Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-tj2md Total loading time: 0 Render date: 2024-04-19T21:55:57.737Z Has data issue: false hasContentIssue false

6 - Biochemical, proteomic and genetic characterization of oxygen survival mechanisms in sulphate-reducing bacteria of the genus Desulfovibrio

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 anaerobes, which derive energy for growth from anaerobic metabolism, coupling the oxidation of organic substrates with the dissimilatory reduction of sulphate to hydrogen sulphide (sulphate respiration). Although generally considered as strict anaerobes, more and more data indicate a higher abundance and metabolic activity in oxic zones of biotopes, such as marine and freshwater sediments, than in neighbouring anoxic zones (Ravenschlag et al., 2000; Sass et al., 1997; 1998). A well-documented example of sulphate reduction under oxic conditions is also provided by cyanobacterial mats. Here a zone of photosynthetic oxygen synthesis overlaps with a zone of sulphide production by SRB and a zone of oxygen-dependent microbial sulphide oxidation, creating steep, opposing gradients of oxygen and sulphide, which fluctuate with the rhythm of day and night (Canfield and des Marais, 1991; Teske et al., 1998; Caumette et al., 1994). In cyanobacterial mats from the saline evaporation pond in Baja California, Desulfobacter and Desulfobacterium are restricted to greater depths while the Desulfococcus and Desulfovibrio groups are predominant in the upper part of the photo-oxic zone (Risatti et al., 1994). The high numbers of SRB found in these oxic environments indicate that these organisms are able to deal with temporal exposures to oxygen concentrations as high as 1.5 mM (Sigalevich and Cohen, 2000).

In these oxygen-exposed systems SRB of the genus Desulfovibrio are among the most oxygen-tolerant, e.g. Desulfovibrio oxyclinae was isolated from Solar Lake microbial mats (Krekeler et al., 1997) and Desulfovibrio desulfuricans strain DvO1 was isolated from activated sludge aerated to atmospheric oxygen saturation (Kjeldsen et al., 2005).

Type
Chapter
Information
Sulphate-Reducing Bacteria
Environmental and Engineered Systems
, pp. 185 - 214
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

Abdollahi, H. and Wimpenny, J. W. T. (1990). Effects of the oxygen on the growth of Desulfovibrio desulfuricans. J Gen Microbiol, 136, 1025–30.CrossRefGoogle Scholar
Abreu, I. A., Xavier, A. V., LeGall, J., Cabelli, D. E. and Teixeira, M. (2002). Superoxide scavenging by neelaredoxin: dismutation and reduction activities in anaerobes. J Biol Inorg Chem, 7, 668–74.CrossRefGoogle ScholarPubMed
Alban, P. S. and Krieg, N. R. (1998). A hydrogen peroxide resistant mutant of Spirillum volutans has NADH peroxidase activity but no increased oxygen tolerance. Can J Microbiol, 44, 87–91.CrossRefGoogle Scholar
Alves, T., Besson, S., Duarte, L. C.et al. (1999). A cytochrome c peroxidase from Pseudomonas nautica 617 active at high ionic strength: expression, purification and characterization. Biochim Biophys Acta, 1434, 248–59.CrossRefGoogle ScholarPubMed
Barynin, V. V., Whittaker, M. M., Antonyuk, S. V.et al. (2001). Crystal structure of manganese catalase from Lactobacillus plantarum. Structure (Camb.), 9, 725–38.CrossRefGoogle ScholarPubMed
Baumgarten, A., Redenius, I., Kranczoch, J. and Cypionka, H. (2001). Periplasmic reduction by Desulfovibrio species. Arch Microbiol, 176, 306–9.CrossRefGoogle ScholarPubMed
Bramlett, R. N. and Peck, H. D. Jr. (1975). Some physical and kinetic properties of adenylyl sulphate reductase from Desulfovibrio vulgaris. J Biol Chem, 250, 2979–86.Google ScholarPubMed
Brioukhanov, A. L. and Netrusov, A. I. (2004). Catalase and superoxide dismutase: distribution, properties, and physiological role in cells of strict anaerobes. Biochemistry (Mosc), 69, 949–62.CrossRefGoogle ScholarPubMed
Brumlik, M. J. and Voordouw, G. (1989). Analysis of the transcriptional unit encoding the genes for rubredoxin (rub) and a putative rubredoxin oxidoreductase (rbo) in Desulfovibrio vulgaris (Hildenborough). J Bacteriol, 171, 4996–5004.CrossRefGoogle Scholar
Canfield, D. E. and Des Marais, D. J. (1991). Aerobic sulphate reduction in microbial mats. Science, 251, 1471–3.CrossRefGoogle ScholarPubMed
Carmel-Harel, O. and Storz, G. (2000). Roles of the glutathione- and thioredoxin-dependent reduction systems in the Escherichia coli and Saccharomyces cerevisiae responses to oxidative stress. Annu Rev Microbiol, 54, 439–61.CrossRefGoogle ScholarPubMed
Caumette, P., Matheron, R., Raymond, N. and Relexans, J.-C. (1994). Microbial mats in the hypersaline ponds of Mediterranean salterns (Salins de Giraud, France). FEMS Microbiol Ecol, 13, 273–86.CrossRefGoogle Scholar
Chen, L., Liu, M.-Y., LeGall, J.et al. (1993a). Purification and characterization of an NADH-rubredoxin oxidoreductase involved in the utilization of oxygen byDesulfovibrio gigas. Eur J Biochem, 216, 443–8.CrossRefGoogle Scholar
Chen, L., Liu, M.-Y., LeGall, J.et al. (1993b). Rubredoxin oxidase, a new flavo-hemo-protein, is the site of oxygen reduction to water by the “strict anaerobe”Desulfovibrio gigas. Biochem Biophys Res Commun, 193, 100–5.CrossRefGoogle Scholar
Chen, L., Sharma, P., Gall, J.et al. (1994). A blue non-heme iron protein from Desulfovibrio gigas. Eur J Biochem, 226, 613–18.CrossRefGoogle ScholarPubMed
Choudhury, S. B., Lee, J. W., Davidson, G.et al. (1999). Examination of the nickel site structure and reaction mechanism in Streptomyces seoulensis superoxide dismutase. Biochemistry, 38, 3744–52.CrossRefGoogle ScholarPubMed
Coulter, E. D. and Kurtz, D. M. Jr. (2001). A role for rubredoxin in oxidative stress protection in Desulfovibrio vulgaris: catalytic electron transfer to rubrerythrin and two-iron superoxide reductase. Arch Biochem Biophys, 394, 76–86.CrossRefGoogle ScholarPubMed
Coulter, E. D., Shenvi, N. V. and Kurtz, D. M. Jr. (1999). NADH peroxidase activity of rubrerythrin. Biochem Biophys Res Commun, 255, 317–23.CrossRefGoogle ScholarPubMed
Cypionka, H. (2000). Oxygen respiration by Desulfovibrio species. Annu Rev Microbiol, 54, 827–48.CrossRefGoogle ScholarPubMed
da Costa, P. N., Romão, C. V., LeGall, J.et al. (2001). The genetic organization of Desulfovibrio desulfuricans ATCC 27774 bacterioferritin and rubredoxin-2 genes. Involvement of rubredoxin in the iron metabolism. Mol Microbiol, 41, 217–29.CrossRefGoogle Scholar
Das, A., Coulter, E. D., Kurtz, D. M. Jr. and Ljungdahl, L. G. (2001). Five-gene cluster in Clostridium thermoaceticum consisting of two divergent operons encoding rubredoxin oxidoreductase-rubredoxin and rubrerythrin-type A flavoprotein-high-molecular-weight rubredoxin. J Bacteriol, 183, 1560–7.CrossRefGoogle ScholarPubMed
deMaré, F., Kurtz, D. M. Jr. and Nordlund, P. (1996). The structure of Desulfovibrio vulgaris rubrerythrin reveals a unique combination of rubredoxin-like FeS4 and ferritin-like diiron domains. Nature Struct Biol, 3, 539–46.CrossRefGoogle ScholarPubMed
Dos Santos, W. G., Pacheco, I., Liu, M. Y.et al. (2000). Purification and characterization of an iron superoxide dismutase and a catalase from the sulphate-reducing bacterium Desulfovibrio gigas. J Bacteriol, 182, 796–804.CrossRefGoogle Scholar
Emerson, J. P., Coulter, E. D., Phillips, R. S. and Kurtz, D. M. Jr. (2003). Kinetics of the superoxide reductase catalytic cycle. J Biol Chem, 278, 39662–8.CrossRefGoogle ScholarPubMed
Eschemann, A., Kèuhl, M. and Cypionka, H. (1999). Aerotaxis in Desulfovibrio. Environ Microbiol, 1, 489–94.CrossRefGoogle ScholarPubMed
Frazão, C., Silva, G., Gomes, C. M.et al. (2000). Structure of a dioxygen reduction enzyme from Desulfovibrio gigas. Nature Struct Biol, 7, 1041–5.Google ScholarPubMed
Fournier, M., Aubert, C., Dermoun, Z.et al. (2006). Response of the anaerobe Desulfovibrio vulgaris Hildenborough to oxidative conditions: proteome and transcript analysis. Biochimie, 88, 85–94.CrossRefGoogle ScholarPubMed
Fournier, M., Dermoun, Z., Durand, M. C. and Dolla, A. (2004). A new function of the Desulfovibrio vulgaris Hildenborough Fe hydrogenase in the protection against oxidative stress. J Biol Chem, 279, 1787–93.CrossRefGoogle ScholarPubMed
Fournier, M., Zhang, Y., Wildschut, J. D.et al. (2003). Function of oxygen resistance proteins in the anaerobic, sulphate-reducing bacterium, Desulfovibrio vulgaris Hildenborough. J Bacteriol, 185, 71–9.CrossRefGoogle Scholar
Fritz, G., Bèuchert, T. and Kroneck, P. M. (2002). The function of the 4Fe-4S clusters and FAD in bacterial and archaeal adenylylsulphate reductases. Evidence for flavin-catalyzed reduction of adenosine 5′-phosphosulphate. J Biol Chem, 277, 26066–73.CrossRefGoogle Scholar
Fu, R. and Voordouw, G. (1997). Targeted gene-replacement mutagenesis of dcrA, encoding an oxygen sensor of the sulphate-reducing bacterium Desulfovibrio vulgaris Hildenborough. Microbiology, 143, 1815–26.Google Scholar
Gardner, A. M., Helmick, R. A. and Gardner, P. R. (2002). Flavorubredoxin, an inducible catalyst for nitric oxide reduction and detoxification in Escherichia coli. J Biol Chem, 277, 8172–7.CrossRefGoogle ScholarPubMed
Gomes, C. M., Giuffrè, A., Forte, E.et al. (2002). A novel type of nitric oxide reductase: Escherichia coli flavorubredoxin, J Biol Chem, 277, 25273–6.CrossRefGoogle ScholarPubMed
Hatchikian, E. C. and Henry, Y. A. (1977). An iron-containing superoxide dismutase from the strict anaerobe Desulfovibrio desulfuricans (Norway 4). Biochimie, 59, 153–61.CrossRefGoogle Scholar
Hatchikian, C. E., LeGall, J. and Bell, G. R. (1977). Significance of superoxide dismutase and catalase activities in the strict anaerobes, sulphate-reducing bacteria. In Michael, A. M., McCord, J. M. and Fridovich, I. (eds.), Superoxide and superoxide dismutase. New York: Academic Press. pp. 159–72.Google 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, 554–9.CrossRefGoogle Scholar
Imlay, J. A. (2002a). How oxygen damages microbes: oxygen tolerance and obligate anaerobiosis. Adv Microb Physiol, 46, 111–53.CrossRefGoogle Scholar
Imlay, J. A. (2002b). What biological purpose is served by superoxide reductases?J Biol Inorg Chem, 7, 659–63.CrossRefGoogle Scholar
Imlay, J. A. (2003). Pathways of oxidative damage. Annu Rev Microbiol, 57, 395–418.CrossRefGoogle ScholarPubMed
Imlay, J. A. and Fridovich, I. (1991). Assay of metabolic superoxide production in Escherichia coli. J Biol Chem, 266, 6957–65.Google ScholarPubMed
Imlay, K. R. C. and Imlay, J. A. (1996). Cloning and analysis of sodC, encoding the copper-zinc superoxide dismutase of Escherichia coli. J Bacteriol, 178, 2564–71.CrossRefGoogle ScholarPubMed
Iyer, R. B., Silaghi-Dumitrescu, R., Kurtz, D. M. Jr. and Lanzilotta, W. N. (2005). High-resolution crystal structures of Desulfovibrio vulgaris (Hildenborough) nigerythrin: facile, redox-dependent iron movement, domain interface variability, and peroxidase activity in the rubrerythrins. J Biol Inorg Chem, 10, 407–16.CrossRefGoogle ScholarPubMed
Jin, S., Kurtz, D. M. Jr., Liu, Z.-J., Rose, J. and Wang, B.-C. (2002). X-ray crystal structures of reduced rubrerythrin and its azide adduct: a structure-based mechanism for a non-heme diiron peroxidase. J Am Chem Soc, 124, 9845–55.CrossRefGoogle ScholarPubMed
Justino, M. C., Vicente, J. B., Teixeira, M. and Saraiva, L. M. (2005). New genes implicated in the protection of anaerobically grown Escherichia coli against nitric oxide. J Biol Chem, 280, 2636–43.CrossRefGoogle ScholarPubMed
Kawasaki, S., Ishikura, J., Watamura, Y. and Niimura, Y. (2004). Identification of O2-induced peptides in an obligatory anaerobe, Clostridium acetobutylicum. FEBS Lett, 571, 21–5.CrossRefGoogle Scholar
Kawasaki, S., Watamura, Y., Ono, M.et al. (2005). Adaptive responses to oxygen stress in obligatory anaerobes Clostridium acetobutylicum and Clostridium aminovalericum. Appl Environ Microbiol, 71, 8442–50.CrossRefGoogle ScholarPubMed
Kirschvink, J. L., Gaidos, E. J., Bertani, L. E.et al. (2000). Paleoproterozoic snowball earth: extreme climatic and geochemical global change and its biological consequences. Proc Natl Acad Sci USA, 97, 1400–5.CrossRefGoogle ScholarPubMed
Kitamura, M., Mizugai, K., Taniguchi, M.et al. (1995). A gene encoding a cytochrome c oxidase-like protein is located closely to the cytochrome c-553 gene in the anaerobic bacterium, Desulfovibrio vulgaris (Miyazaki F). Microbiol Immunol, 39, 75–80.CrossRefGoogle Scholar
Kjeldsen, K. U., Joulian, C. and Ingvorsen, K. (2005). Effects of oxygen exposure on respiratory activities of Desulfovibrio desulfuricans strain DvO1 isolated from activated sludge. FEMS Microbiol Ecol, 53, 275–84.CrossRefGoogle ScholarPubMed
Krekeler, D., Sigalevich, P., Teske, A., Cypionka, H. and Cohen, Y. (1997). Sulphate-reducing bacterium form the oxic layer of a microbial mat from Solar Lake (Sinai), Desulfovibrio oxyclinae sp. nov. Arch Microbiol, 167, 369–75.CrossRefGoogle Scholar
Kumagai, H., Fujiwara, T., Matsubara, H. and Saeki, K. (1997). Membrane localization, topology, and mutual stabilization of the rnfABC gene products in Rhodobacter capsulatus and implications for a new family of energy-coupling NADH oxidoreductases. Biochemistry, 36, 5509–21.CrossRefGoogle ScholarPubMed
Kurtz, D. M. Jr. (2004). Microbial detoxification of superoxide: The non-heme iron reductive paradigm for combating oxidative stress. Acc Chem Res, 37, 902–8.CrossRefGoogle ScholarPubMed
LeGall, J., Prickril, B. C., Moura, I.et al. (1988). Isolation and characterization of rubrerythrin, a non-heme iron protein from Desulfovibrio vulgaris that contains rubredoxin centers and a hemerythrin-like binuclear iron cluster. Biochemistry, 27, 1636–42.CrossRefGoogle Scholar
Lemos, R. S., Gomes, C. M., LeGall, J., Xavier, A. V., Teixeira, M. (2002). The quinol:fumarate oxidoreductase from the sulphate reducing bacterium Desulfovibrio gigas: spectroscopic and redox studies. J Bioenerg Biomemb, 34, 21–30.CrossRefGoogle ScholarPubMed
Lemos, R. S., Gomes, C. M., Santana, M.et al. (2001). The ‘strict’ anaerobe Desulfovibrio gigas contains a membrane-bound oxygen-reducing respiratory chain. FEBS Lett, 496, 40–3.CrossRefGoogle ScholarPubMed
Liochev, S. I. and Fridovich, I. (1997). A mechanism for complementation of the sodA sodB defect in Escherichia coli by overproduction of the rbo gene product (desulfoferrodoxin) from Desulfoarculus baarsii. J Biol Chem, 272, 25573–5.CrossRefGoogle ScholarPubMed
Liochev, S. I. and Fridovich, I. (2000). Copper- and zinc-containing superoxide dismutase can act as a superoxide reductase and a superoxide oxidase. J Biol Chem, 275, 38482–5.CrossRefGoogle Scholar
Loewen, P. C., Klotz, M. G. and Hassett, D. J. (2000). Catalase – an “old” enzyme that continues to surprise us. ASM News, 66, 76–82.Google Scholar
Lombard, M., Fontecave, M., Touati, D. and Nivière, V. (2000). Reaction of the desulfoferrodoxin from Desulfoarculus baarsii with superoxide anion. Evidence for a superoxide reductase activity. J Biol Chem, 275, 115–21.CrossRefGoogle ScholarPubMed
Lumppio, H. L., Shenvi, N. V., Garg, R. P., Summers, A. O. and Kurtz, D. M. Jr. (1997). A rubrerythrin operon and nigerythrin gene in Desulfovibrio vulgaris (Hildenborough). J Bacteriol, 179, 4607–15.CrossRefGoogle Scholar
Lumppio, H. L., Shenvi, N. V., Summers, A. O., Voordouw, G. and Kurtz, D. M. Jr. (2001). Rubrerythrin and rubredoxin oxidoreductase in Desulfovibrio vulgaris. A novel oxidative stress protection system. J Bacteriol, 183, 101–8, and correction 2970.CrossRefGoogle ScholarPubMed
Macedo, S., Romão, C. V., Mitchell, E.et al. (2003). The nature of the diiron site in the bacterioferritin from Desulfovibrio desulfuricans. Nature Struct Biol, 10, 285–90.CrossRefGoogle Scholar
Marschall, C., Frenzel, P. and Cypionka, H. (1993). Influence of oxygen on sulphate reduction and growth of sulphate-reducing bacteria. Arch Microbiol, 159, 168–73.CrossRefGoogle Scholar
May, A., Hillmann, F., Riebe, O., Fischer, R. J. and Bahl, H. (2004). A rubrerythrin-like oxidative stress protein of Clostridium acetobutylicum is encoded by a duplicated gene and identical to the heat shock protein Hsp21. FEMS Microbiol Lett, 238, 249–54.Google ScholarPubMed
McCord, J. M., Keele, B. B. Jr. and Fridovich, I. (1971). An enzyme-based theory of obligate anaerobiosis: the physiological function of superoxide dismutase. Proc Natl Acad Sci USA, 68, 1024–7.CrossRefGoogle ScholarPubMed
Miller, A.-F. (2004). Superoxide Processing. In Que, Jr L.. and Tolman, W. B. (eds.), Comprehensive coordination chemistry II – from biology to nanotechnology, Oxford, UK: Elsevier. pp. 479–506.Google Scholar
Moura, I., Tavares, P., Moura, J. J.et al. (1990). Purification and characterization of desulfoferrodoxin. A novel protein from Desulfovibrio desulfuricans (ATCC 27774) and from Desulfovibrio vulgaris (strain Hildenborough) that contains a distorted rubredoxin center and a mononuclear ferrous center. J Biol Chem, 265, 21596–602.Google Scholar
Moura, I., Tavares, P. and Ravi, N. (1994). Characterization of 3 proteins containing multiple iron sites – rubrerythrin, desulfoferrodoxin, and a protein containing a six-iron cluster. Methods Enzymol, 243, 216–40.CrossRefGoogle Scholar
Nakanishi, T., Inoue, H. and Kitamura, M. (2003). Cloning and expression of the superoxide dismutase gene from the obligate anaerobic bacterium Desulfovibrio vulgaris (Miyazaki F). J Biochem (Tokyo), 133, 387–93.CrossRefGoogle Scholar
Park, S., You, X. and Imlay, J. A. (2005). Substantial DNA damage from submicromolar intracellular hydrogen peroxide detected in Hpx-mutants of Escherichia coli. Proc Natl Acad Sci USA, 102, 9317–22.CrossRefGoogle ScholarPubMed
Parsonage, D., Youngblood, D. S., Sarma, G. N.et al. (2005). Analysis of the link between enzymatic activity and oligomeric state in AhpC, a bacterial peroxiredoxin. Biochemistry, 44, 10583–92.CrossRefGoogle Scholar
Pereira, M. M. and Teixeira, M. (2004). Proton pathways, ligand binding and dynamics of the catalytic site haem-copper superfamily of oxygen reductases. Biochim Biophys Acta, 1655, 340–6.CrossRefGoogle Scholar
Pereira, M. M., Bandeiras, T. M., Fernandes, A. S.et al. (2004). Respiratory chains from aerobic thermophilic prokaryotes. J Bioenerg Biomemb, 36, 93–105.CrossRefGoogle ScholarPubMed
Pereira, M. M., Carita, J. N. and Teixeira, M. (1999). Membrane-bound electron transfer chain of the thermohalophilic bacterium Rhodothermus marinus: a novel multihemic cytochrome bc, a new complex III. Biochemistry, 38, 1268–75.CrossRefGoogle ScholarPubMed
Pereira, M. M., Santana, M. and Teixeira, M. (2001). A novel scenario for the evolution of haem-copper oxidases. Biochim Biophys Acta, 1505, 185–208.CrossRefGoogle Scholar
Pianzzola, M. J., Soubes, M. and Touati, D. (1996). Overproduction of the rbo gene product from Desulfovibrio species suppresses all deleterious effects of lack of superoxide dismutase in Escherichia coli. J Bacteriol, 178, 6736–42.CrossRefGoogle ScholarPubMed
Pierik, A. J., Wolbert, R. B. G., Portier, G. L., Verhagen, M. F. J. M. and Hagen, W. R. (1993). Nigerythrin and rubrerythrin from Desulfovibrio vulgaris each contain two mononuclear iron centers and two dinuclear iron clusters. Eur J Biochem, 212, 237–45.CrossRefGoogle ScholarPubMed
Pütz, S., Gelius-Dietrich, G., Piotrowski, M. and Henze, K. (2005). Rubrerythrin and peroxiredoxin: two novel putative peroxidases in the hydrogenosomes of the microaerophilic protozoon Trichomonas vaginalis. Mol Biochem Parasitol, 142, 212–23.CrossRefGoogle ScholarPubMed
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, 887–902.CrossRefGoogle Scholar
Ravenschlag, K., Sahm, K., Knoblauch, C., Jorgensen, B. B. and Amann, R. (2000). Community structure, cellular rRNA content, and activity of sulphate-reducing bacteria in marine arctic sediments. Appl Environ Microbiol, 66, 3592–602.CrossRefGoogle Scholar
Risatti, J. B., Capman, W. C. and Stahl, D. A. (1994). Community structure of a microbial mat: the phylogenetic dimension. Proc Natl Acad Sci USA, 91, 10173–7.CrossRefGoogle ScholarPubMed
Rodrigues, J. V., Abreu, I. A., Saraiva, L. M. and Teixeira, M. (2005). Rubredoxin acts as an electron donor for neelaredoxin in Archaeoglobus fulgidus. Biochem Biophys Res Commun, 329, 1300–5.CrossRefGoogle ScholarPubMed
Romão, C. V., Louro, R., Timkovich, R.et al. (2000a). Iron-coproporphyrin III is a natural cofactor in bacterioferritin from the anaerobic bacterium Desulfovibrio desulfuricans. FEBS Lett, 480, 213–16.CrossRefGoogle Scholar
Romão, C. V., Regalla, M., Xavier, A. V.et al. (2000b). A bacterioferritin from the strict anaerobe Desulfovibrio desulfuricans ATCC 27774. Biochemistry, 39, 6841–9.CrossRefGoogle Scholar
Salvador, A., Sousa, J. and Pinto, R. E. (2001). Hydroperoxyl, superoxide and pH gradients in the mitochondrial matrix: a theoretical assessment. Free Radic Biol Med, 31, 1208–15.CrossRefGoogle ScholarPubMed
Saraiva, L. M., Vicente, J. B. and Teixeira, M. (2004). The role of the flavodiiron proteins in microbial nitric oxide detoxification. Advances in Microbial Physiology, 49, 77–129.CrossRefGoogle ScholarPubMed
Saraste, M. (1999). Oxidative phosphorylation at the fin de siècle. Science, 283, 1488–93.CrossRefGoogle ScholarPubMed
Sass, H., Berchtold, M., Branke, J.et al. (1998). Psychrotolerant sulphate-reducing bacteria from an oxic freshwater sediment, description of Desulfovibrio cuneatus sp. nov. and Desulfovibrio litoralis sp. nov. Syst Appl Microbiol, 21, 212–19.CrossRefGoogle Scholar
Sass, H., Cypionka, H. and Babenzien, H.-D. (1997). Vertical distribution of sulphate-reducing bacteria at the oxic–anoxic interface in sediments of the oligotrophic Lake Stechlin. FEMS Microbiol Ecol, 22, 245–55.CrossRefGoogle Scholar
Seaver, L. C. and Imlay, J. A. (2001). Hydrogen peroxide fluxes and compartmentalization inside growing Escherichia coli. J Bacteriol, 183, 7182–19.CrossRefGoogle ScholarPubMed
Sigalevich, P. and Cohen, Y. (2000). Oxygen dependent growth of the sulphate-reducing bacterium Desulfovibrio oxyclinae in coculture with Marinobacter sp. strain MB in a aerated sulphate-depleted chemostat. Appl Environ Microbiol, 66, 5019–23.CrossRefGoogle Scholar
Silaghi-Dumitrescu, R., Coulter, E. D., Das, A.et al. (2003). A flavodiiron protein and high molecular weight rubredoxin from Moorella thermoacetica with nitric oxide reductase activity. Biochemistry, 42, 2806–15.CrossRefGoogle ScholarPubMed
Silaghi-Dumitrescu, R., Ng, K. Y., Viswanathan, R. and Kurtz, D. M. Jr. (2005). A flavodiiron protein from Desulfovibrio vulgaris with oxidase and nitric oxide reductase activities. Evidence for an in vivo nitric oxide scavenging function. Biochemistry, 44, 3572–9.CrossRefGoogle Scholar
Silva, G., LeGall, J., Xavier, A. V., Teixeira, M. and Rodrigues-Pousada, C. (2001). Molecular characterization of Desulfovibrio gigas neelaredoxin, a protein involved in oxygen detoxification in anaerobes. J Bacteriol, 183, 4413–20.CrossRefGoogle ScholarPubMed
Srinivasan, V., Rajendran, C., Sousa, F. L.et al. (2005). Structure at 1.3A resolution of Rhodothermus marinus caa3 cytochrome c domain. J Mol Biol, 345, 1047–57.CrossRefGoogle Scholar
Storz, G. and Imlay, J. A. (1999). Oxidative stress. Curr Opin Microbiol, 2, 188–94.CrossRefGoogle ScholarPubMed
Sztukowska, M., Bugno, M., Potempa, J., Travis, J. and Kurtz, D. M. Jr. (2002). Role of rubrerythrin in the oxidative stress response of Porphyromonas gingivalis. Mol Microbiol, 44, 479–88.CrossRefGoogle ScholarPubMed
Teske, A., Ramsing, N. B., Habicht, K.et al. (1998). Sulphate-reducing bacteria and their activities in cyanobacterial mats of Solar Lake (Sinai, Egypt). Appl Env Microbiol, 64, 2943–51.Google Scholar
Vliet, A. H., Baillon, M. L., Penn, C. W. and Ketley, J. M. (1999). Campylobacter jejuni contains two Fur homologs: characterization of iron-responsive regulation of peroxide stress defense genes by the PerR repressor. J Bacteriol, 181, 6371–6.Google ScholarPubMed
Vance, C. K. and Miller, A. F. (1998). A simple proposal that can explain the inactivity of metal-substituted superoxide dismutases. J Am Chem Soc, 120, 461–7.CrossRefGoogle Scholar
Voordouw, G. (2002). Carbon monoxide cycling by Desulfovibrio vulgaris Hildenborough. J Bacteriol, 184, 5903–11.CrossRefGoogle 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, 7888–95.CrossRefGoogle ScholarPubMed
Wood, Z. A., Schröder, E., Robin Harris, J. and Poole, L. B. (2003). Structure, mechanism and regulation of peroxiredoxins. Trends Biochem Sci, 28, 32–40.CrossRefGoogle ScholarPubMed
Yoder, D. W., Hwang, J., and Penner-Hahn, J. E. (2000). Manganese catalases. In Sigel, A. and Sigel, H.. (eds.), Metal Ions in Biological Systems, Vol. 37. New York:Marcel Dekker, Inc. pp. 527–57.Google Scholar
Zámocky, M. and Koller, F. (1999). Understanding the structure and function of catalases: clues from molecular evolution and in vitro mutagenesis. Prog Biophys Mol Biol, 72, 19–66.CrossRefGoogle ScholarPubMed
Zou, P.-J., Borovok, I., Ortiz de Orué Lucana, D., Müller, D. and Schrempf, H. (1999). The mycelium-associated Streptomyces reticuli catalase-peroxidase, its gene and regulation by FurS. Microbiology, 145, 549–59.CrossRefGoogle 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
×