Hostname: page-component-848d4c4894-8bljj Total loading time: 0 Render date: 2024-06-20T04:10:38.314Z Has data issue: false hasContentIssue false

Growth, nitrogen fixation, respiration, and a nitrogen budget for cultures of a cosmopolitan diazotrophic endosymbiont (Teredinibacter turnerae) of shipworms

Published online by Cambridge University Press:  07 November 2013

Rachel E.A. Horak*
Affiliation:
Georgia Institute of Technology, School of Biology, 311 Ferst Drive, Atlanta, GA 30332, USA
Joseph P. Montoya
Affiliation:
Georgia Institute of Technology, School of Biology, 311 Ferst Drive, Atlanta, GA 30332, USA
*
Correspondence should be addressed to: R.E.A. Horak, University of Washington, School of Oceanography, Box 357940, Seattle, WA 98195, USA email: rahorak@gmail.com

Abstract

Wood-boring bivalves (Bivalvia, family Teredinidae), also known as shipworms, host dinitrogen-fixing and cellulolytic symbiotic bacteria in gill bacteriocytes, which may be a necessary adaptation to a wooden diet. Although oxygen (O2) inhibits nitrogenase in other species, symbionts are able to fix nitrogen (N) within the gill tissue and provide newly fixed N to the host shipworm. The recent direct evidence of new N incorporation into the host tissue indicates that there are potentially complex nutrient cycles in this symbiosis and uninvestigated controls upon these cycles.To elucidate the mechanisms of this unique N2-fixing symbiosis and determine whether symbionts can excrete newly fixed N, we measured rates of growth, N2-fixation, respiration, and inorganic N content for the cultivated symbiont Teredinibacter turnerae (γ-proteobacteria, strain T7901) under a range of headspace O2 conditions. In all conditions, headspace O2 did not affect maximum specific N2-fixation and respiration activity, but did influence the rate and timing of growth. These results are consistent with the development of microaerobic conditions through an oxygen gradient in the culture medium, which facilitates N2-fixation and growth. The medium accumulated a small amount of NH4+, which represented 0.5–2.5% of the total N fixed by the culture. We constructed a simple N budget for T. turnerae to assess the role of the major known N sources and sinks. The N budget was not closed, indicating that new N is allocated to currently unidentified sinks, which may include excreted dissolved organic nitrogen.

Type
Research Article
Copyright
Copyright © Marine Biological Association of the United Kingdom 2013 

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

REFERENCES

Beshay, U. (2003) Production of alkaline protease by Teredinobacter turnirae cells immobilized in Ca-alginate beads. African Journal of Biotechnology 2, 6065.Google Scholar
Bradford, M. (1976) A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principal of protein-dye binding. Analytical Biochemistry 72, 248254.CrossRefGoogle Scholar
Brune, A. and Friedrich, M. (2000) Microecology of the termite gut: structure and function on a microscale. Current Opinion in Microbiology 3, 263269.CrossRefGoogle ScholarPubMed
Capone, D. (1994) Amino acid cycling in colonies of the planktonic marine cyanobacterium Trichdesmium thiebautii. Applied and Environmental Microbiology 60, 39893995.CrossRefGoogle ScholarPubMed
Capone, D. and Montoya, J.P. (2001) Nitrogen fixation and denitrification. Methods in Microbiology 30, 501515.CrossRefGoogle Scholar
Delatolla, R., Berk, D. and Tufenkji, N. (2008) Rapid and reliable quantification of biofilm weight and nitrogen content of biofilm attached to polystyrene beads. Water Research 42, 30823088.CrossRefGoogle ScholarPubMed
Distel, D. (2003) The biology of marine wood-boring bivalves and their bacterial endosymbionts. In Goodell, B., Nicholas, D. and Schultz, T. (eds) Wood deterioration and preservation. Washington, DC: American Chemical Society, pp. 253271.CrossRefGoogle Scholar
Distel, D., DeLong, E. and Waterbury, J. (1991) Phylogenetic characterization and in situ localization of the bacterial symbiont of shipworms (Teredinidae: Bivalvia) by using 16S rRNA sequence analysis and oligodeoxynucleotide probe hybridization. Applied and Environmental Microbiology 57, 23762382.CrossRefGoogle ScholarPubMed
Distel, D., Morrill, N., MacLaren-Toussaint, N., Franks, D. and Waterbury, J. (2002) Teredinibacter turnerae gen. nov., sp. nov., a dinitrogen-fixing, cellulolytic, endosymbiotic gamma-proteobacterium isolated from the gills of wood-boring molluscs (Bivalvia: Tereinidae). International Journal of Systematics and Evolutionary Microbiology 52, 22612269.Google Scholar
Dixon, R. and Kahn, D. (2004) Genetic regulation of biological nitrogen fixation. Nature Reviews Microbiology 2, 621631.CrossRefGoogle ScholarPubMed
Ferreira, G., Ahuja, S., Sierks, M. and Moreira, A. (2001) Pleomorphism of the marine bacterium Teredinobacter turnirae. Letters in Applied Microbiology 33, 5660.CrossRefGoogle ScholarPubMed
Foster, R., Kuypers, M., Vagner, T., Paerl, R., Musat, N. and Zehr, J. (2011) Nitrogen fixation and transfer in open ocean diatom-cyanobacterial symbioses. ISME Journal 5, 14841493.CrossRefGoogle ScholarPubMed
Gallager, S., Turner, R. and Berg, C.J. Jr (1981) Physiological aspects of wood consumption, growth, and reproduction in the shipworm Lyrodus pedicellatus (Bivalvia: Teredinidae). Journal of Experimental Marine Biology and Ecology 52, 6377.CrossRefGoogle Scholar
Garcia, H. and Gordon, L. (1992) Oxygen solubility in seawater: better fitting equations. Limnology and Oceanography 37, 13071312.CrossRefGoogle Scholar
Glibert, P. and Bronk, D. (1994) Release of dissolved organic nitrogen by marine diazotrophic cyanobacteria, Trichodesmium spp. Applied and Environmental Microbiology 60, 39964000.CrossRefGoogle ScholarPubMed
Greene, R. (1994) Challenges from the sea: marine shipworms and their symbiotic bacterium. Society of Industrial Microbiology News 44, 5159.Google Scholar
Greene, R., Cotta, M. and Griffin, H. (1989) A novel, symbiotic bacteria isolated from marine shipworm secretes proteolytic activity. Current Microbiology 19, 353356.CrossRefGoogle Scholar
Greene, R. and Freer, S. (1986) Growth characteristics of a novel nitrogen-fixing cellulolytic bacterium. Applied and Environmental Microbiology 52, 982986.CrossRefGoogle ScholarPubMed
Griffin, H., Greene, R. and Cotta, M. (1992) Isolation and characterization of an alkaline protease from the marine shipworm bacterium. Current Microbiology 24, 111117.CrossRefGoogle Scholar
Großkopf, T. and LaRoche, J. (2012) Direct and indirect costs of nitrogen fixation in Crocosphaera watsonii WH8501 and possible implications for the nitrogen cycle. Frontiers in Microbiology 3, article 236.CrossRefGoogle ScholarPubMed
Hill, S. (1988) How is nitrogenase regulated by oxygen? FEMS Microbiology Reviews 54, 111130.CrossRefGoogle Scholar
Holl, C. and Montoya, J. (2008) Diazotrophic growth of the marine cyanobacterium Trichodesmium IMS101 in continuous culture: effects of growth rate on N2-fixation rate, biomass, and C:N:P stoichiometry. Journal of Phycology 44, 929937.CrossRefGoogle ScholarPubMed
Kuhla, J., Dingler, C. and Oelze, J. (1985) Production of extracellular nitrogen-containing components by Azotobacter vinelandii fixing dinitrogen in oxygen-controlled continuous culture. Archives of Microbiology 141, 297302.CrossRefGoogle Scholar
Lechene, C., Luyten, Y., McMahon, G. and Distel, D. (2007) Quantitative imaging of nitrogen fixation by individual bacteria within animal cells. Science 317, 15631566.CrossRefGoogle ScholarPubMed
Lechene, C., Hillion, F., McMahon, G., Benson, D., Kleinfeld, A., Kampf, J., Distel, D., Luyten, Y., Bonventre, J., Hentschel, D., Park, K., Ito, S., Schwart, M., Begichou, G. and Slodzian, G. (2006) High-resolution quantitative imaging of mammalian and bacterial cells using stable isotope mass spectrometry. Journal of Biology 5, 20.CrossRefGoogle ScholarPubMed
Loladze, I. and Elser, J. (2011) The origins of the Redfield nitrogen-to-phosphorus ratio are in a homeostatic protein-to-rRNA ratio. Ecology Letters 14, 244250.CrossRefGoogle Scholar
Marchal, K. and Vanderleyden, J. (2000) The ‘oxygen paradox’ of dinitrogen-fixing bacteria. Biology and Fertility of Soils 30, 363373.CrossRefGoogle Scholar
Meeks, J., Enderlin, C., Joseph, C., Chapman, J. and Lollar, M. (1985) Fixation of [13N]N2 and transfer of fixed nitrogen in the Anthoceros–Nostoc symbiotic association. Planta 164, 406414.CrossRefGoogle ScholarPubMed
Mulholland, M. and Bernhardt, P. (2005) The effect of growth rate, phosphorus concentration, and temperature on N2 fixation, carbon fixation, and nitrogen release in continuous cultures of Trichodesmium IMS101. Limnology and Oceanography 50, 839849.CrossRefGoogle Scholar
Mulholland, M. and Capone, D. (1999) Nitrogen fixation, uptake, and metabolism in natural and cultured populations of Trichodesmium spp. Marine Ecology Progress Series 188, 3349.CrossRefGoogle Scholar
Neidhardt, F.C. and Umbarger, H. (1996) Chemical composition of Escherichia coli. In Neidhardt, F.C. (ed) Escherichia coli and salmonella: cellular and molecular biology, Volume 1. 2nd edition. Washington, DC: American Society of Microbiology Press, pp. 13–16.Google Scholar
Ohkuma, M. (2001) Symbiosis within the gut microbial community of termites. RIKEN Review 41, 6972.Google Scholar
Peters, G. (1977) The Azolla-Anabaena azollae symbiosis. In Hollaender, A., Burns, R., Day, P., Hardy, R., Helinski, D., Lamborg, M., Owens, L. and Volentine, R. (eds) Genetic engineering for nitrogen fixation. New York: Plenum Press, pp. 231257.CrossRefGoogle Scholar
Silvester, W., Parsons, R. and Watt, P. (1996) Direct measurement of release and assimilation of ammonia in the Gunnera–Nostoc symbiosis. New Phytologist 132, 617625.CrossRefGoogle ScholarPubMed
Sterner, R. and Elser, J. (2002) Ecological stoichiometry: the biology of elements from molecules to the biosphere. Princeton, NJ: Princeton University Press.Google Scholar
Streams, M., Fisher, C. and Fiala-Medioni, A. (1997) Methanotrophic symbiont location and fate of carbon incorportated from methane in a hydrocarbon seep mussel. Marine Biology 129, 465476.CrossRefGoogle Scholar
Strickland, J. and Parsons, T. (1972) A practical handbook of seawater analysis. 2nd edition. Ottawa: Bulletin of Fisheries Research Board of Canada.Google Scholar
Trindade-Silva, A., Machado-Forrera, E., Senra, M., Vizzoni, V., Ypanaguine, L., Leoncini, O. and Soares, C. (2009) Physiological traits of the symbiotic bacterium Teredinibacter turnerae isolated from the mangrove shipworm Neoteredo reynei. Genetics and Molecular Biology 32, 572581.CrossRefGoogle ScholarPubMed
Trytek, R. and Allen, W. (1980) Synthesis of essential amino acids by bacterial symbionts in the gills of the shipworm Bankia setacea (Tryon). Comparative Biochemistry and Physiology 67A, 419427.Google Scholar
Turner, R. (1984) An overview of research on marine borers: past progress and future direction. In Costlow, J. and Tipper, R. (eds) Marine biodeterioration: an interdisciplinary study. Annapolis, MD: Naval Institute Press, pp. 316.CrossRefGoogle Scholar
Van Dover, C. (2000) The ecology of deep-sea hydrothermal vents. Princeton, NJ: Princeton University Press.CrossRefGoogle Scholar
Waterbury, J., Calloway, C. and Turner, R. (1983) A cellulolytic nitrogen-fixing bacterium cultured from the gland of Deshayes in shipworms (Bivalvia: Teredinidae). Science 221, 14011403.CrossRefGoogle ScholarPubMed
Waters, J., Hughes II, B., Purcell, L., Gerhardt, K., Mawhinney, T. and Emerich, D. (1998) Alanine, not ammonia, is excreted from N2-fixing soybean nodule bacteroids. Proceedings of the National Academy of Sciences of the United States of America 95, 1203812042.CrossRefGoogle Scholar