Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-wq484 Total loading time: 0 Render date: 2024-04-27T22:33:41.099Z Has data issue: false hasContentIssue false

1 - Energy metabolism and phylogenetic diversity of sulphate-reducing bacteria

Published online by Cambridge University Press:  22 August 2009

Larry L. Barton
Affiliation:
University of New Mexico
W. Allan Hamilton
Affiliation:
University of Aberdeen
Get access

Summary

INTRODUCTION

Sulphate-reducing bacteria (SRB) are those prokaryotic microorganisms, both bacteria and archaea, that can use sulphate as the terminal electron acceptor in their energy metabolism, i.e. that are capable of dissimilatory sulphate reduction. Most of the SRB described to date belong to one of the four following phylogenetic lineages (with some examples of genera): (i) the mesophilic δ-proteobacteria with the genera Desulfovibrio, Desulfobacterium, Desulfobacter, and Desulfobulbus; (ii) the thermophilic Gram-negative bacteria with the genus Thermodesulfovibrio; (iii) the Gram-positive bacteria with the genus Desulfotomaculum; and (iv) the Euryarchaeota with the genus Archaeoglobus (Castro et al., 2000). A fifth lineage, the Thermodesulfobiaceae, has been described recently (Mori et al., 2003).

Many SRB are versatile in that they can use electron acceptors other than sulphate for anaerobic respiration. These include elemental sulphur (Bottcher et al., 2005; Finster et al., 1998), fumarate (Tomei et al., 1995), nitrate (Krekeler and Cypionka, 1995), dimethylsulfoxide (Jonkers et al., 1996), Mn(IV) (Myers and Nealson, 1988) and Fe(III) (Lovley et al., 1993; 2004). Some SRB are even capable of aerobic respiration (Dannenberg et al., 1992; Lemos et al., 2001) although this process appears not to sustain growth, and probably provides these organisms only with energy for maintenance. Since dissimilatory sulphate reduction is inhibited under oxic conditions, SRB can grow at the expense of sulphate reduction only in the complete absence of molecular oxygen. SRB are thus considered to be strictly anaerobic microorganisms and are mainly found in sulphate-rich anoxic habitats (Cypionka, 2000; Fareleira et al., 2003; Sass et al., 1992).

Type
Chapter
Information
Sulphate-Reducing Bacteria
Environmental and Engineered Systems
, pp. 1 - 38
Publisher: Cambridge University Press
Print publication year: 2007

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Akagi, J. M. (1995). Respiratory sulphate reduction. In Barton, L. L. (ed.), Sulphate-Reducing Bacteria, Vol. 8. New York: Plenum Press. pp. 89–111.CrossRefGoogle Scholar
Badziong, W. and Thauer, R. K. (1978). Growth yields and growth rates of Desulfovibrio vulgaris (Marburg) growing on hydrogen plus sulphate and hydrogen plus thiosulphate as the sole energy sources. Arch Microbiol, 117, 209–14CrossRefGoogle ScholarPubMed
Bak, F. and Cypionka, H. (1987). A novel type of energy-metabolism involving fermentation of inorganic sulfur-compounds. Nature, 326, 891–2CrossRefGoogle Scholar
Baron, E. J., Summanen, P., Downes, J.et al. (1989). Bilophila wadsworthia, gen. nov. and sp. nov., a unique gram-negative anaerobic rod recovered from appendicitis specimens and human faeces. J Gen Microbiol, 135, 3405–11Google ScholarPubMed
Beijerinck, M. W. (1895). Über Spirillum desulfuricans als Ursache von Sulfatreduktion. Zentralbl Bakteriol Parasitkd Infekt Abt II, 1, 49–59Google Scholar
Bottcher, M. E., Thamdrup, B., Gehre, M. and Theune, A. (2005). S-34/S-32 and O-18/O-16 fractionation during sulfur disproportionation by Desulfobulbus propionicus. Geomicrobiol J, 22, 219–26CrossRefGoogle Scholar
Broco, M., Rousset, M., Oliveira, S. and Rodrigues-Pousada, C. (2005). Deletion of flavoredoxin gene in Desulfovibrio gigas reveals its participation in thiosulphate reduction. FEBS Lett, 579, 4803–7CrossRefGoogle Scholar
Castro, H., Reddy, K. R. and Ogram, A. (2002). Composition and function of sulphate-reducing prokaryotes in eutrophic and pristine areas of the Florida Everglades. Appl Environ Microbiol, 68, 6129–37CrossRefGoogle Scholar
Castro, H. F., Williams, N. H. and Ogram, A. (2000). Phylogeny of sulphate-reducing bacteria. FEMS Microbiol Ecol, 31, 1–9Google Scholar
Conrad, R., Phelps, T. J. and Zeikus, J. G. (1985). Gas metabolism evidence in support of the juxtaposition of hydrogen-producing and methanogenic bacteria in sewage sludge and lake sediments. Appl Environ Microbiol, 50, 595–601Google ScholarPubMed
Crane, B. R., Siegel, L. M. and Getzoff, E. D. (1997). Structures of the siroheme- and Fe4S4-containing active center of sulfite reductase in different states of oxidation: Heme activation via reduction-gated exogenous ligand exchange. Biochemistry, 36, 12101–19CrossRefGoogle ScholarPubMed
Cypionka, H. (1987). Uptake of sulphate, sulfite and thiosulphate by proton-anion symport in Desulfovibrio desulfuricans. Arch Microbiol, 148, 144–9CrossRefGoogle Scholar
Cypionka, H. (2000). Oxygen respiration by Desulfovibrio species. Annu Rev Microbiol, 54, 827–48CrossRefGoogle ScholarPubMed
Dannenberg, S., Kroder, M., Dilling, W. and Cypionka, H. (1992). Oxidation of H2, organic-compounds and inorganic sulfur-compounds coupled to reduction of O2 or nitrate by sulphate-reducing bacteria. Arch Microbiol, 158, 93–9CrossRefGoogle Scholar
Dehning, I. and Schink, B. (1989). Malonomonas rubra gen. nov. sp. nov., a microaerotolerant anaerobic bacterium growing by decarboxylation of malonate. Arch Microbiol, 151, 427–33CrossRefGoogle Scholar
Dimroth, P. and Cook, G. M. (2004). Bacterial Na+- or H+-coupled ATP synthases operating at low electrochemical potential. Adv Microb Physiol, 49, 175–218CrossRefGoogle ScholarPubMed
Dolla, A., Pohorelic, B. K. J., Voordouw, J. K. and Voordouw, G. (2000). Deletion of the hmc operon of Desulfovibrio vulgaris subsp vulgaris Hildenborough hampers hydrogen metabolism and low-redox-potential niche establishment. Arch Microbiol, 174, 143–51CrossRefGoogle ScholarPubMed
Ehrenreich, A. and Widdel, F. (1994). Anaerobic oxidation of ferrous iron by purple bacteria, a new-type of phototrophic metabolism. Appl Environ Microb, 60, 4517–26Google ScholarPubMed
Ehrlich, H. L. (1999). Microbes as geologic agents: their role in mineral formation. Geomicrobiol J, 16, 135–53CrossRefGoogle Scholar
Fareleira, P., Santos, B. S., Antonio, C.et al. (2003). Response of a strict anaerobe to oxygen: survival strategies in Desulfovibrio gigas. Microbiolog-SGM, 149, 1513–22CrossRefGoogle ScholarPubMed
Farquhar, J. and Wing, B. A. (2003). Multiple sulfur isotopes and the evolution of the atmosphere. Earth and Planetary Science Letters, 213, 1–13CrossRefGoogle Scholar
Finster, K., Liesack, W. and Thamdrup, B. (1998). Elemental sulfur and thiosulphate disproportionation by Desulfocapsa sulfoexigens sp nov, a new anaerobic bacterium isolated from marine surface sediment. Appl Environ Microbiol, 64, 119–25Google ScholarPubMed
Fitz, R. M. and Cypionka, H. (1990). Formation of thiosulphate and trithionate during sulfite reduction by washed cells of Desulfovibrio desulfuricans. Arch Microbiol, 154, 400–6CrossRefGoogle Scholar
Forzi, L., Koch, J., Guss, A. M.et al. (2005). Assignment of the 4Fe-4S clusters of Ech hydrogenase from Methanosarcina barkeri to individual subunits via the characterization of site-directed mutants. FEBS J, 272, 4741–53CrossRefGoogle ScholarPubMed
Frederiksen, T. M. and Finster, K. (2003). Sulfite-oxido-reductase is involved in the oxidation of sulfite in Desulfocapsa sulfoexigens during disproportionation of thiosulphate and elemental sulfur. Biodegradation, 14, 189–98CrossRefGoogle ScholarPubMed
Fricke, W. F., Seedorf, H., Henne, A.et al. (2006). The genome sequence of Methanosphaera stadtmanae reveals why this human intestinal archaeon is restricted to methanol and H2 for methane formation and ATP synthesis. J Bacteriol, 188, 642–58CrossRefGoogle ScholarPubMed
Friedrich, M. W. (2002). Phylogenetic analysis reveals multiple lateral transfers of adenosine-5′-phosphosulphate reductase genes among sulphate-reducing microorganisms. J Bacteriol, 184, 278–89CrossRefGoogle ScholarPubMed
Fritz, G., Roth, A., Schiffer, A.et al. (2002). Structure of adenylylsulphate reductase from the hyperthermophilic Archaeoglobus fulgidus at 1.6-A resolution. Proc Natl Acad Sci USA, 99, 1836–41CrossRefGoogle Scholar
Galouchko, A. S. and Rozanova, E. P. (1996). Sulfidogenic oxidation of acetate by a syntrophic association of anaerobic mesophilic bacteria. Microbiology, 65, 134–9Google Scholar
Garrity, G. M., Bell, J. A. and Lilburn, T. G. (2003). Taxonomic outline of the procaryotes. Bergey's Manual of Systematic Bacteriology. Second Edition. Release 5.0, New York: Springer Verlag 401 pages. DOI: 10.1007/bergeysoutline200405 (http://dx.doi.org/10.1007/bergeysoutline200405) New York: Springer-Verlag.Google Scholar
Ghiorse, W. C. (1984). Biology of iron- and manganese-depositing bacteria. Annu Rev Microbiol, 38, 515–50CrossRefGoogle ScholarPubMed
Goenka, A., Voordouw, J. K., Lubitz, W., Gartner, W. and Voordouw, G. (2005). Construction of a NiFe-hydrogenase deletion mutant of Desulfovibrio vulgaris Hildenborough. Biochem Soc Trans, 33, 59–60CrossRefGoogle ScholarPubMed
Hamilton, W. A. (2003). Microbially influenced corrosion as a model system for the study of metal microbe interactions: a unifying electron transfer hypothesis. Biofouling, 19, 65–76CrossRefGoogle Scholar
Hansen, T. A. (1994). Metabolism of sulphate-reducing prokaryotes. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology, 66, 165–85CrossRefGoogle Scholar
Haveman, S. A., Brunelle, V., Voordouw, J. K.et al. (2003). Gene expression analysis of energy metabolism mutants of Desulfovibrio vulgaris Hildenborough indicates an important role for alcohol dehydrogenase. J Bacteriol, 185, 4345–53CrossRefGoogle ScholarPubMed
Haynes, T. S., Klemm, D. J., Ruocco, J. J. and Barton, L. L. (1995). Formate dehydrogenase activity in cells and outer-membrane blebs of Desulfovibrio gigas. Anaerobe, 1, 175–82CrossRefGoogle ScholarPubMed
Hedderich, R. (2004). Energy-converting NiFe hydrogenases from archaea and extremophiles: ancestors of complex I. J Bioenerg Biomembr, 36, 65–75CrossRefGoogle ScholarPubMed
Hedderich, R., Klimmek, O., Kroeger, A.et al. (1998). Anaerobic respiration with elemental sulfur and with disulfides. FEMS Microbiol Rev, 22, 353–81CrossRefGoogle Scholar
Heidelberg, J. F., Seshadri, R., Haveman, S. A.et al. (2004). The genome sequence of the anaerobic, sulphate-reducing bacterium Desulfovibrio vulgaris Hildenborough. Nat Biotechnol, 22, 554–9CrossRefGoogle Scholar
Hemme, C. L. and Wall, J. D. (2004). Genomic insights into gene regulation of Desulfovibrio vulgaris Hildenborough. Omics, 8, 43–55CrossRefGoogle ScholarPubMed
Hines, M. E., Evans, R. S., Sharak Genthner, B. R.et al. (1999). Molecular phylogenetic and biogeochemical studies of sulphate-reducing bacteria in the rhizosphere of Spartina alterniflora. Appl Environ Microbiol, 65, 2209–16Google Scholar
Hoehler, T., Alperin, M. J., Albert, D. B. and Martens, C. S. (2001). Apparent minimum free energy requirements for methanogenic Archaea and sulphate-reducing bacteria in an anoxic marine sediment. FEMS Microbiol Ecol, 38, 33–41CrossRefGoogle Scholar
Holmer, M. and Storkholm, P. (2001). Sulphate reduction and sulphur cycling in lake sediments: a review. Freshwater Biol, 46, 431–51CrossRefGoogle Scholar
Houwen, F. P., Dijkema, C., Stams, A. J. M. and Zehnder, A. J. B. (1991). Propionate metabolism in anaerobic bacteria – determination of carboxylation reactions with C-13-NMR spectroscopy. Biochim Biophys Acta, 1056, 126–32CrossRefGoogle Scholar
Johnston, D. T., Wing, B. A., Farquhar, J.et al. (2005). Active microbial sulfur disproportionation in the Mesoproterozoic. Science, 310, 1477–9CrossRefGoogle ScholarPubMed
Jonkers, H. M., Maarel, M. J. E. C., Gemerden, H. and Hansen, T. A. (1996). Dimethylsulfoxide reduction by marine sulphate-reducing bacteria. FEMS Microbiol Lett, 136, 283–7CrossRefGoogle Scholar
Jorgensen, B. B. (1982). Ecology of the bacteria of the sulfur cycle with special reference to anoxic oxic interface environments. Philo Trans Roy Soc Ser B, 298, 543–61CrossRefGoogle ScholarPubMed
Keon, R. G. and Voordouw, G. (1996). Identification of the HmcF and topology of the HmcB subunit of the Hmc complex of Desulfovibrio vulgaris. Anaerobe, 2, 231–8CrossRefGoogle Scholar
Klenk, H. P., Clayton, R. A., Tomb, J. F.et al. (1997). The complete genome sequence of the hyperthermophilic, sulphate-reducing archaeon Archaeoglobus fulgidus. Nature, 390, 364–70CrossRefGoogle ScholarPubMed
Kobayashi, K., Morisawa, Y., Ishituka, T. and Ishimoto, M. (1975). Biochemical studies on sulphate-reducing bacteria. 14. Enzyme levels of adenylylsulphate reductase, inorganic pyrophosphatase, sulfite reductase, hydrogenase, and adenosine-triphosphatase in cells grown on sulphate, sulfite, and thiosulphate. J Biochem (Tokyo), 78, 1079–85CrossRefGoogle Scholar
Kopp, R. E., Kirschvink, J. L., Hilburn, I. A. and Nash, C. Z. (2005). The paleoproterozoic snowball Earth: a climate disaster triggered by the evolution of oxygenic photosynthesis. Proc Natl Acad Sci USA, 102, 11131–6CrossRefGoogle ScholarPubMed
Koizumi, Y., Kelly, J. J., Nakagawa, T.et al. (2002). Parallel characterization of anaerobic toluene- and ethylbenzene-degrading microbial consortia by PCR-denaturing gradient gel electrophoresis, RNA-DNA membrane hybridization, and DNA microarray technology. Appl Environ Microbiol, 68, 3215–25CrossRefGoogle ScholarPubMed
Kramer, M. and Cypionka, H. (1989). Sulphate formation via ATP sulfurylase in thiosulphate-disproportionating and sulfite-disproportionating bacteria. Arch Microbiol, 151, 232–7CrossRefGoogle Scholar
Kreke, B. and Cypionka, H. (1992). Proton motive force in fresh-water sulphate-reducing bacteria, and its role in sulphate accumulation in Desulfobulbus propionicus. Arch Microbiol, 158, 183–7CrossRefGoogle Scholar
Kreke, B. and Cypionka, H. (1994). Role of sodium-ions for sulphate transport and energy-metabolism in Desulfovibrio salexigens. Arch Microbiol, 161, 55–61CrossRefGoogle Scholar
Krekeler, D. and Cypionka, H. (1995). The preferred electron-acceptor of Desulfovibrio desulfuricans Csn. FEMS Microbiol Ecol, 17, 271–7CrossRefGoogle Scholar
Kremer, D. R. and Hansen, T. A. (1988). Pathway of propionate degradation in Desulfobulbus propionicus. FEMS Microbiol Lett, 49, 273–7CrossRefGoogle Scholar
Larsen, O., Lien, T. and Birkeland, N. K. (1999). Dissimilatory sulfite reductase from Archaeoglobus profundus and Desulfotomaculum thermocisternum: phylogenetic and structural implications from gene sequences. Extremophiles, 3, 63–70CrossRefGoogle ScholarPubMed
Leaphart, A. B., Friez, M. J. and Lovell, C. R. (2003). Formyltetrahydrofolate synthetase sequences from salt marsh plant roots reveal a diversity of acetogenic bacteria and other bacterial functional groups. Appl Environ Microbiol, 69, 693–6CrossRefGoogle ScholarPubMed
Le, M. J. and Zinder, S. H. (1988). Isolation and characterization of a thermophilic bacterium which oxidizes acetate in syntrophic association with a methanogen and which grows acetogenically on H2-CO2. Appl Environ Microbiol, 54, 124–9Google Scholar
Lemos, R. S., Gomes, C. M., Santana, M.et al. (2001). The “strict” anaerobe Desulfovibrio gigas contains a membrane-bound oxygen respiratory chain. J Inorg Biochem, 86, 314Google Scholar
Liu, M. Y. and Legall, J. (1990). Purification and characterization of 2 proteins with inorganic pyrophosphatase activity from Desulfovibrio vulgaris – rubrerythrin and a new, highly-active,enzyme. Biochem Biophys Res Commun, 171, 313–18CrossRefGoogle Scholar
Lopez-Cortes, A., Bursakov, S., Figueiredo, A.et al. (2005). Purification and preliminary characterization of tetraheme cytochrome c(3) and adenylylsulphate reductase from the peptidolytic sulphate-reducing bacterium Desulfovibrio aminophilus DSM 12254. Bioinorg Chem Appl, 3, 81–91CrossRefGoogle Scholar
Lovley, D. R., Holmes, D. E. and Nevin, K. P. (2004). Dissimilatory Fe(III) and Mn(IV) reduction. Adv Microb Physiol, 49, 221–86Google ScholarPubMed
Lovley, D. R., Roden, E. E., Phillips, E. J. P. and Woodward, J. C. (1993). Enzymatic iron and uranium reduction by sulphate-reducing bacteria. Mar Geol, 113, 41–53CrossRefGoogle Scholar
Malki, S., DeLuca, G., Fardeau, M. L.et al. (1997). Physiological characteristics and growth behavior of single and double hydrogenase mutants of Desulfovibrio fructosovorans. Arch Microbiol, 167, 38–45CrossRefGoogle ScholarPubMed
Mander, G. J., Duin, E. C., Linder, D., Stetter, K. O. and Hedderich, R. (2002). Purification and characterization of a membrane-bound enzyme complex from the sulphate-reducing archaeon Archaeoglobus fulgidus related to heterodisulfide reductase from methanogenic archaea. Eur J Biochem, 269, 1895–904CrossRefGoogle Scholar
Mander, G. J., Pierik, A. J., Huber, H. and Hedderich, R. (2004). Two distinct heterodisulfide reductase-like enzymes in the sulphate-reducing archaeon Archaeoglobus profundus. Eur J Biochem, 271, 1106–16CrossRefGoogle Scholar
Matias, P. M., Pereira, I. A. C., Soares, C. M. and Carrondo, M. A. (2005). Sulphate respiration from hydrogen in Desulfovibrio bacteria: a structural biology overview. Prog Biophys Mol Biol, 89, 292–329CrossRefGoogle ScholarPubMed
McOrist, S., Gebhart, C. J., Boid, R. and Barns, S. M. (1995). Characterization of Lawsonia intracellularis gen. nov., sp. nov., the obligately intracellular bacterium of porcine proliferative enteropathy. Int J Syst Bacteriol, 45, 820–5CrossRefGoogle ScholarPubMed
Meier, T., Polzer, P., Diederichs, K., Welte, W. and Dimroth, P. (2005). Structure of the rotor ring of F-type Na+-ATPase from Ilyobacter tartaricus. Science, 308, 659–62CrossRefGoogle ScholarPubMed
Meuer, J., Kuettner, H. C., Zhang, J. K., Hedderich, R. and Metcalf, W. W. (2002). Genetic analysis of the archaeon Methanosarcina barkeri Fusaro reveals a central role for Ech hydrogenase and ferredoxin in methanogenesis and carbon fixation. Proc Natl Acad Sci USA, 99, 5632–7CrossRefGoogle ScholarPubMed
Molitor, M., Dahl, C., Molitor, I.et al. (1998). A dissimilatory sirohaem-sulfite-reductase-type protein from the hyperthermophilic archaeon Pyrobaculum islandicum. Microbiology-SGM, 144, 529–41CrossRefGoogle ScholarPubMed
Monster, J., Appel, P. W. U., Thode, H. G.et al. (1979). Sulfur isotope studies in early archaean sediments from Isua, West Greenland – implications for the antiquity of bacterial sulphate reduction. Geochim Cosmochim Acta, 43, 405–13CrossRefGoogle Scholar
Mori, K., Kim, H., Kakegawa, T. and Hanada, S. (2003). A novel lineage of sulphate-reducing microorganisms: Thermodesulfobiaceae fam. nov., Thermodesulfobium narugense, gen. nov., sp nov., a new thermophilic isolate from a hot spring. Extremophiles, 7, 283–90CrossRefGoogle Scholar
Mueller, V. (2004). An exceptional variability in the motor of archaeal A(1)A(0) ATPases: from multimeric to monomeric rotors comprising 6–13 ion binding sites. J Bioenerg Biomembr, 36, 115–25CrossRefGoogle Scholar
Murata, T., Yamato, I., Kakinuma, Y., Leslie, A. G. W. and Walker, J. E. (2005). Structure of the rotor of the V-type Na+-ATPase from Enterococcus hirae. Science, 308, 654–9CrossRefGoogle ScholarPubMed
Myers, C. R. and Nealson, K. H. (1988). Bacterial manganese reduction and growth with manganese oxide as the sole electron-acceptor. Science, 240, 1319–21CrossRefGoogle ScholarPubMed
Nealson, K. H. and Saffarini, D. (1994). Iron and manganese in anaerobic respiration – environmental significance, physiology, and regulation. Annu Rev Microbiol, 48, 311–43CrossRefGoogle ScholarPubMed
Nielsen, P. H., Lee, W., Lewandowski, Z., Morrison, M. and Characklis, W. G. (1993). Corrosion of mild steel in an alternating oxic and anoxic biofilm system. Biofouling, 7, 267–84CrossRefGoogle Scholar
Odom, J. M. and Peck, H. D. (1984). Hydrogenase, electron-transfer proteins, and energy coupling in the sulphate-reducing bacteria Desulfovibrio. Annu Rev Microbiol, 38, 551–92CrossRefGoogle Scholar
Ogata, M., Arihara, K. and Yagi, T. (1981). D-Lactate dehydrogenase of Desulfovibrio vulgaris. J Biochem (Tokyo), 89, 1423–31CrossRefGoogle ScholarPubMed
Pankhania, I. P., Spormann, A. M., Hamilton, W. A. and Thauer, R. K. (1988). Lactate conversion to acetate, CO2 and H2 in cell suspensions of Desulfovibrio vulgaris (Marburg): indications for the involvement of an energy driven reaction. Arch Microbiol, 150, 26–31CrossRefGoogle Scholar
Paulsen, J., Kröger, A. and Thauer, R. K. (1986). ATP-driven succinate oxidation in the catabolism of Desulfuromonas acetoxidans. Arch Microbiol, 144, 78–83CrossRefGoogle Scholar
Peck, H. D. (1993). Bioenergetic strategies of the sulphate-reducing bacteria. In Odom, J. M. and Singleton, J. Rivers (eds.), The Sulphate-Reducing Bacteria: Contemporary Perspectives. New York, London: Springer-Verlag. pp. 41–76.CrossRefGoogle Scholar
Pires, R. H., Lourenco, A. I., Morais, F.et al. (2003). A novel membrane-bound respiratory complex from Desulfovibrio desulfuricans ATCC 27774. Biochim Biophys Acta-Bioenergetics, 1605, 67–82CrossRefGoogle ScholarPubMed
Pohorelic, B. K. J., Voordouw, J. K., Lojou, E.et al. (2002). Effects of deletion of genes encoding Fe-only hydrogenase of Desulfovibrio vulgaris Hildenborough on hydrogen and lactate metabolism. J Bacteriol, 184, 679–86CrossRefGoogle ScholarPubMed
Rabus, R., Fukui, M., Wilkes, H. and Widdel, F. (1996). Degradative capacities and 16S rRNA-targeted whole-cell hybridization of sulphate-reducing bacteria in an anaerobic enrichment culture utilizing alkylbenzenes from crude oil. Appl Environ Microbiol, 62, 3605–13Google Scholar
Rabus, R., Hansen, T., and Widdel, F. (2001). An evolving electronic resource for the microbiological community. In Dworkin, S., Falkow, M., Rosenberg, E., Schleifer, K.-H. and Stackebrandt, E. (eds.), The Prokaryotes. New York: Springer-Verlag. pp. release 3.3, http://link.springer-ny.com/link/service/books/10125.Google Scholar
Rabus, R., Ruepp, A., Frickey, T.et al. (2004). The genome of Desulfotalea psychrophila, a sulphate-reducing bacterium from permanently cold Arctic sediments. Environ Microbiol, 6, 887–902CrossRefGoogle Scholar
Reed, D. W. and Hartzell, P. L. (1999). The Archaeoglobus fulgidus D-lactate dehydrogenase is a Zn2+ flavoprotein. J Bacteriol, 181, 7580–7Google Scholar
Reguera, G., McCarthy, K. D., Mehta, T.et al. (2005). Extracellular electron transfer via microbial nanowires. Nature, 435, 1098–101CrossRefGoogle ScholarPubMed
Rodrigues, R., Valente, F. M. A., Pereira, I. A. C., Oliveira, S. and Rodrigues-Pousada, C. (2003). A novel membrane-bound Ech NiFe hydrogenase in Desulfovibrio gigas. Biochem Biophys Res Commun, 306, 366–75CrossRefGoogle ScholarPubMed
Rossi, M., Pollock, W. B. R., Reij, M. W.et al. (1993). The Hmc operon of Desulfovibrio vulgaris Subsp vulgaris Hildenborough encodes a potential transmembrane redox protein complex. J Bacteriol, 175, 4699–711CrossRefGoogle ScholarPubMed
Sapra, R., Bagramyan, K. and Adams, M. W. W. (2003). A simple energy-conserving system: proton reduction coupled to proton translocation. Proc Natl Acad Sci USA, 100, 7545–50CrossRefGoogle ScholarPubMed
Sass, H., Steuber, J., Kroder, M., Kroneck, P. M. H. and Cypionka, H. (1992). Formation of thionates by fresh-water and marine strains of sulphate-reducing bacteria. Arch Microbiol, 158, 418–21CrossRefGoogle Scholar
Sato, K., Nishina, Y., Setoyama, C., Miura, R. and Shiga, K. (1999). Unusually high standard redox potential of acrylyl-CoA/propionyl-CoA couple among enoyl-CoA/acyl-CoA couples: a reason for the distinct metabolic pathway of propionyl-CoA from longer acyl-CoAs. J Biochem (Tokyo), 126, 668–75CrossRefGoogle ScholarPubMed
Schiffers, A. and Jorgensen, B. B. (2002). Biogeochemistry of pyrite and iron sulfide oxidation in marine sediments. Geochim Cosmochim Acta, 66, 85–92Google Scholar
Schink, B. (1992). The genus Pelobacter. In Balows, A., Trüper, H. G., Dworkin, M., Harder, W. and Schleifer, K.-H. (eds.), The Prokaryotes. New York: Springer-Verlag. pp. 3393–9.CrossRefGoogle Scholar
Schink, B. and Stams, A. J. (2002). Syntrophism among Prokaryotes. In Dworkin, M. (ed.), The Prokaryotes (electronic version). New York: Springer Verlag. pp. 309–35.Google Scholar
Shen, Y. N. and Buick, R. (2004). The antiquity of microbial sulphate reduction. Earth-Science Reviews, 64, 243–72CrossRefGoogle Scholar
Shigematsu, T., Tang, Y. Q. Kobayashi, T. et al. (2004). Effect of dilution rate on metabolic pathway shift between aceticlastic and nonaceticlastic methanogenesis in chemostat cultivation. Appl Environ Microbiol, 70, 4048–52CrossRefGoogle ScholarPubMed
Shima, S. and Thauer, R. K. (2005). Methyl-coenzyme M reductase (MCR) and the anaerobic oxidation of methane (AOM) in methanotrophic archaea. Curr Opin Microbiol, 8, 643–8CrossRefGoogle Scholar
Soboh, B., Linder, D. and Hedderich, R. (2002). Purification and catalytic properties of a CO-oxidizing: H2-evolving enzyme complex from Carboxydothermus hydrogenoformans. Eur J Biochem, 269, 5712–21CrossRefGoogle ScholarPubMed
Sperling, D., Kappler, U., Wynen, A., Dahl, C. and Truper, H. G. (1998). Dissimilatory ATP sulfurylase from the hyperthermophilic sulphate reducer Archaeoglobus fulgidus belongs to the group of homo-oligomeric ATP sulfurylases. FEMS Microbiol Lett, 162, 257–64CrossRefGoogle Scholar
Stackebrandt, E. (1995). Origin and evolution of prokaryotes. In Gibbs, A. J., Calisher, C. H. and Garcia-Arenal, F. (eds.), Molecular Basis of Virus Evolution. Cambridge: Cambridge University Press, pp. 224–52.CrossRefGoogle Scholar
Stackebrandt, E. (2004). The phylogeny and classification of anaerobic bacteria. In Nakano, M. M. and Zuber, P. (eds.), Strict and Facultative Anaerobes. Medical and Environmental Aspects. Wymondham, UK: Horizon Bioscience. pp. 1–25.Google Scholar
Steger, J. L., Vincent, C., Ballard, J. D. and Krumholz, L. R. (2002). Desulfovibrio sp genes involved in the respiration of sulphate during metabolism of hydrogen and lactate. Appl Environ Microbiol, 68, 1932–7CrossRefGoogle ScholarPubMed
Taguchi, Y., Sugishima, M. and Fukuyama, K. (2004). Crystal structure of a novel zinc-binding ATP sulfurylase from Thermus thermophilus HB8. Biochemistry, 43, 4111–18CrossRefGoogle ScholarPubMed
Thamdrup, B. and Canfield, D. E. (1996). Pathways of carbon oxidation in continental margin sediments off central Chile. Limnol Oceanogr, 41, 1629–50CrossRefGoogle ScholarPubMed
Thamdrup, B., Fossing, H. and Jorgensen, B. B. (1994). Manganese, iron and sulfur cycling in a coastal marine sediment, Aarhus Bay, Denmark. Geochim Cosmochim Acta, 58, 5115–29CrossRefGoogle Scholar
Thauer, R. K. (1988). Citric acid cycle, 50 years on: modifications and an alternative pathway in anaerobic bacteria. Eur. J. Biochem., 176, 497–508CrossRefGoogle ScholarPubMed
Thauer, R. K., Jungermann, K. and Decker, K. (1977). Energy conservation in chemotrophic anaerobic bacteria. Bacteriol. Rev. 41, 100–80Google ScholarPubMed
Thauer, R. K. and Kunow, J. (1995). Sulphate reducing Archaea. In Clark, N. (ed.), Biotechnology Handbook. London: Plenum Publishing. pp. 33–48.Google Scholar
Thauer, R. K., Möller-Zinkhan, D. and Spormann, A. (1989). Biochemistry of acetate catabolism in anaerobic chemotrophic bacteria. Annu. Rev. Microbiol., 43, 43–67CrossRefGoogle ScholarPubMed
Thauer, R. K. and Morris, J. G. (1984). Metabolism of chemotrophic anaerobes: old views and new aspects. In Kelly, D. P. and Carr, N. G. (eds.), The Microbe: 1984 Part II: Prokaryotes and Eukaryotes. Society for General Microbiology Symposium 36. Cambridge: Cambridge University Press. pp. 123–68.Google Scholar
Tomei, F. A., Barton, L. L., Lemanski, C. L.et al. (1995). Transformation of selenate and selenite to elemental selenium by Desulfovibrio desulfuricans. J Ind Microbiol, 14, 329–36CrossRefGoogle Scholar
Venter, J. C., Remington, K., Heidelberg, J. F.et al. (2004). Environmental genome shotgun sequencing of the Sargasso Sea. Science, 304, 66–74CrossRefGoogle ScholarPubMed
Voordouw, G. (2002). Carbon monoxide cycling by Desulfovibrio vulgaris Hildenborough. J Bacteriol, 184, 5903–11CrossRefGoogle ScholarPubMed
Voordouw, G., Armstrong, S. M., Reimer, M. F.et al. (1996). Characterization of 16S rRNA genes from oil field microbial communities indicates the presence of a variety of sulphate-reducing, fermentative, and sulfide-oxidizing bacteria. Appl Environ Microb, 62, 1623–9Google Scholar
Wächtershäuser, G. (1992). Groundworks for an evolutionary biochemistry: the iron-sulphur world. Prog Biophys Mol Biol, 58, 85–201CrossRefGoogle ScholarPubMed
Ware, D. A. and Postgate, J. R. (1971). Physiological and chemical properties of a reductant-activated inorganic pyrophosphatase from Desulfovibrio desulfuricans. J Gen Microbiol, 67, 145–60CrossRefGoogle ScholarPubMed
Weinberg, M. V., Jenney, F. E., Cui, X. Y. and Adams, M. W. W. (2004). Rubrerythrin from the hyperthermophilic archaeon Pyrococcus furiosus is a rubredoxin-dependent, iron-containing peroxidase. J Bacteriol, 186, 7888–95CrossRefGoogle ScholarPubMed
Widdel, F. and Bak, F. (1992). Gram-negative mesophilic sulphate-reducing bacteria. In Balows, A., Trüper, H. G., Dworkin, M., Harder, W. and Schleifer, K.-H. (eds.), The Prokaryotes. New York: Springer-Verlag. pp. 3352–78.CrossRefGoogle Scholar
Widdel, F. and Pfennig, N. (1982). Studies on dissimilatory sulphate-reducing bacteria that decompose fatty-acids. 2. Incomplete oxidation of propionate by Desulfobulbus propionicus Gen-Nov, Sp-Nov. Arch Microbiol, 131, 360–5CrossRefGoogle Scholar
Widdel, F. and Pfennig, N. (1984). Dissimilatory sulphate- and sulfur-reducing bacteria. In Krieg, N. R. and Holt, J. G. (eds.), Bergey's Manual of Systematic Bacteriology. Baltimore, MD: Williams and Wilkins. pp. 663–79.Google Scholar
Winogradsky, S. (1890). Recherches sur les organismes de la nitrification. Compt Rendue, 110, 1013–16. In Brock, T. D. (ed.), Milestones in Microbiology: 1556 to 1940. ASM Press: Washington DC (1998). pp. 231–33.Google Scholar
Yagi, T. and Ogata, M. (1996). Catalytic properties of adenylylsulphate reductase from Desulfovibrio vulgaris Miyazaki. Biochimie, 78, 838–46CrossRefGoogle Scholar
Zengler, K., Richnow, H. H., Rossello-Mora, R., Michaelis, W. and Widdel, F. (1999). Methane formation from long-chain alkanes by anaerobic microorganisms. Nature, 401, 266–9CrossRefGoogle ScholarPubMed
Zverlov, V., Klein, M., Lucker, S.et al. (2005). Lateral gene transfer of dissimilatory (bi)sulfite reductase revisited. J Bacteriol, 187, 2203–8CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×