acetic acid to respectively produce CO2 and methane gas.

Archaea that catabolize acetate for energy are referred to as acetotrophic or aceticlastic. Methylotrophic archaea utilize methylated compounds such as methylamines, methanol, and methanethiol as well.

Strains of methanogens

See also

References

  1. Joseph W. Lengeler (1999). Biology of the Prokaryotes. Stuttgart: Thieme. p. 796. ISBN 0-632-05357-7.
  2. J.K. Kristjansson; et al. (1982). "Different Ks values for hydrogen of methanogenic bacteria and sulfate-reducing bacteria: an explanation for the apparent inhibition of methanogenesis by sulfate". Arch. Microbiol. 131 (3): 278–282. doi:10.1007/BF00405893.
  3. Tabatabaei, Meisam; Rahim, Raha Abdul; Abdullah, Norhani; Wright, André-Denis G.; Shirai, Yoshihito; Sakai, Kenji; Sulaiman, Alawi; Hassan, Mohd Ali (2010). "Importance of the methanogenic archaea populations in anaerobic wastewater treatments". Process Biochemistry. 45 (8): 1214–1225. doi:10.1016/j.procbio.2010.05.017.
  4. Peters V; Conrad R (1995). "Methanogenic and other strictly anaerobic bacteria in desert soil and other oxic sois". Applied and Environmental Microbiology. 61 (4): 1673–1676. PMC 1388429Freely accessible. PMID 16535011.
  5. http://spacecenter.uark.edu/JillJabstract.doc
  6. Kvenvolden, K. (1995). "A review of the geochemistry of methane in natural gas hydrate". Organic Geochemistry. 23 (11–12): 997–1008. doi:10.1016/0146-6380(96)00002-2.
  7. Milkov, Alexei V (2004). "Global estimates of hydrate-bound gas in marine sediments: how much is really out there?". Earth-Science Reviews. 66 (3–4): 183–197. Bibcode:2004ESRv...66..183M. doi:10.1016/j.earscirev.2003.11.002.
  8. Tung, H. C.; Bramall, N. E.; Price, P. B. (2005). "Microbial origin of excess methane in glacial ice and implications for life on Mars". Proceedings of the National Academy of Sciences. 102 (51): 18292. doi:10.1073/pnas.0507601102.
  9. Icarus (vol. 178, p. 277)cs:Methanogen
  10. Extreme bugs back idea of life on Mars
  11. "Crater Critters: Where Mars Microbes Might Lurk". Space.com. Retrieved 16 December 2014.
  12. Thauer, R. K. & Shima, S. (2006). "Biogeochemistry: Methane and microbes". Nature. 440 (7086): 878–879. Bibcode:2006Natur.440..878T. doi:10.1038/440878a. PMID 16612369.
  13. 1 2 3 4 Gao B; Gupta RS (2007). "Phylogenomic analysis of proteins that are distinctive of Archaea and its main subgroups and the origin of methanogenesis". BMC Genomics. 8: 86. doi:10.1186/1471-2164-8-86.
  14. Gribaldo S; Brochier-Armanet C (2006). "The origin and evolution of Archaea: a state of the art". Philos Trans R Soc Lond B Biol Sci. 361 (1470): 1007–1022. doi:10.1098/rstb.2006.1841. PMC 1578729Freely accessible. PMID 16754611.
  15. M.J. Whiticar; et al. (1986). "Biogenic methane formation in marine and freshwater environments: CO2 reduction vs. acetate fermentation — isotope evidence". Geochim. Cosmochim. Acta. 50 (5): 393–709. Bibcode:1986GeCoA..50..693W. doi:10.1016/0016-7037(86)90346-7.
  16. Ferry, J.G. (1992). "Methane from acetate". Journal of Bacteriology. 174 (17): 5489–5495. PMC 206491Freely accessible. PMID 1512186. Retrieved 2011-11-05.
  17. Vogels, G.D.; Keltjens J.T.; Van Der Drift C. (1988). "Biochemistry of methane production". In Zehnder A.J.B. Biology of anaerobic microorganisms. New York: Wiley. pp. 707–770.
  18. Mondav, Rhiannon; Woodcroft, Ben J.; Kim, Eun-Hae; McCalley, Carmody K.; Hodgkins, Suzanne B.; Crill, Patrick M.; Chanton, Jeffrey; Hurst, Gregory B.; Verberkmoes, Nathan C.; Saleska, Scott R.; Hugenholtz, Philip; Rich, Virginia I.; Tyson, Gene W. (2014). "Discovery of a novel methanogen prevalent in thawing permafrost". Nature Communications. 5. doi:10.1038/ncomms4212.
This article is issued from Wikipedia - version of the 12/1/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.