Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-26T17:26:25.635Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 March 2013

Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Wetland Ecology
Principles and Conservation
, pp. 427 - 475
Publisher: Cambridge University Press
Print publication year: 2010

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

Abraham, K. F. and Keddy, C. J. (2005). The Hudson Bay Lowland. In The World's Largest Wetlands: Ecology and Conservation, eds. Fraser, L. H. and Keddy, P. A., pp. 118–48. Cambridge, UK: Cambridge University Press.Google Scholar
Adam, P. (1990). Saltmarsh Ecology. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Adams, G. D. (1988). Wetlands of the prairies of Canada. In Wetlands of Canada, National Wetlands Working Group, Ecological Land Classification Series No. 24, pp. 158–98. Montreal, QC: Polyscience Publications for Sustainable Development Branch, Environment Canada.Google Scholar
Adamus, P. R. (1992). Choices in monitoring wetlands. In Ecological Indicators, eds. McKenzie, D. H., Hyatt, D. E., and McDonald, V. J., pp. 571–92. London: Elsevier.CrossRefGoogle Scholar
Adamus, P. R. (1996). Bioindicators for Assessing Ecological Integrity of Prairie Wetlands, EPA/600/R-96/082. Corvallis, OR: U.S. Environmental Protection Agency, National Health and Environmental Effects Research Laboratory, Western Ecology Division.Google Scholar
Adamus, P. R. and Stockwell, L. T. (1983). A Method for Wetland Functional Assessment, Vol. 1 Critical Review and Evaluation Concepts, Report No. FHA-PI-82-23, and Vol. 2 Federal Highway Administration Assessment Method, Report No. FHA-PI-82-24. Springfield, VA: National Technical Information Service.Google Scholar
Adamus, P. R., , ARA Inc., Clairain, E. J., Smith, R. D., and Young, R. E. (1987). Wetland Evaluation Technique (WET), Vol. 2, Methodology. Vicksburg, MS: U.S. Army Corps of Engineers.Google Scholar
Aerts, R. and Berendse, F. (1988). The effect of increased nutrient availability on vegetation dynamics in wet heathlands. Vegetatio, 76, 63–9.Google Scholar
Agrawala, S., Ota, T., Ahmed, A. U., Smith, J., and Aalst, M. (2003). Development and Climate Change in Bangladesh: Focus on Coastal Flooding and The Sundarbans. Paris: Environment Directorate, OECD.Google Scholar
Agren, G. I. and Fagerstrom, T. (1984). Limiting dissimilarity in plants: randomness prevents exclusion of species with similar competitive abilities. Oikos, 43, 369–75.CrossRefGoogle Scholar
Alestalio, J. and Haikio, J.. (1979). Forms created by the thermal movement of lake ice in Finland in winter 1972–73. Fennia, 157, 51–92.Google Scholar
Alho, C. J. R. (2005). The Pantanal. In The World's Largest Wetlands: Ecology and Conservation, eds. Fraser, L. H. and Keddy, P. A., pp. 271–303. Cambridge, UK: Cambridge University Press.Google Scholar
Alho, C. J. R., Lacher, T. E., Jr., and Goncalves, H. C. (1988). Environmental degradation in the Pantanal ecosystem. BioScience, 38, 164–71.CrossRefGoogle Scholar
Allison, M. A. (1998). Historical changes in the Ganges–Brahmaputra delta front. Journal of Coastal Research, 14, 1269–75.Google Scholar
Allison, S. K. (1995). Recovery from small-scale anthropogenic disturbances by northern California salt marsh plant assemblages. Ecological Applications, 5, 693–702.CrossRefGoogle Scholar
Anderson, R. C., Liberta, A. E., and Dickman, L. A. (1984). Interaction of vascular plants and vesicular–arbuscular mycorrhizal fungi across a soil moisture–nutrient gradient. Oecologia, 64, 111–17.CrossRefGoogle ScholarPubMed
Anthoni, J. F. (2006). The chemical composition of seawater. www.seafriends.org.nz/oceano/seawater.htm. (accessed June 4, 2008)
Archibold, O. W. (1995). Ecology of World Vegetation. London: Chapman and Hall.CrossRefGoogle Scholar
Aresco, M. J. (2004). Highway mortality of turtles and other herpetofauna at Lake Jackson, Florida, USA, and the efficacy of a temporary fence/culvert system to reduce roadkills. In Proceedings of the 2003 International Conference on Ecology and Transportation, eds. Irwin, C. L., Garrett, P., and McDevmott, K. P., pp. 433–49. Raleigh, NC: Center for Transportation and the Environment, North Carolina State University.Google Scholar
Armentano, T. V. and Verhoeven, J. T. A. (1990). Biogeochemical cycles: global. In Wetlands and Shallow Continental Water Bodies, Vol. 1, Natural and Human Relationships, ed. Patten, B. C., pp. 281–311. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Armstrong, W., Armstrong, J., Beckett, P. M. and Justin, S. H. F. W. (1991). Convective gas-flows in wetland plant aeration. In Plant Life under Oxygen Deprivation, eds. Jackson, M. B., Davies, D. D., and Lambers, H., pp. 283–302. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Armstrong, J., Armstrong, W. and Beckett, P. M.. (1992). Phragmites australis: Venturi- and humidity-induced pressure flows enhance rhizome aeration and rhizosphere oxidation. New Phytologist, 120, 197–207.CrossRefGoogle Scholar
Arnold, S. J. (1972). Species densities of predators and their prey. The American Naturalist, 106, 220–35.CrossRefGoogle Scholar
Arnold, T. W. and Frytzell, E. K. (1990). Habitat use by male mink in relation to wetland characteristics and avian prey abundances. Canadian Journal of Zoology, 68, 2205–8.CrossRefGoogle Scholar
Arrhenius, O. (1921). Species and area. Journal of Ecology, 9, 95–9.CrossRefGoogle Scholar
Arroyo, M. T. K., Pliscoff, P., Mihoc, M., and Arroyo-Kalin, M. (2005). The Magellanic moorland. In The World's Largest Wetlands, eds. Fraser, L. H. and Keddy, P. A., pp. 424–45. Cambridge, UK: Cambridge University Press.Google Scholar
Aselman, I. and Crutzen, P. J. (1989). Global distribution of natural freshwater wetlands and rice paddies, their net primary productivity, seasonality and possible methane emissions. Journal of Atmospheric Chemistry, 8, 307–58.CrossRefGoogle Scholar
Atwood, E. L. (1950). Life history studies of the nutria, or coypu, in coastal Louisiana. Journal of Wildlife Management, 14, 249–65.CrossRefGoogle Scholar
Auclair, A. N. D., Bouchard, A. and Pajaczkowski, J. (1976a). Plant standing crop and productivity relations in a Scirpus–Equisetum wetland. Ecology, 57, 941–52.CrossRefGoogle Scholar
Auclair, A. N. D., Bouchard, A. and Pajaczkowski, J. (1976b). Productivity relations in a Carex-dominated ecosystem. Oecologia, 26, 9–31.CrossRefGoogle Scholar
Austin, M. P. (1982). Use of a relative physiological performance value in the prediction of performance in multispecies mixtures from monoculture performance. Journal of Ecology, 70, 559–70.CrossRefGoogle Scholar
Austin, M. P., Pausas, J. G., and Nicholls, A. O. (1996). Patterns of tree species richness in relation to environment in southeastern New South Wales, Australia. AustralianJournal of Ecology, 21, 154–64.Google Scholar
Bacon, P. R. (1978). Flora and Fauna of the Caribbean. Trinidad: Key Caribbean Publications.Google Scholar
Baedke, S. J. and Thompson, T. A.. (2000). A 4700-year record of lake level and isostasy for Lake Michigan. Journal of Great Lakes Research, 26, 416–26.CrossRefGoogle Scholar
Bakker, J. P. (1985). The impact of grazing on plant communities, plant populations and soil conditions on salt marshes. Vegetatio, 62, 391–8.CrossRefGoogle Scholar
Bakker, S. A., Jasperse, C. and Verhoeven, J. T. A. (1997). Accumulation rates of organic matter associated with different successional stages from open water to carr forest in former turbaries. Plant Ecology, 129, 113–20.CrossRefGoogle Scholar
Baldwin, A. H. and Mendelssohn, I. A. (1998a). Response of two oligohaline marsh communities to lethal and nonlethal disturbance. Oecologia, 116, 543–555.CrossRefGoogle ScholarPubMed
Baldwin, A. H. and Mendelssohn, I. A. (1998b). Effects of salinity and water level on coastal marshes: an experimental test of disturbance as a catalyst for vegetation change. Aquatic Botany, 61, 255–68.CrossRefGoogle Scholar
Baldwin, A. H., McKee, K. L., and Mendelssohn, I. A. (1996). The influence of vegetation, salinity and inundation of seedbanks of oligohaline coastal marshes. American Journal of Botany, 83, 470–9.CrossRefGoogle Scholar
Ball, P. J. and Nudds, T. D. (1989). Mallard habitat selection: an experiment and implications for management. In Freshwater Wetlands and Wildlife, eds. Sharitz, R. R., and Gibbons, J. W., pp. 659–71. US Department of Energy. Proceedings of a Symposium held at Charleston, South Carolina, March 24–27, 1986. Washington, DC:U.S. Department of Energy.Google Scholar
Barbour, C. D. and Brown, J. H. (1974). Fish species diversity in lakes. The American Naturalist, 108, 473–89.CrossRefGoogle Scholar
Bardecki, M. J., Bond, W. K., and Manning, E. W. (1989). Assessing Greenock Swamp: functions benefits and values. In Wetlands: Inertia or Momentum?, pp. 235–44. Conference Proceedings, Oct 21–22. Toronto, ON: Federation of Ontario Naturalists.Google Scholar
Barko, J. W. and Smart, R. M. (1978). The growth and biomass distribution of two emergent freshwater plants, Cyperus esculentus and Scirpus validus, on different sediments. Aquatic Botany, 5, 109–17.CrossRefGoogle Scholar
Barko, J. W. and Smart, R. M. (1979). The nutritional ecology of Cyperus esculentus, an emergent aquatic plant, grown on different sediments. Aquatic Botany, 6, 13–28.CrossRefGoogle Scholar
Barko, J. W. and Smart, R. M. (1980). Mobilization of sediment phosphorus by submersed freshwater macrophytes. Freshwater Biology, 10, 229–38.CrossRefGoogle Scholar
Barnard, J. R. (1978). Externalities from urban growth: the case of increased storm runoff and flooding. Land Economics, 54, 298–315.CrossRefGoogle Scholar
Barry, J. M. (1997). Rising Tide: The Great Mississippi Flood of 1927 and How It Changed America. New York: Simon and Schuster.Google Scholar
Barthelemy, A. (1874). De la respiration et de la circulation des gaz dans les végétaux. Annales des Sciences Naturelles Botaniques, 19, 131–75.Google Scholar
Bartram, W. (1791). Travels through North & South Carolina, Georgia, East & West Florida, the Cherokee Country, the Extensive Territories of the Muscogulges, or Creek Confederacy, and the Country of the Chactaws: Containing an Account of the Soil and Natural Productions of These Regions, Together with Observations on the Manners of the Indians. Philadelphia, PA: James and Johnson. (Digital edition, 2001, in Documenting the South, Chapel Hill, NC: University of North Carolina.)Google Scholar
Batt, B. D. J., Anderson, M. G., Anderson, C. D., and Caswell, F. D. (1989). The use of prairie potholes by North American ducks. In Northern Prairie Wetlands, ed. Valk, A. G., pp. 204–27. Ames, IA: Iowa State University Press.Google Scholar
Bauder, E. T. (1989). Drought stress and competition effects on the local distribution of Pogogyne abramsii. Ecology, 70, 1083–9.CrossRefGoogle Scholar
Bazely, D. R. and Jefferies, R. L. (1989). Lesser snow geese and the nitrogen economy of a grazed salt marsh. Journal of Ecology, 77, 24–34.CrossRefGoogle Scholar
Bazilevich, N. I., Rodin, L. Y., and Rozov, N. N. (1971). Geophysical aspects of biological productivity. Soviet Geography, Review and Translations, 12, 293–317.CrossRefGoogle Scholar
Beanland, G. E. and Duinker, P. N. (1983). An Ecological Framework for Environmental Impact Assessment in Canada. Halifax, NS: Institute for Resource and Environmental Studies, Dalhousie University, and Federal Environmental Assessment Review Office.Google Scholar
Beard, J. S. (1949). The Natural Vegetation of the Windward and Leeward Islands. Oxford, UK: Clarendon Press.Google Scholar
Bedford, B. L. (1996). The need to define hydrologic equivalence at the landscape scale for freshwater wetland mitigation. Ecological Applications, 6, 57–68.CrossRefGoogle Scholar
Bedford, B. L. and Preston, E. M. (1988). Developing the scientific basis for assessing cumulative effects of wetland loss and degradation on landscape functions: status, perspectives and prospects. Environmental Management, 12, 751–71.CrossRefGoogle Scholar
Beebee, T. J. C. (1996). Ecology and Conservation of Amphibians. London: Chapman and Hall.Google Scholar
Beeftink, W. G. (1977). The coastal salt marshes of western and northern Europe: an ecological and phytosociological approach. In Wet Coastal Ecosystems, ed. Chapman, V. J., pp. 109–55. Amsterdam, the Netherlands: Elsevier.Google Scholar
Begin, Y., Arseneault, S., and Lavoie, J. (1989). Dynamique d'une bordure forestière par suite de la hausse récente du niveau marin, rive sud-ouest du Golfe du Saint-Laurent, Nouveau-Brunswick. Géographie physique et Quaternaire, 43, 355–66.CrossRefGoogle Scholar
Belanger, L. and Bedard, J. (1994). Role of ice scouring and goose grubbing in marsh plant dynamics. Journal of Ecology, 82, 437–45.CrossRefGoogle Scholar
Belkin, D. A. (1963). Anoxia: tolerance in reptiles. Science, 139, 492–3.CrossRefGoogle ScholarPubMed
Bender, E. A, Case, T. J., and Gilpin, M. E. (1984). Perturbation experiments in community ecology: theory and practice. Ecology, 65, 1–13.CrossRefGoogle Scholar
Benson, L. (1959). Plant Classification. Lanham, MD: Lexington Books.Google Scholar
Berenbaum, M. R. (1991). Coumarins. In Herbivores: Their Interactions with Secondary Plant Metabolites, eds. Rosenthal, G. A. and Berenbaum, M. R., pp. 221–49. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Berendse, F. and Aerts, R. (1987). Nitrogen-use efficiency: a biologically meaningful definition?Functional Ecology, 1, 293–6.Google Scholar
Bernatowicz, S. and Zachwieja, J. (1966). Types of littoral found in the lakes of the Masurian and Suwalki Lakelands. Komitet Ekolgiezny-Polska Akademia Nauk, 14, 519–45.Google Scholar
Bertness, M. D. (1991). Interspecific interactions among high marsh perennials in a New England salt marsh. Ecology, 72, 125–37.CrossRefGoogle Scholar
Bertness, M. D. and Ellison, A. E. (1987). Determinants of pattern in a New England salt marsh plant community. Ecological Monographs, 57, 12–147.CrossRefGoogle Scholar
Bertness, M. D. and Hacker, S. D. (1994). Physical stress and positive associations among marsh plants. The American Naturalist, 144, 363–72.CrossRefGoogle Scholar
Bertness, M. D. and Leonard, G. H. (1997). The role of positive interactions in communities: lessons from intertidal habitats. Ecology, 78, 1976–89.CrossRefGoogle Scholar
Bertness, M. D. and Shumway, S. W. (1993). Competition and facilitation in marsh plants. The American Naturalist, 142, 718–34.CrossRefGoogle ScholarPubMed
Bertness, M. D. and Yeh, S. M. (1994). Cooperative and competitive interactions in the recruitment of marsh elders. Ecology, 75, 2416–29.CrossRefGoogle Scholar
Bertness, M. D., Gough, L., and Shumway, S. W. (1992a). Salt tolerances and the distribution of fugitive salt marsh plants. Ecology, 73, 1842–51.CrossRefGoogle Scholar
Bertness, M. D., Wikler, K., and Chatkupt, T. (1992b). Flood tolerance and the distribution of Iva frutescens across New England salt marshes. Oecologia, 91, 171–8.CrossRefGoogle ScholarPubMed
Best, E. P. H., Verhoeven, J. T. A., and Wolff, W. J. (1993). The ecology of The Netherlands wetlands: characteristics, threats, prospects and perspectives for ecological research. Hydrobiologia, 265, 305–20.CrossRefGoogle Scholar
Bethke, R. W. and Nudds, T. D. (1993). Variation in the diversity of ducks along a gradient of environmental variability. Oecologia, 93, 242–50.CrossRefGoogle ScholarPubMed
Biesterfeldt, J. M., Petranka, J. W., and Sherbondy, S. (1993). Prevalence of chemical interference competition in natural populations of wood frogs, Rana sylvatica. Copeia, 3, 688–95.CrossRefGoogle Scholar
Bilby, R. E., and Ward, J. (1991). Characteristics and function of large woody debris in streams draining old-growth, clear-cut, and 2nd-growth forests in southwestern Washington. Canadian Journal of Fisheries and Aquatic Sciences, 48, 2499–508.CrossRefGoogle Scholar
Binford, M. W., Brenner, M., Whitmore, T. J., Higuera-Gundy, , Deevey, A., , E. S., and Leyden, B. (1987). Ecosystems, paleoecology and human disturbance in subtropical and tropical America. Quaternary Scientific Review, 6, 115–28.Google Scholar
Bliss, L. C. and Gold, W. G. (1994). The patterning of plant communities and edaphic factors along a high arctic coastline: implications for succession. Canadian Journal of Botany, 72, 1095–107.CrossRefGoogle Scholar
Blizard, D. (1993). The Normandy Landings D-Day: The Invasion of Europe 6 June 1944. London: Reed International.Google Scholar
Bloom, S. A. (1980). Multivariate quantification of community recovery. In The Recovery Process in Damaged Ecosystems, ed. Cairns, J., pp. 141–51. Ann Arbor, MI: Ann Arbor Science Publishers.Google Scholar
Bodsworth, F. (1963). Last of the Curlews. Toronto, ON: McClelland and Stewart.Google Scholar
Boers, A. M., Veltman, R. L. D., and Zedler, J. B. (2007) Typha × glauca dominance and extended hydroperiod constrain restoration of wetland diversity. Ecological Engineering, 29, 232–44.CrossRefGoogle Scholar
Boesch, D. F., Josselyn, M. N., Mehta, A. J., Morris, J. T., Nuttle, W. K., Simenstad, C. A., and Swift, D. P. J. (1994). Scientific assessment of coastal wetland loss, restoration and management in Louisiana. Journal of Coastal Research, Special Issue No. 20.Google Scholar
Bogan, A. E. (1996). Margaritifera hembeli. In: IUCN (2007). 2007 IUCN Red List of Threatened Species. www.iucnredlist.org. (accessed June 30, 2008)
Bolen, E. G., Smith, L. M., and Schramm, H. L.. (1989). Playa lakes: prairie wetlands of the southern High Plains. BioScience, 39, 615–23.CrossRefGoogle Scholar
Bond, G. (1963). In Plant Physiology, eds. Salisbury, F. B. and Ross, C. W. (1985), 3rd edn, p. 254, Figure 13.3. Belmont, CA: Wadsworth.Google Scholar
Bondavalli, C. and Ulanowicz, R. E. (1999). Unexpected effects of predators upon their prey: the case of the American Alligator. Ecosystems, 2, 49–63.CrossRefGoogle Scholar
Bonetto, A. A. (1986). The Parana River system. In The Ecology of River Systems, eds. Davies, B. R. and Walker, K. F., pp. 541–55. Dordrecht, the Netherlands: Dr. W. Junk Publishers.CrossRefGoogle Scholar
Bonnicksen, T. M. (1988). Restoration ecology: philosophy, goals and ethics. The Environmental Professional, 10, 25–35.Google Scholar
Bormann, E. H. and Likens, G. E. (1981). Patterns and Process in a Forested Ecosystem. New York: Springer-Verlag.Google Scholar
Boston, H. L. (1986). A discussion of the adaptation for carbon acquisition in relation to the growth strategy of aquatic isoetids. Aquatic Botany, 26, 259–70.CrossRefGoogle Scholar
Boston, H. L. and Adams, M. S. (1986). The contribution of crassulacean acid metabolism to the annual productivity of two aquatic vascular plants. Oecologia, 68, 615–22.CrossRefGoogle ScholarPubMed
Botch, M. S. and Masing, V. V. (1983). Mire ecosystems in the USSR. In Ecosystems of the World, Vol. 4B, Mires: Swamp, Bog, Fen and Moor – Regional Studies, ed. Gore, A. J. P., pp. 95–152. Amsterdam, the Netherlands: Elsevier.Google Scholar
Botkin, D. B. (1990). Discordant Harmonies. A New Ecology for the Twenty-first Century. New York: Oxford University Press.Google Scholar
Boucher, D. H. (1985). The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Boutin, C. and Keddy, P. A. (1993). A functional classification of wetland plants. Journal of Vegetation Science, 4, 591–600.CrossRefGoogle Scholar
Bowden, W. B. (1987). The biogeochemistry of nitrogen in freshwater wetlands. Biogeochemistry, 4, 313–48.CrossRefGoogle Scholar
Bowers, M. D. (1991). Iridoid glycosides. In Herbivores: Their Interactions with Secondary Plant Metabolites, eds. Rosenthal, G. A. and Berenbaum, M. R., pp. 297–325. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Boyd, C. E. (1978). Chemical composition of wetland plants. In Freshwater Wetlands: Ecological Processes and Management Potential, eds. Good, R. E., Whigham, D. F., and Simpson, R. L., pp. 155–68. New York: Academic Press.Google Scholar
Boyd, R. and Penland, S. (1988). A geomorphologic model for Mississippi River Delta evolution, Transactions Gulf Coast Association of Geological Societies, 38, 443–52.Google Scholar
Bradley, C. E. and Smith, D. G. (1986). Plains cottonwood recruitment and survival on a prairie meandering river floodplain, Milk River, southern Alberta and northern Montana. Canadian Journal of Botany, 64, 1433–42.CrossRefGoogle Scholar
Brandle, R. A. (1991). Flooding resistance of rhizomatous amphibious plants. In Plant Life under Oxygen Deprivation, eds. Jackson, M. B., Davis, D. D., and Lambers, H., pp. 35–46. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Brasher, S. and Perkins, D. F. (1978). The grazing intensity and productivity of sheep in the grassland ecosystem. In Production Ecology of British Moors and Montane Grasslands, Ecological Studies Vol. 27, eds. Heal, O. W. and Perkins, D. F., pp. 354–74. Berlin, Germany: Springer-Verlag.CrossRefGoogle Scholar
Brewer, J. S. and Grace, J. B. (1990). Plant community structure in an oligohaline tidal marsh. Vegetatio, 90, 93–107.CrossRefGoogle Scholar
Bridgham, S. D., Pastor, J., Janssens, J. A., Chapin, C., and Malterer, T. J. (1996). Multiple limiting gradients in peatlands: a call for a new paradigm. Wetlands, 16, 45–65.CrossRefGoogle Scholar
Brinkman, R. and Diepen, C. A. (1990). Mineral soils. In Wetlands and Shallow Continental Water Bodies, Vol. 1, Natural and Human Relationships, ed. Patten, B. C., pp. 37–59. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Brinson, M. M. (1993a). Changes in the functioning of wetlands along environmental gradients. Wetlands, 13, 65–74.CrossRefGoogle Scholar
Brinson, M. M. (1993b). A Hydrogeomorphic Classification for Wetlands, Technical Report No. WRP-DE-4. Washington, DC: U.S. Army Corps of Engineers.Google Scholar
Brinson, M. M. (1995). Functional classifications of wetlands to facilitate watershed planning. In Wetlands and Watershed Management: Science Applications and Public Policy, eds. Kusler, J. A., Willard, D. E., and Hull, H. C., pp. 65–71. A collection of papers from a national symposium and several workshops at Tampa, FL, Apr 23–26. Berne, NY: Association of State Wetland Managers.Google Scholar
Brinson, M. M., Lugo, A. E. and Brown, S. (1981). Primary productivity, decomposition and consumer activity in freshwater wetlands. Annual Review of Ecology and Systematics, 12, 123–61.CrossRefGoogle Scholar
Brinson, M. M., Christian, R. R. and Blum, L. K. (1995). Multiple states in the sealevel induced transition from terrestrial forest to estuary. Estuaries, 18, 648–59.CrossRefGoogle Scholar
Bronmark, C. (1985). Interactions between macrophytes, epiphytes and herbivores: an experimental approach. Oikos, 45, 26–30.CrossRefGoogle Scholar
Bronmark, C. (1990). How do herbivorous freshwater snails affect macrophytes? – a comment. Ecology, 71, 1213–15.Google Scholar
Brosnan, D., Courtney, S., Sztukowski, L., Bedford, B., Burkett, V., Collopy, M., Derrickson, S., Elphick, C., Hunt, R., Potter, K., Sedinger, J. and Walters, J. (2007). Everglades Multi-Species Avian Ecology and Restoration Review: Final Report. Portland, OR: Sustainable Ecology Institute.Google Scholar
Brown, J. F. (1997). Effects of experimental burial on survival, growth, and resource allocation of three species of dune plants. Journal of Ecology, 85, 151–8.CrossRefGoogle Scholar
Brown, L. R. (2001). Paving the Planet: Cars and Crops Competing for Land. Washington, DC: Earth Policy Institute.Google Scholar
Brown, S., Brinson, M. M., and Lugo, A. E. (1979). Structure and function of riparian wetlands. In Strategies for Protection and Management of Floodplain Wetlands and Other Riparian Ecosystems, Gen. Tech. Rep. No. WO-12, tech. coord. Johnson, R. R. and McCormick, J. F., pp. 17–31. Washington, DC: U.S. Department of Agriculture, Forest Service.Google Scholar
Bruland, G. L. and Richardson, C. J. (2005). Hydrologic, edaphic, and vegetative responses to microtopographic reestablishment in a restored wetland. Restoration Ecology, 13, 515–23.CrossRefGoogle Scholar
Brunton, D. F. and Di Labio, B. M. (1989). Diversity and ecological characteristics of emergent beach flora along the Ottawa River in the Ottawa–Hull region, Quebec and Ontario. Naturaliste Canadien, 116, 179–91.Google Scholar
Brutsaert, W. (2005). Hydrology: An Introduction. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Bubier, J. L. (1995). The relationship of vegetation to methane emission and hydrochemical gradients in northern peatlands. Journal of Ecology, 83, 403–20.CrossRefGoogle Scholar
Bucher, E. H., Bonetto, A., Boyle, T. P., Canevari, P., Castro, G., Huszar, P., and Stone, T. (1993). Hidrovia: An Initial Environmental Examination of the Paraguay–Paraná Waterway. Manomet, MA and Buenos Aires, Argentina: Wetlands for the Americas.Google Scholar
Bump, S. R. (1986). Yellow-headed blackbird nest defense: aggressive responses to marsh wrens. The Condor, 88, 328–35.CrossRefGoogle Scholar
Burger, J., Shisler, J., and Lesser, F. H. (1982). Avian utilization on six salt marshes in New Jersey. Biological Conservation, 23, 187–212.CrossRefGoogle Scholar
Burnett, J. H. (1964). The study of Scottish vegetation. In The Vegetation of Scotland, ed. Burnett, J. H., pp. 1–11. Edinburgh, UK: Oliver and Boyd.Google Scholar
Bury, B. R. (1979). Population ecology of freshwater turtles. In Turtles: Perspectives and Research, eds. Harless, M. and Morlock, H., pp. 571–602. New York: John Wiley.Google Scholar
Cade, B. S. and Noon, B. R. (2003). A gentle introduction to quantile regression for ecologists. Frontiers in Ecology and the Environment, 1, 412–20.CrossRefGoogle Scholar
Cade, B. S., Terrell, J. W., and Schroeder, R. L. (1999). Estimating effects of limiting factors with regression quantiles. Ecology, 80, 311–23.CrossRefGoogle Scholar
Cairns, , J. (ed.) (1980). The Recovery Process in Damaged Ecosystems. Ann Arbor, MI: Ann Arbor Science Publishers.
Cairns, , J. (ed.) (1988). Rehabilitating Damaged Ecosystems, Vols. 1 and 2. Boca Raton, FL: CRC Press.
Cairns, J. (1989). Restoring damaged ecosystems: is predisturbance condition a viable option?The Environmental Professional, 11, 152–9.Google Scholar
Cairns, J., Niederlehner, B. R., and Orvos, D. R. (1992). Predicting Ecosystem Risk. Princeton, NJ: Princeton Scientific Publishing.Google Scholar
Callaway, R. M. and King, L. (1996). Temperature-driven variation in substrate oxygenation and the balance of competition and facilitation. Ecology, 77, 1189–95.CrossRefGoogle Scholar
Campbell, D. (2005). The Congo River basin. In The World's Largest Wetlands: Ecology and Conservation, eds. Fraser, L. H. and Keddy, P. A., pp. 149–65. Cambridge, UK: Cambridge University Press.Google Scholar
Campbell, D. R. and Rochefort, L. (2003). Germination and seedling growth of bog plants in relation to the recolonization of milled peatlands. Plant Ecology, 169, 71–84.CrossRefGoogle Scholar
,Canadian Hydrographic Service. (2009). Historical water level data. www.waterlevels.gc.ca/C&A/historical_e.html (accessed May 4, 2009)
Canny, M. J. (1998). Transporting water in plants. American Scientist, 86, 152–9.CrossRefGoogle Scholar
Carignan, R. and Kalff, J. (1980). Phosphorus sources for aquatic weeds: water or sediments?Science, 207, 987–9.CrossRefGoogle ScholarPubMed
Carpenter, S. R. and Kitchell, J. F. (1988). Consumer control of lake productivity. BioScience, 38, 764–9.CrossRefGoogle Scholar
Carpenter, S. R. and Lodge, D. M. (1986). Effects of submersed macrophytes on ecosystem processes. Aquatic Botany, 26, 341–70.CrossRefGoogle Scholar
Carpenter, S. R., Kitchell, J. F., Hodgson, J. R., Cochran, P. A., Elser, J. J., Elser, M. M., Lodge, D. M., Kretchmer, D., He, X., and Ende, C. N. (1987). Regulation of lake primary productivity by food web structure. Ecology, 68, 1863–76.CrossRefGoogle ScholarPubMed
Carpenter, S. R., Chisholm, S. W., Krebs, C. J., Schindler, D. W., and Wright, R. F. (1995). Ecosystem experiments. Science, 269, 324–7.CrossRefGoogle ScholarPubMed
Carvalho, A. R. (2007). An ecological economics approach to estimate the value of a fragmented wetland in Brazil (Mato Grosso do Sul state). Brazilian Journal of Biology, 67, 663–71.CrossRefGoogle Scholar
Carver, E. and Caudill, J. (2007). Banking on Nature: The Economic Benefits to Local Communites of National Wildlife Refuge Visitation. Washington, DC: U.S. Fish and Wildlife Service.Google Scholar
Castellanos, E. M., Figueroa, M. E., and Davy, A. J. (1994). Nucleation and facilitation in saltmarsh succession: interactions between Spartina maritima and Arthrocnemum perenne. Journal of Ecology, 82, 239–48.CrossRefGoogle Scholar
Catling, P. M., Spicer, K. W., and Lefkovitch, L. P. (1988). Effects of the introduced floating vascular aquatic, Hydrocharis morsus-ranae (Hydrocharitaceae), on some North American aquatic macrophytes. Naturaliste Canadien, 115, 131–7.Google Scholar
Cavalieri, A. J. and Huang, A. H. C. (1979). Evaluation of proline accumulation in the adaptation of diverse species of marsh halophytes to the saline environment. American Journal of Botany, 66, 307–12.CrossRefGoogle Scholar
Cazenave, A. and Nerem, R. (2004). Present-day sea level change: observations and causes. Reviews of Geophysics, 42, 139–50.CrossRefGoogle Scholar
Chaneton, E. J. and Facelli, J. M. (1991). Disturbance effects on plant community diversity: spatial scales and dominance hierarchies. Vegetatio, 93, 143–56.CrossRefGoogle Scholar
Chapin, F. S., III. (1980). The mineral nutrition of wild plants. Annual Review of Ecology and Systematics, 11, 233–60.CrossRefGoogle Scholar
Chapman, V. J. (1940). The functions of the pneumatophores of Avicennia nitida Jacq. Proceedings of the Linnean Society of London, 152, 228–33.CrossRefGoogle Scholar
Chapman, V. J. (1974). Salt Marshes and Salt Deserts of the World. Lehre, Germany: J. Cramer.CrossRefGoogle Scholar
Chapman, , V. J. (ed.) (1977). Wet Coastal Ecosystems. Amsterdam, the Netherlands: Elsevier.
Charlton, D. L. and Hilts, S. (1989). Quantitative evaluation of fen ecosystems on the Bruce Peninsula. In Ontario Wetlands: Inertia or Momentum, eds. Bardecki, M. J. and Patterson, N., pp. 339–54. Proceedings of Conference, Ryerson Polytechnical Institute, Toronto, Oct 21–22, 1988. Toronto, ON: Federation of Ontario Naturalists.Google Scholar
Cherry-Garrard, A. (1922). The Worst Journey in the World. London: Constable.Google Scholar
Chesson, P. L. and Warner, R. R. (1981). Environmental variability promotes coexistence in lottery competitive systems. The American Naturalist, 117, 923–43.CrossRefGoogle Scholar
Chimney, M. and Goforth, G. (2006). History and description of the Everglades Nutrient Removal Project. Ecological Engineering, 27, 268–78.CrossRefGoogle Scholar
,China Development Brief. (2004). Ploughshares into fishing nets. www.chinadevelopmentbrief.com/node/204 (accessed Dec 3, 2007)
Christensen, N. L. (1999). Vegetation of the Coastal Plain of the southeastern United States. In Vegetation of North America, 2nd edn, eds. Barbour, M. and Billings, W. D., pp. 397–448. Cambridge, UK: Cambridge University Press.Google Scholar
Christensen, N. L., Burchell, R. B., Liggett, A., and Simms, E. L. (1981). The structure and development of pocosin vegetation. In Pocosin Wetlands: An Integrated Analysis of Coastal Plain Freshwater Bogs in North Carolina, ed. Richardson, C. J., pp. 43–61. Stroudsburg, PA: Hutchinson Ross.Google Scholar
Christensen, N. L., Bartuska, A. M., Brown, J. H., Carpenter, S., D'Antonio, C., Francis, R., Franklin, J. F., MacMahon, J. A., Noss, R. F., Parsons, D. J., Peterson, C. H., Turner, M. G., and Woodmansee, R. G. (1996). The report of the Ecological Society of America Committee on the Scientific Basis for Ecosystem Management. Ecological Applications, 6, 665–91.CrossRefGoogle Scholar
Christie, W. J. (1974). Changes in the fish species composition of the Great Lakes. Journal of the Fisheries Research Board of Canada, 31, 827–54.CrossRefGoogle Scholar
Chung, C. (1982). Low marshes, China. In Creation and Restoration of Coastal Plant Communities, ed. Lewis, R. R., pp. 131–45. Boca Raton, FL: CRC Press.Google Scholar
Cicerone, R. J. and Ormland, R. S. (1988). Biogeochemical aspects of atmospheric methane. Global Biogeochemical Cycles, 2, 299–327.CrossRefGoogle Scholar
Clapham, W. B.. (1973). Natural Ecosystems. New York: Macmillan.Google Scholar
Clark, M. A., Siegrist, J., and Keddy, P. A. (2008). Patterns of frequency in species-rich vegetation in pine savannas: effects of soil moisture and scale. Ecoscience, 15, 529–35.CrossRefGoogle Scholar
Clarke, L. D. and Hannon, N. J. (1967). The mangrove swamp and salt marsh communities of the Sydney district. I. Vegetation, soils and climate. Journal of Ecology, 55, 753–71.CrossRefGoogle Scholar
Clarke, L. D. and Hannon, N. J. (1969). The mangrove swamp and salt marsh communities of the Sydney district. II. The holocoenotic complex with particular reference to physiography. Journal of Ecology, 57, 213–34.CrossRefGoogle Scholar
Clegg, J. (1986). Pond Life. London: Frederick Warne.Google Scholar
Clements, F. E. (1916). Plant Succession: An Analysis of the Development of Vegetation. Washington, DC: Carnegie Institution of Washington.CrossRefGoogle Scholar
Clements, F. E. (1935). Experimental ecology in the public service. Ecology, 16, 342–63.CrossRefGoogle Scholar
Clements, F. E. (1936). Nature and structure of climax. Journal of Ecology, 24, 254–82.CrossRefGoogle Scholar
Clements, F. E., Weaver, J. E., and Hanson, H. C. (1929). Plant Competition. Washington, DC: Carnegie Institution of Washington.Google Scholar
Clymo, R. S. and Duckett, J. G. (1986). Regeneration of Sphagnum. New Phytologist, 102, 589–614.CrossRefGoogle Scholar
Clymo, R. S. and Hayward, P. M. (1982). The ecology of Sphagnum. In Bryophyte Ecology, ed. Smith, A. J. E., pp. 229–89. London: Chapman and Hall.CrossRefGoogle Scholar
Cobbaert, D, Rochefort, L., and Price, J. S. (2004). Experimental restoration of a fen plant community after peat mining. Applied Vegetation Science, 7, 209–20.CrossRefGoogle Scholar
Coleman, J. M., Roberts, H. H., and Stone, G. W. (1998). Mississippi River Delta: an overview. Journal of Coastal Research, 14, 698–716.Google Scholar
Coles, B. and Coles, J. (1989). People of the Wetlands: Bogs, Bodies and Lake-Dwellers. London: Thames and Hudson.Google Scholar
Coley, P. D. (1983). Herbivory and defense characteristics of tree species in a lowland tropical forest. Ecological Monographs, 53, 209–33.CrossRefGoogle Scholar
Colinvaux, P. (1978). Why Big Fierce Animals Are Rare: An Ecologist's Perspective. Princeton, NJ: Princeton University Press.Google Scholar
Colwell, R. K. and Fuentes, E. R. (1975). Experimental studies of the niche. Annual Review of Ecology and Systematics, 6, 281–309.CrossRefGoogle Scholar
,Committee on Characterization of Wetlands. (1995). Wetlands: Characteristics and Boundaries. Washington, DC: National Academy of Sciences Press.
,Committee on Ecological Land Classification. (1988). Wetlands of Canada, Ecological Land Classification Series No. 24. Ottawa, ON: National Wetlands Working Group, Environment Canada.
Conant, R. and Collins, J. T. (1998). A Field Guide to Reptiles and Amphibians, Eastern/Central North America, 3rd edn. New York: Houghton Mifflin.Google Scholar
Connell, J. H. (1978). Diversity in tropical rain forests and coral reefs. Science, 199, 1302–10.CrossRefGoogle ScholarPubMed
Connell, J. H. (1980). Diversity and the coevolution of competitors, or the ghost of competition past. Oikos, 35, 131–8.CrossRefGoogle Scholar
Connell, J. H. (1987). Maintenance of species diversity in biotic communities. In Evolution and Coadaptation in Biotic Communities, eds. Kawano, S., Connell, J. H., and Hidaka, T., pp. 208–18. Tokyo: University of Tokyo Press.Google Scholar
Connell, J. H. and Orias, E. (1964). The ecological regulation of species diversity. The American Naturalist, 98, 399–414.CrossRefGoogle Scholar
Conner, W. H. and Buford, M. A. (1998). Southern deepwater swamps. In Southern Forested Wetlands: Ecology and Management, eds. Messina, M. G. and Conner, W. H., pp. 261–87. Boca Raton, FL: Lewis Publishers.Google Scholar
Conner, W. H., Day, J. W., Baumann, R. H., and Randall, J. M. (1989). Influence of hurricanes on coastal ecosystems along the northern Gulf of Mexico. Wetlands Ecology and Management, 1, 45–56.CrossRefGoogle Scholar
Connor, E. F. and McCoy, E. D. (1979). The statistics and biology of the species–area relationship. The American Naturalist, 113, 791–833.CrossRefGoogle Scholar
Connor, E. F. and Simberloff, D. (1979). The assembly of species communities: chance or competition?Ecology, 69, 1132–40.CrossRefGoogle Scholar
Cordone, A. J. and Kelley, D. W. (1961). The influences of inorganic sediment on the aquatic life of streams. California Fish and Game, 47, 189–228.Google Scholar
Cornwell, W. K., Bedford, B. L., and Chapin, C. T. (2001). Occurrence of arbuscular mycorrhizal fungi in a phosphorus-poor wetland and mycorrhizal response to phosphorus fertilization. American Journal of Botany, 88, 1824–9.CrossRefGoogle Scholar
Costanza, R., Cumberland, J., Daly, H., Goodland, R., and Norgaard, R. (1997). An Introduction to Ecological Economics. Boca Raton, FL: St. Lucie Press.CrossRefGoogle Scholar
Cowardin, L. M. and Golet, F. C. (1995). US Fish and Wildlife Service 1979 wetland classification: a review. Vegetatio, 118, 139–52.CrossRefGoogle Scholar
Cowardin, L. M., Carter, V., Golet, F. C., and LaRoe, E. T. (1979). Classification of Wetlands and Deepwater Habitats of the United States, FWS/OBS-79/31. Washington, DC: U.S. Department of the Interior Fish and Wildlife Service.Google Scholar
Cowling, R. M., Rundel, P. W., Lamont, B. B., Arroyo, M. K., and Arianoutsou, M. (1996a). Plant diversity in Mediterranean-climate regions. Trends in Ecology and Evolution, 11, 362–6.CrossRefGoogle ScholarPubMed
Cowling, R. M., MacDonald, I. A. W., and Simmons, M. T. (1996b). The Cape Peninsula, South Africa: physiographical, biological and historical background to an extraordinary hot-spot of biodiversity. Biodiversity and Conservation, 5, 527–50.CrossRefGoogle Scholar
Craft, C. B., Vymazal, J., and Richardson, C. J. (1995). Response of everglades plant communities to nitrogen and phosphorus additions. Wetlands, 15, 258–71.CrossRefGoogle Scholar
Craighead, F. C.. (1968). The role of the alligator in shaping plant communities and maintaining wildlife in the southern Everglades. The Florida Naturalist, 41, 2–7, 69–74.Google Scholar
Crawford, R. M. M. (1982). Physiological response to flooding. In Encyclopedia of Plant Physiology, new series Vol. 12B, Physiological Plant Ecology II, eds. Large, O. L., Nobel, P. S., Osmond, C. B., and Ziegler, H., pp. 453–77. Berlin, Germany: Springer-Verlag.Google Scholar
Crawford, R. M. M. and Braendle, R. (1996). Oxygen deprivation stress in a changing environment. Journal of Experimental Botany, 47, 145–59.CrossRefGoogle Scholar
Crawford, R. M. M. and McManmon, M (1968). Inductive responses of alcohol and malic acid dehydrogenases in relation to flooding tolerance in roots. Journal of Experimental Botany, 19, 435–41.CrossRefGoogle Scholar
Crawley, M. J. (1983). Herbivory: The Dynamics of Plant/Animal Interactions. Oxford, UK: Blackwell Scientific Publications.Google Scholar
Crocker, L. P. (1990). Army Officer's Guide, 45th edn. Harrisburg, PA: Stackpole Books.Google Scholar
Crook, D. A. and Robertson, A. I. (1999). Relationships between riverine fish and woody debris: implications for lowland rivers. Marine and Freshwater Research, 50, 941–53.CrossRefGoogle Scholar
Crosby, A.W. (1993). Economic Imperialism: The Biological Expansion of Europe 900–1900. Cambridge, UK: Cambridge University Press.Google Scholar
Crow, G. E. (1993). Species diversity in aquatic angiosperms: latitudinal patterns. Aquatic Botany, 44, 229–58.CrossRefGoogle Scholar
Crowder, A. A. and Bristow, J. M. (1988). Report: the future of waterfowl habitats in the Canadian lower Great Lakes wetlands. Journal of Great Lakes Research, 14, 115–27.CrossRefGoogle Scholar
Cummins, K. W. (1973). Trophic relationships of aquatic insects. Annual Review of Entomology, 18, 83–206.CrossRefGoogle Scholar
Cummins, K. W. and Klug, M. J. (1979). Feeding ecology of stream invertebrates. Annual Review of Ecology and Systematics, 10, 147–72.CrossRefGoogle Scholar
Currie, D. J. (1991). Energy and large-scale patterns of animal- and plant-species richness. The American Naturalist, 137, 27–49.CrossRefGoogle Scholar
Cyr, H. and Pace, M. L. (1993). Magnitude and patterns of herbivory in aquatic and terrestrial ecosystems. Nature, 361, 148–50.CrossRefGoogle Scholar
Czaya, E. (1983). Rivers of the World. Cambridge, UK: Cambridge University Press.Google Scholar
Dacey, J. W. H. (1980). Internal winds in water lillies: an adaptation for life in anaerobic sediments. Science, 210, 1017–19.CrossRefGoogle Scholar
Dacey, J. W. H. (1981). Pressurized ventilation in the yellow water lily. Ecology, 62, 1137–47.CrossRefGoogle Scholar
Dacey, J. W. H. (1988). In Plant Physiology, 3rd edn, eds. Salisbury, F. B. and Ross, C. W., pp. 68–70. Belmont, CA: Wadsworth.Google Scholar
Dahm, C. N., Cummins, K. W., Valett, H. M., and Coleman, R. L. (1995). An ecosystem view of the restoration of the Kissimmee River. Restoration Ecology, 3, 225–38.CrossRefGoogle Scholar
Daily, G. C. (1997). Nature's Services: Societal Dependence Upon Natural Ecosystems. Washington, DC: Island Press.Google Scholar
Damman, A. W. H. (1986). Hydrology, development, and biogeochemistry of ombrogenous bogs with special reference to nutrient relocation in a western Newfoundland bog. Canadian Journal of Botany, 64, 384–94.CrossRefGoogle Scholar
Damman, A. and Dowhan, J. (1981). Vegetation and habitat conditions in Western Head Bog, a southern Nova Scotian plateau bog. Canadian Journal of Botany, 59, 1343–59.CrossRefGoogle Scholar
Dansereau, P. (1959). Vascular aquatic plant communities of southern Quebec: a preliminary analysis. Transactions of the Northeast Wildlife Conference, 10, 27–54.Google Scholar
Dansereau, P. and Segadas-Vianna, F. (1952). Ecological study of the peat bogs of eastern North America. Canadian Journal of Botany, 30, 490–520.CrossRefGoogle Scholar
Darlington, P. J. (1957). Zoogeography: The Geographical Distribution of Animals. New York: John Wiley.Google Scholar
Davis, D. W. (2000). Historical perspective on crevasses, levees, and the Mississippi River. In Transforming New Orleans and Its Environs: Centuries of Change, ed. Colten, C. E., pp. 84–108. Pittsburgh, PA: University of Pittsburgh Press.Google Scholar
Davis, S.M. and Ogden, J.C. (eds.) (1994). Everglades: The Ecosystem and its Restoration. Delray Beach, FL: St. Lucie Press.
Day, J. W., Boesch, D. F., Clairain, E. J., Kemp, G. P., Laska, S. B., Mitsch, W. J., Orth, K., Mashriqui, H., Reed, D. J., Shabman, L., Simenstad, C. A., Streever, B. J., Twilley, R. R., Watson, C. C., Wells, J. T., and Whigham, D. F. (2007). Restoration of the Mississippi Delta: lessons from Hurricanes Katrina and Rita. Science, 315, 1679–84.CrossRefGoogle ScholarPubMed
Day, R. T., Keddy, P. A., McNeill, J., and Carleton, T. (1988). Fertility and disturbance gradients: a summary model for riverine marsh vegetation. Ecology, 69, 1044–54.CrossRefGoogle Scholar
Day, W. (1984). Genesis on Planet Earth, 2nd edn. New Haven, CT: Yale University Press.Google Scholar
Dayton, P. K. (1979). Ecology: a science and a religion. In Ecological Processes in Coastal and Marine Systems, ed. Livingston, R. J., pp. 3–18. New York: Plenum Press.CrossRefGoogle Scholar
DeBenedictis, P. A. (1974). Interspecific competition between tadpoles of Rana pipiens and Rana sylvatica: an experimental field study. Ecological Monographs, 44, 129–51.CrossRefGoogle Scholar
Groot, R. S. (1992). Functions of Nature. Groningen, the Netherlands: Wolters-Noordhoff.Google Scholar
Delany, S. N. and Scott, D. A. (2006). Waterbird Population Estimates, 4th edn. Wageningen, the Netherlands: Wetlands International.Google Scholar
Delcourt, H. R. and Delcourt, P. A. (1988). Quaternary landscape ecology: relevant scales in space and time. Landscape Ecology, 2, 23–44.CrossRefGoogle Scholar
Delcourt, H. R. and Delcourt, P. A. (1991). Quaternary Ecology: A Paleoecological Perspective. London: Chapman and Hall.CrossRefGoogle Scholar
del Moral, R., Titus, J. H., and Cook, A. M. (1995). Early primary succession on Mount St. Helens, Washington, USA. Journal of Vegetation Science, 6, 107–20.CrossRefGoogle Scholar
Luc, J. A. (1810). Geologic travels. In Gorham, E. (1953). Some early ideas concerning the nature, origin and development of peat lands. Journal of Ecology, 41, 257–74.
Denny, P. (1972). Sites of nutrient absorption in aquatic macrophytes. Journal of Ecology, 60, 819–29.CrossRefGoogle Scholar
Denny, P. (1985). The Ecology and Management of African Wetland Vegetation. Dordrecht, the Netherlands: Dr. W. Junk Publishers.CrossRefGoogle Scholar
Denny, P. (1993a).Wetlands of Africa: Introduction. In Wetlands of the World, Vol. 1, eds. Whigham, D. F., D. Dykyjova, and S. Hejny, pp. 1–31. Dordrecht, the Netherlands: Kluwer.Google Scholar
Denny, P. (1993b). Eastern Africa. In Wetlands of the World, Vol. 1, ed. Whigham, D. F., D. Dykyjova, and S. Hejny, pp. 32–46. Dordrecht, the Netherlands: Kluwer.Google Scholar
Denny, P. (1995). Benefits and priorities for wetland conservation: the case for national conservation strategies. In Wetlands. Archaeology and Nature Conservation, eds. Cox, M., Straker, V., and Taylor, D., pp. 249–74. London: HMSO.Google Scholar
Desmukh, I. (1986). Ecology and Tropical Biology. Palo Alto, CA: Blackwell Scientific Publications.Google Scholar
Desrochers, D. W., Keagy, J. C., and Cristol, D. A. (2008). Created versus natural wetlands: avian communities in Virgina salt marshes. Ecoscience, 15, 36–43.CrossRefGoogle Scholar
Diamond, J. M. (1975). Assembly of species communities. In Ecology and Evolution of Communities, eds. Cody, M. L. and Diamond, J. M., pp. 342–444. Cambridge, MA: Belknap Press of Harvard University Press.Google Scholar
Diamond, J. M. (1983). Laboratory, field and natural experiments. Nature, 304, 586–7.CrossRefGoogle Scholar
Diamond, J. (1994). Ecological collapses of past civilisations. Proceedings of the American Philosophical Society, 138, 363–70.Google Scholar
Diamond, J. (2005). Collapse: How Societies Choose to Fail or Succeed. New York: Penguin Books.Google Scholar
Dickinson, C. H. (1983). Micro-organisms in peatlands. In Ecosystems of the World Vol. 4A, Mires: Swamp, Bog, Fen and Moor–General Studies, ed. Gore, A. J. P., pp. 225–45. Amsterdam, the Netherlands: Elsevier.Google Scholar
Digby, P. G. N. and Kempton, R. A. (1987). Multivariate Analysis of Ecological Communities. London: Chapman and Hall.Google Scholar
Dinerstein, E. (1991). Seed dispersal by greater one-horned rhinoceros (Rhinoceros unicornis) and the flora of Rhinoceros latrines. Mammalia, 55, 355–62.CrossRefGoogle Scholar
Dinerstein, E. (1992). Effects of Rhinoceros unicornis on riverine forest structure in lowland Nepal. Ecology, 73, 701–4.CrossRefGoogle Scholar
Dittmar, L. A. and Neely, R. K. (1999). Wetland seed bank response to sedimentation varying in loading rate and texture. Wetlands, 19, 341–51.CrossRefGoogle Scholar
Douglas, B. C. (1997). Global sea rise: a redetermination. Surveys in Geophysics, 18, 279–92.CrossRefGoogle Scholar
Dowdeswell, J. A. (2006). The Greenland ice sheet and global sea-level rise. Science, 311, 963–4.CrossRefGoogle ScholarPubMed
Doyle, T. W., Garrett, F. G., and Books, M. A. (2003). Modeling mangrove forest migration along the southwest coast of Florida under climate change. In Integrated Assessment of the Climate Change Impacts on the Gulf Coast Region, eds. Ning, Z. H., Tumer, R. E., Doyle, T., and Abdollahi, K. K., pp. 211–21. Baton Rouge, LA: Gulf Coast Climate Change Assessment Council (GCRCC) and Louisiana State University (LSU) Graphic Services.Google Scholar
Dray, F. A., Bennett, B. C., and Center, T. D. (2006). Invasion history of Melaleuca quinquenervia (Cav.) S. T. Blake in Florida. Castanea, 71, 210–25.CrossRefGoogle Scholar
Dugan, , P. (ed.) (1993). Wetlands in Danger. New York: Oxford University Press.
Dugan, , P. (ed.) (2005). Guide to Wetlands. Buffalo, NY: Firefly Books.
Dumortier, M., Verlinden, A., Beeckman, H., and Mijnsbrugge, K. (1996). Effects of harvesting dates and frequencies on above- and below-ground dynamics in Belgian wet grasslands. Ecoscience, 3, 190–8.CrossRefGoogle Scholar
Duncan, R. P. (1993). Flood disturbance and the coexistence of species in a lowland podocarp forest, south Westland, New Zealand. Journal of Ecology, 81, 403–16.CrossRefGoogle Scholar
Durant, W. (1944). The Story of Civilization III: Caesar and Christ. New York: Simon and Schuster.Google Scholar
du Rietz, G. E. (1931). Life-Forms of Terrestrial Flowering Plants. Uppsala, Sweden: Almqvist & Wiksell.Google Scholar
Dynesius, M. and Nilsson, C. (1994). Fragmentation and flow regulation of river systems in the northern third of the world. Science, 266, 753–62.CrossRefGoogle Scholar
Edmonds, , J. (ed.) (1997). Oxford Atlas of Exploration. New York: Oxford University Press.
Ehrenfeld, J. G. (1983). The effects of changes in land-use on swamps of the New Jersey pine barrens. Biological Conservation, 25, 353–75.CrossRefGoogle Scholar
Ehrlich, A. and Ehrlich, P. (1981). Extinction: The Causes and Consequences of the Disappearance of Species. New York: Random House.Google Scholar
Elakovich, S. D. and Wootten, J. W. (1989). Allelopathic potential of sixteen aquatic and wetland plants. Journal of Aquatic Plant Management, 27, 78–84.Google Scholar
Ellenberg, H. (1985). Veranderungen der Flora Mitteleuropas unter dem Einflus von Dungung und Immissionen. Schweizerische Zeitschrift für Forstwesen, 136, 19–39.Google Scholar
Ellenberg, H. (1988). Floristic changes due to nitrogen deposition in central Europe. In Critical Loads for Sulfur and Nitrogen, eds. Nilsson, J. and Grennfelt, P., pp. 375–83. Report from a workshop held at Skokloster, Sweden, Mar 19–24, 1988. Copenhagen: Nordic Council of Ministers.Google Scholar
Ellenberg, H. (1989). Eutrophierung: das gravierendste Problem im Naturschutz?Norddeutsche Naturschutzakademie, 2, 9–12.Google Scholar
Ellery, W. N., Ellery, K., Rogers, K. H., McCarthy, T. S., and Walker, B. H. (1993). Vegetation, hydrology and sedimentation processes as determinants of channel form and dynamics in the northeastern Okavango Delta, Botswana. African Journal of Ecology, 31, 10–25.CrossRefGoogle Scholar
Ellison, A. M. and Farnsworth, E. J. (1996). Spatial and temporal variability in growth of Rhizophora mangle saplings on coral cays: links with variation in insolation, herbivory, and local sedimentation rate. Journal of Ecology, 84, 717–31.CrossRefGoogle Scholar
Elton, C. (1927). Animal Ecology. London: Sidgwick and Jackson.Google Scholar
Elveland, J. (1978). Management of Rich Fens in Northern Sweden: Studies of Various Factors Influencing the Vegetational Dynamics, Statens naturvardsverk PM 1007. Solna, Sweden: Forskningsnamnden.Google Scholar
Elveland, J. (1979). Irrigated and Naturally Flooded Hay-Meadows in North Sweden: A Nature Conservancy Problem, Statens naturvardsverk PM 1174. Solna, Sweden: Forskningssekretariatet.Google Scholar
Elveland, J. and Sjoberg, K. (1982). Some Effects of Scything and Other Management Procedures on the Plant and Animal Life of N. Swedish Wetlands Formerly Mown for Hay, Statens naturvardsverket PM 1516. Solna, Sweden: Forskningssekretariatet.Google Scholar
,Encyclopaedia Britannica. (1991). Vol. 16, p. 481. Chicago, IL: Encyclopaedia Britannica Inc.
,Environment Canada. (1976). Marine Environmental Data Service, Ocean and Aquatic Sciences: Monthly and Yearly Mean Water Levels, Vol. 1, Inland. Ottawa, ON: Department of Environment.
,Environment Canada. (2000). The Importance of Nature to Canadians: The Economic Significance of Nature-Related Activities. Ottawa, ON: Environment Canada.
Eriksson, O. (1993). The species-pool hypothesis and plant community diversity. Oikos, 68, 371–4.CrossRefGoogle Scholar
Essame, H. (1974). Patton: A Study in Command. New York: Charles Scribner's Sons.Google Scholar
Ewel, J. J. (1986). Invasibility: lessons from south Florida. In Ecology of Biological Invasions of North America and Hawaii, eds. Mooney, H. A. and Drake, J. A., pp. 214–30. New York: Springer-Verlag.CrossRefGoogle Scholar
Facelli, J. M., Leon, R. J. C., and Deregibus, V. A. (1989). Community structure in grazed and ungrazed grassland sites in the flooding Pampa, Argentina. American Midland Naturalist, 121, 125–33.CrossRefGoogle Scholar
Faith, D. P., Minchin, P. R. and Belbin, L. (1987). Compositional dissimilarity as a robust measure of ecological distance. Vegetatio, 69, 57–68.CrossRefGoogle Scholar
Farney, R. A. and Bookhout, T. A. (1982). Vegetation changes in a Lake Erie marsh (Winous Point, Ottawa County, Ohio) during high water years. Ohio Journal of Science, 82, 103–7.Google Scholar
Faulkner, S. P. and Richardson, C. J. (1989). Physical and chemical characteristics of freshwater wetland soils. In Constructed Wetlands for Wastewater Treatment, ed. Hammer, D. A., pp. 41–72. Chelsea, MI: Lewis Publishers.Google Scholar
Fernandez-Armesto, F. (1989). The Spanish Armada: The Experience of War in 1588. Oxford, UK: Oxford University Press.Google Scholar
Field, R., Stuzeski, E. J., Masters, H. E., and Tafuri, A. N. (1974). Water pollution and associated effects from street salting. Journal of Environmental Engineering Division, 100, 459–77.Google Scholar
Findlay, S. C. and Houlahan, J. (1997). Anthropogenic correlates of biodiversity in southeastern Ontario wetlands. Conservation Biology, 11, 1000–9.CrossRefGoogle Scholar
Finney, B. P. and Johnson, T. C. (1991). Sedimentation in Lake Malawi (East Africa) during the past 10,000 years: a continuous paleoclimatic record from the southern tropics. Palaeogeography, Palaeoclimatology, Palaeoecology, 85, 351–66.CrossRefGoogle Scholar
Fitter, A. and Hay, R. (2002). Environmental Physiology of Plants, 3rd edn. San Diego, CA: Academic Press.Google Scholar
Flint, R. F. (1971). Glacial and Quaternary Geology. New York: John Wiley.Google Scholar
Flores, D. L. (ed.) (1984). Jefferson and Southwestern Exploration: The Freeman and Custis Accounts of the Red River Expedition of 1806. Norman, OK: University of Oklahoma Press.Google Scholar
,Food and Agriculture Organization of the United Nations (FAO). (2009). Commodities by Country. http://faostat.fao.org/site/339/default.aspx (accessed Dec 4, 2009)
Forman, A. T. and Alexander, L. E. (1998). Roads and their ecological effects. Annual Review of Ecology and Systematics, 29, 207–31.CrossRefGoogle Scholar
Forman, , R. T. T. (ed.) (1998). Pine Barrens: Ecosystem and Landscape. Rutgers, NJ: Rutgers University Press.
Forman, R. T. T., Sperling, D., Bissonette, J., Clevenger, A. P., Cutshall, C. D., Dale, V. H., Fahrig, L., France, R., Goldman, C. R., Heanue, K., Jones, J. A., Swanson, F. J., Turrentine, T., and Winter, T. C. (2002). Road Ecology: Science and Solutions. Washington, DC: Island Press.Google Scholar
Forster, P., Ramaswamy, V., Artaxo, P., Berntsen, T., Betts, R., Fahey, D. W., Haywood, J., Lean, J., Lowe, D. C., Myhre, G., Nganga, J., Prinn, R., Raga, G., Schulz, M., and Dorland, R. (2007). Changes in atmospheric constituents and in radiative forcing. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, eds. Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K. B., Tignor, M. M. B. and Miller, H. L., pp. 129–234. Cambridge, UK: Cambridge University Press.Google Scholar
Foster, D. R. and Glaser, P. H. (1986). The raised bogs of south-eastern Labrador, Canada: classification, distribution, vegetation and recent dynamics. Journal of Ecology, 74, 47–71.CrossRefGoogle Scholar
Foster, D. R. and Wright, H. E. (1990). Role of ecosystem development and climate change in bog formation in central Sweden. Ecology, 71, 450–63.CrossRefGoogle Scholar
Foster, D. R., King, G. A., Glaser, P. H., and Wright, H. E. (1983). Origin of string patterns in boreal peatlands. Nature, 306, 256–7.CrossRefGoogle Scholar
Fox, A. D. and Kahlert, J. (1999). Adjustments to nitrogen metabolism during wing moult in Greylag Geese, Anser anser. Functional Ecology, 13, 661–9.CrossRefGoogle Scholar
Fragoso, J. M. V. (1998). Home range and movement patterns of white-lipped Peccary (Tayassu pecari) herds in the northern Brazilian Amazon. Biotropica, 30, 458–69.CrossRefGoogle Scholar
Francis, T. B. and Schindler, D. E. (2006). Degradation of littoral habitats by residential development: woody debris in lakes of the Pacific Northwest and Midwest, United States. AMBIO: A Journal of the Human Environment, 35, 274–80.CrossRefGoogle ScholarPubMed
Fraser, A. (1973). Cromwell: The Lord Protector. New York: Konecky and Konecky.Google Scholar
Fraser, , L. H.andKeddy, P. A. (eds.) (2005). The World's Largest Wetlands: Ecology and Conservation. Cambridge, UK: Cambridge University Press.CrossRef
Freedman, B. (1995). Environmental Ecology, 2nd edn. San Diego, CA: Academic Press.Google Scholar
Fremlin, G. (ed. in chief) (1974). The National Atlas of Canada, 4th edn, revd. Toronto, ON: Macmillan.Google Scholar
Frenzel, B. (1983). Mires: repositories of climatic information or self-perpetuating ecosystems? In Ecosystems of the World Vol. 4A, Mires: Swamp, Bog, Fen and Moor – General Studies, ed. Gore, A. J. P., pp. 35–65. Amsterdam, the Netherlands: Elsevier.Google Scholar
Fretwell, S. D. (1977). The regulation of plant communities by food chains exploiting them. Perspectives in Biology and Medicine, 20, 169–85.CrossRefGoogle Scholar
Frey, R. W. and Basan, P. B. (1978). Coastal salt marshes. In Coastal Sedimentary Environments, ed. Davis, R. A., pp. 101–69. New York: Springer-Verlag.CrossRefGoogle Scholar
Fritzell, E. K. (1989). Mammals in prairie wetlands. In Northern Prairie Wetlands, ed. Valk, A., pp. 268–301. Ames, IA: Iowa State University Press.Google Scholar
Galatowitsch, S. M. and Valk, A. G. (1994). Restoring Prairie Wetlands: An Ecological Approach. Ames, IA: Iowa State University Press.Google Scholar
Galatowitsch, S. M. and Valk, A. G. (1996). The vegetation of restored and natural prairie wetlands. Ecological Applications, 6, 102–12.CrossRefGoogle Scholar
Galinato, M. and Valk, A. (1986). Seed germination of annuals and emergents recruited during drawdowns in the Delta Marsh, Manitoba, Canada. Aquatic Botany, 26, 89–102.CrossRefGoogle Scholar
Garcia, L. V., Maranon, T., Moreno, A., and Clemente, L. (1993). Above-ground biomass and species richness in a Mediterranean salt marsh. Journal of Vegetation Science, 4, 417–24.CrossRefGoogle Scholar
Gastescu, P. (1993). The Danube Delta: geographical characteristics and ecological recovery. Earth and Environmental Science, 29, 57–67.Google Scholar
Gaston, K. J. (2000). Global patterns in biodiversity. Nature, 405, 220–7.CrossRefGoogle ScholarPubMed
Gaston, K. J., Williams, P. H., Eggleton, P., and Humphries, C. J. (1995). Large scale patterns of biodiversity: spatial variation in family richness. Proceedings of the Royal Society of London Series B, 260, 149–54.CrossRefGoogle Scholar
Gaudet, C. L. and Keddy, P. A. (1988). A comparative approach to predicting competitive ability from plant traits. Nature, 334, 242–3.CrossRefGoogle Scholar
Gaudet, C. L. and Keddy, P. A. (1995). Competitive performance and species distribution in shoreline plant communities: a comparative approach. Ecology, 76, 280–91.CrossRefGoogle Scholar
Geho, E. M., Campbell, D., and Keddy, P. A. (2007). Quantifying ecological filters: the relative impact of herbivory, neighbours, and sediment on an oligohaline marsh. Oikos, 116, 1006–16.CrossRefGoogle Scholar
Geis, J. W. (1985). Environmental influences on the distribution and composition of wetlands in the Great Lakes basin. In Coastal Wetlands, eds. Prince, H. H. and D'Itri, F. M., pp. 15–31. Chelsea, MI: Lewis Publishers.Google Scholar
Gentry, A. H. (1988). Changes in plant community diversity and floristic composition on environmental and geographical gradients. Annals of the Missouri Botanical Garden, 75, 1–34.CrossRefGoogle Scholar
,German Advisory Council on Global Change. (2006). The Future Oceans: Warming Up, Rising High, Turning Sour, Special Report. Berlin, Germany: German Advisory Council on Global Change.
Gibson, D. J., Zampella, R. A., and Windisch, A. G. (1999). New Jersey Pine Plains: the “true barrens” of the New Jersey Pine Barrens. In Savannas, Barrens, and Rock Outcrop Communities of North America, eds. Anderson, R. C.. Fralish, J. S., and Bastin, J. M., pp. 52–66. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Gignac, L. D. and Vitt, D. H. (1990). Habitat limitations of Sphagnum along climatic, chemical, and physical gradients in mires of western Canada. The Bryologist, 93, 7–22.CrossRefGoogle Scholar
Gilbert, J. J. (1988). Suppression of rotifer populations by Daphnia: a review of the evidence, the mechanisms, and the effects on zooplankton community structure. Limnology and Oceanography, 33, 1286–303.CrossRefGoogle Scholar
Gilbert, J. J. (1990). Differential effects of Anabaena affinis on cladoceran and rotifers: mechanisms and implications. Ecology, 71, 1727–40.CrossRefGoogle Scholar
Gilbert, R. and Glew, J. R. (1986). A wind-driven ice-push event in eastern Lake Ontario. Journal of Great Lakes Research, 12, 326–31.CrossRefGoogle Scholar
Gill, D. (1973). Modification of northern alluvial habitats by river development. The Canadian Geographer, 17, 138–53.CrossRefGoogle Scholar
Giller, K. E. and Wheeler, B. D. (1986). Past peat cutting and present vegetation patterns in an undrained fen in the Norfolk Broadland. Journal of Ecology, 74, 219–47.CrossRefGoogle Scholar
Givnish, T. J. (1982). On the adaptive significance of leaf height in forest herbs. The American Naturalist, 120, 353–81.CrossRefGoogle Scholar
Givnish, T. J. (1988). Ecology and evolution of carnivorous plants. In Plant–Animal Interactions, ed. Abrahamson, W. B., pp. 243–90. New York: McGraw-Hill.Google Scholar
Gladwell, M. (2002). The Tipping Point: How Little Things Can Make a Big Difference. New York: Little, Brown.Google Scholar
Glaser, P. H. (1992). Raised bogs in eastern North America: regional controls for species richness and floristic assemblages. Journal of Ecology, 80, 535–54.CrossRefGoogle Scholar
Glaser, P. H., Janssens, J. A., and Siegel, D. I. (1990). The response of vegetation to chemical and hydrological gradients in the Lost River peatland, northern Minnesota. Journal of Ecology, 78, 1021–48.CrossRefGoogle Scholar
Gleason, H. A. (1926). The individualistic concept of the plant association. Bulletin of the Torrey Botanical Club, 53, 7–26.CrossRefGoogle Scholar
Gleason, H. A. (1939). The individualistic concept of the plant association. American Midland Naturalist, 21, 92–110.CrossRefGoogle Scholar
Glob, P. V. (1969). The Bog People. Iron-Age Man Preserved, translated from the Danish by Bruce-Mitford, R.. Ithaca, NY: Cornell University Press.Google Scholar
Glooschenko, W. A. (1980). Coastal ecosystems of the James/Hudson Bay area of Ontario, Canada. Zeitschrift für Geomorphologie, NF, 34, 214–24.Google Scholar
Godfrey, W. E. (1966). The Birds of Canada. Ottawa, ON: Information Canada.Google Scholar
Godwin, Sir H. (1981). The Archives of the Peat Bogs. Cambridge, UK: Cambridge University Press.Google Scholar
Godwin, K. S., Shallenberger, J., Leopold, D. J., and Bedford, B. L. (2002). Linking landscape properties to local hydrogeologic gradients and plant species occurrence in New York fens: a hydrogeologic setting (HGS) framework. Wetlands, 22, 722–37.CrossRefGoogle Scholar
Goethe, J. W. (1831). Goethe's Faust, Part 2, translated by Taylor, B., revised and edited by Atkins, S., 1962. New York: Collier Books.Google Scholar
Goin, C. J. and Goin, O. B. (1971). Introduction to Herpetology, 2nd edn. San Francisco, CA: W. H. Freeman.Google Scholar
Goldsmith, F. B. (1973). The vegetation of exposed sea cliffs at South Stack, Anglesey. II. Experimental studies. Journal of Ecology, 61, 819–29,CrossRefGoogle Scholar
Goldsmith, , F. B. (ed.) (1991). Monitoring for Conservation and Ecology. London: Chapman and Hall.CrossRef
Goldsmith, F. B. and Harrison, C. M. (1976). Description and analysis of vegetation. In Methods in Plant Ecology, ed. Chapman, S. B., pp. 85–155. Oxford, UK: Blackwell Scientific Publications.Google Scholar
Golet, F. C. and Parkhurst, J. A. (1981). Freshwater wetland dynamics in South Kingston, Rhode Island, 1939–1972. Environmental Management, 5, 245–51.CrossRefGoogle Scholar
Good, , R. E., Whigham, D. F., and Simpson, R. L (eds.) (1978). Freshwater Wetlands: Ecological Processes and Management Potential. New York: Academic Press.
Gopal, B. (1990). Nutrient dynamics of aquatic plant communities. In Ecology and Management of Aquatic Vegetation in the Indian Subcontinent, ed. Gopal, B., pp. 177–97. Dordrecht, the Netherlands: Kluwer.CrossRefGoogle Scholar
Gopal, B. and Goel, U. (1993). Competition and allelopathy in aquatic plant communities. Botanical Review, 59, 155–210.CrossRefGoogle Scholar
Gopal, B., Kvet, J., Loffler, H., Masing, V. and Patten, B. (1990). Definition and classification. In Wetlands and Shallow Continental Water Bodies, Vol. 1, Natural and Human Relationships, ed. Patten, B. C., pp. 9–15. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Gore, , A. J. P. (ed.) (1983). Ecosystems of the World, Vol. 4A, Mires: Swamp, Bog, Fen and Moor – General Studies. Amsterdam, the Netherlands: Elsevier.
Gore, A. J. P. (1983). Introduction. In Ecosystems of the World, Vol. 4A, Mires: Swamp, Bog, Fen and Moor – General Studies, ed. Gore, A. J. P.. Amsterdam, the Netherlands: Elsevier.Google Scholar
Gorham, E. (1953). Some early ideas concerning the nature, origin and development of peat lands. Journal of Ecology, 41, 257–74.CrossRefGoogle Scholar
Gorham, E. (1957). The development of peatlands. Quarterly Review of Biology, 32, 145–66.CrossRefGoogle Scholar
Gorham, E. (1961). Water, ash, nitrogen and acidity of some bog peats and other organic soils. Journal of Ecology, 49, 103–6.CrossRefGoogle Scholar
Gorham, E. (1990). Biotic impoverishment in northern peatlands. In The Earth in Transition, ed. Woodwell, G. M., pp. 65–98. Cambridge, UK: Cambridge University Press.Google Scholar
Gorham, E. (1991). Northern peatlands role in the carbon cycle and probable responses to climatic warming. Ecological Applications, 1, 182–95.CrossRefGoogle ScholarPubMed
Gosselink, J. G. and Turner, R. E. (1978). The role of hydrology in freshwater wetland ecosystems. In Freshwater Wetlands: Ecological Processes and Management Potential, eds. Good, R. E., Whigham, D. F., and Simpson, R. L., pp. 63–79. New York: Academic Press.Google Scholar
Gottlieb, A. D., Richards, J. H., and Gaiser, E. E. (2006). Comparative study of periphyton community structure in long and short hydroperiod Everglades marshes. Hydrobiologia, 569, 195–207.CrossRefGoogle Scholar
Gough, J. (1793). Reasons for supposing that lakes have been more numerous than they are at present; with an attempt to assign the causes whereby they have been defaced. Memoirs of the Literary and Philosophical Society of Manchester, 4, 1–19. In Walker, D. (1970). Direction and Rate in Some British Post-Glacial Hydroseres. In Studies in the Vegetational History of the British Isles, eds. Walker, D. and West, R. G., pp. 117–39. Cambridge, UK: Cambridge University Press.Google Scholar
Gough, L. G., Grace, J. B., and Taylor, K. L. (1994). The relationship between species richness and community biomass: the importance of environmental variables. Oikos, 70, 271–9.CrossRefGoogle Scholar
Goulding, M. (1980). The Fishes and the Forest: Explorations in Amazonian Natural History. Berkeley, CA: University of California Press.Google Scholar
Grace, J. B. (1990). On the relationship between plant traits and competitive ability. In Perspectives on Plant Competition, eds. Grace, J. B. and Tilman, D., pp. 51–65. San Diego, CA: Academic Press.Google Scholar
Grace, J. B. (1999). The factors controlling species density in herbaceous plant communities: an assessment. Perspectives in Plant Ecology, Evolution and Systematics, 2, 1–28.CrossRefGoogle Scholar
Grace, J. B. and Ford, M. A. (1996). The potential impact of herbivores on the susceptibility of the marsh plant Sagittaria lancifolia to saltwater intrusion in coastal wetlands. Estuaries, 19, 13–20.CrossRefGoogle Scholar
Grace, J. B. and Wetzel, R. G. (1981). Habitat partitioning and competitive displacement in cattails (Typha): experimental field studies. The American Naturalist, 118, 463–74.CrossRefGoogle Scholar
Graf, D. L. and Cummings, K. S. (2007). Review of the systematics and global diversity of freshwater mussel species (Bivalvia: Unionoida). Journal of Molluscan Studies, 73, 291–314.CrossRefGoogle Scholar
Graham, J. B. (1997). Air Breathing Fishes. San Diego, CA: Academic Press.Google Scholar
Greening, H. (1995). Resource-based watershed management in Tampa Bay. In Wetlands and Watershed Management: Science Applications and Public Policy, eds. Kusler, J. A., Willard, D. E., and Hull, H. C., pp. 172–81. A collection of papers from a national symposium and several workshops at Tampa, FL, Apr 23–26. Berne, NY: Association of State Wetland Managers.Google Scholar
Griffiths, R. A., Denton, J., and Wong, A. L. (1993). The effect of food level on competition in tadpoles: interference mediated by protothecan algae?Journal of Animal Ecology, 62, 274–9.CrossRefGoogle Scholar
Grime, J. P. (1973). Competitive exclusion in herbaceous vegetation. Nature, 242, 344–7.CrossRefGoogle Scholar
Grime, J. P. (1974). Vegetation classification by reference to strategies. Nature, 250, 26–31.CrossRefGoogle Scholar
Grime, J. P. (1977). Evidence for the existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. The American Naturalist, 111, 1169–94.CrossRefGoogle Scholar
Grime, J. P. (1979). Plant Strategies and Vegetation Processes. Chichester, UK: John Wiley.Google Scholar
Grime, J. P. and Hunt, R. (1975). Relative growth-rate: its range and adaptive significance in a local flora. Journal of Ecology, 63, 393–422.CrossRefGoogle Scholar
Grime, J. P., Mason, G., Curtis, A. V., Rodman, J., Band, S. R., Mowforth, M. A. G., Neal, A. M., and Shaw, S. (1981). A comparative study of germination characteristics in a local flora. Journal of Ecology, 69, 1017–59.CrossRefGoogle Scholar
Grimes, W. (2006). Visionaries and rascals in Florida's wetlands: review of The Swamp: The Everglades, Florida and the Politics of Paradise. The Washington Post, Mar 8, 2006.Google Scholar
Grishin, S. Y., del Moral, R., Krestov, P. V., and Verkholat, V. P. (1996). Succession following the catastrophic eruption of Ksudach volcano (Kamchatka, 1907). Vegetatio, 127, 129–53.CrossRefGoogle Scholar
Groombridge, , B. (ed.) (1992). Global Biodiversity: Status of the Earth's Living Resources, a report compiled by the World Conservation Monitoring Centre. London: Chapman and Hall.CrossRef
Grootjans, A. P., Diggelen, R., Everts, H. F., Schipper, P. C., Streefkerk, J., Vries, N. P., and Wierda, A. (1993). Linking ecological patterns to hydrological conditions on various spatial scales: a case study of small stream valleys. In Landscape Ecology of a Stressed Environment, eds. Vos, C. C. and Opdam, P., pp. 60–99. London: Chapman and Hall.CrossRefGoogle Scholar
Grosse, W., Buchel, H. B., and Tiebel, H. (1991). Pressurized ventilation in wetland plants. Aquatic Botany, 39, 89–98.CrossRefGoogle Scholar
Grover, A. M. and Baldassarre, G. A. (1995). Bird species richness within beaver ponds in south-central New York. Wetlands, 15, 108–18.CrossRefGoogle Scholar
Grubb, P. J. (1977). The maintenance of species-richness in plant communities: the importance of the regeneration niche. Biological Review, 52, 107–45.CrossRefGoogle Scholar
Grubb, P. J. (1985). Plant populations and vegetation in relation to habitat disturbance and competition: problems of generalizations. In The Population Structure of Vegetation, ed. White, J., pp. 595–621. Dordrecht, the Netherlands: Dr. W. Junk Publishers.CrossRefGoogle Scholar
Grubb, P. J. (1986). Problems posed by sparse and patchily distributed species in species-rich plant communities. In Community Ecology, eds. Diamond, J. M. and Case, T. J., pp. 207–25. New York: Harper and Row.Google Scholar
Grubb, P. J. (1987). Global trends in species-richness in terrestrial vegetation: a view from the northern hemisphere. In Organization of Communities Past and Present, eds. Gee, J. H. R. and Giller, P. S., pp. 99–118. 27th Symposium of the British Ecological Society, Aberystwyth. Oxford, UK: Blackwell Scientific Publications.Google Scholar
Grumbine, R. E. (1994). What is ecosystem management?Conservation Biology, 8, 27–38.CrossRefGoogle Scholar
Grumbine, R. E. (1997). Reflections on ‘What is ecosystem management?’Conservation Biology, 11, 41–7.CrossRefGoogle Scholar
Grunwald, M. (2006). The Swamp: The Everglades, Florida and the Politics of Paradise. New York: Simon and Schuster.Google Scholar
Gurevitch, J., Morrow, L., Wallace, A., and Walsh, A. (1992). A meta-analysis of competition in field experiments. The American Naturalist, 140, 539–72.CrossRefGoogle Scholar
Guy, H. P. (1973). Sediment problems in urban areas. In Focus on Environmental Geology, ed. Tank, R. W., pp. 186–92. New York: Oxford University Press.Google Scholar
Guyer, C. and Bailey, M. A. (1993). Amphibians and reptiles of longleaf pine communities. In The Longleaf Pine Ecosystem: Ecology, Restoration and Management, ed. Hermann, S. M., pp. 139–58. Proceedings of the Tall Timbers Fire Ecology Conference No. 18. Tallahassee, FL: Tall Timbers Research Station.Google Scholar
Hacker, S. D. and Bertness, M. D. (1999). Experimental evidence for factors maintaining plant species diversity in a New England salt marsh. Ecology, 80, 2064–73.CrossRefGoogle Scholar
Hacker, S. D. and Gaines, S. D. (1997). Some implications of direct positive interactions for community species diversity. Ecology, 78, 1990–2003.CrossRefGoogle Scholar
Haeuber, R. and Franklin, J. (eds.) (1996). Perspectives on ecosystem management. Ecological Applications, 6, 692–747.CrossRef
Hairston, N. G., Smith, F. E., and Slobodkin, L. B. (1960). Community structure, population control, and competition. The American Naturalist, 94, 421–5.CrossRefGoogle Scholar
Haith, D. A. and Shoemaker, L. L. (1987). Generalized watershed loading functions for stream-flow nutrients. Water Resources Bulletin, 23, 471–8.CrossRefGoogle Scholar
Hamilton, S. K., Sipel, S. J., and Melack, J. M. (1996). Inundation patterns in the Pantanal wetland of South America determined from passive microwave remote sensing. Archiv für Hydrobiologie, 137, 1–23.Google Scholar
Hammer, D. A. (1969). Parameters of a marsh snapping turtle population Lacreek refuge, South Dakota. Journal of Wildlife Management, 33, 995–1005.CrossRefGoogle Scholar
Hammer, , D. A. (ed.) (1989). Constructed Wetlands for Wastewater Treatment: Municipal, Industrial and Agricultural. Chelsea, MI: Lewis Publishers.
Hanski, I. (1994). Patch-occupancy dynamics in fragmented landscapes. Trends in Ecology and Evolution, 9, 131–5.CrossRefGoogle ScholarPubMed
Hanski, I. and Gilpin, M. (1991). Metapopulation dynamics: a brief history and conceptual domain. Biological Journal of the Linnean Society, 42, 3–16.CrossRefGoogle Scholar
Hardin, G. (1968). The tragedy of the commons. Science, 162, 1243–8.Google ScholarPubMed
Hardin, G. and Baden, J. (1977). Managing the Commons. San Francisco, CA: W. H. Freeman.Google Scholar
Harington, C. R. (1996). Giant beaver. (Reproduced courtesy of the Canadian Museum of Nature, Ottawa). www.beringia.com/02/02maina6.html (accessed July 28, 2008)
Harmon, M. E., Franklin, J. F., Swanson, F. J., Sollins, P., Gregory, S. V., Lattin, J. D., Anderson, N. H., and Cline, S. P. (1986). Ecology of coarse woody debris in temperate ecosystems. Advances in Ecological Research, 15, 133–302.CrossRefGoogle Scholar
Harper, J. L. (1977). Population Biology of Plants. London: Academic Press.Google Scholar
Harper, J. L., Williams, J. T., and Sagar, G. R. (1965). The behavior of seeds in soil. I. The heterogeneity of soil surfaces and its role in determining the establishment of plants from seed. Journal of Ecology, 53, 273–86.CrossRefGoogle Scholar
Harris, R. R., Fox, C. A., and Risser, R. (1987). Impact of hydroelectric development on riparian vegetation in the Sierra Nevada region, California, USA. EnvironmentalManagement, 11, 519–27.Google Scholar
Harris, S. W. and Marshall, W. H. (1963). Ecology of water-level manipulations on a northern marsh. Ecology, 44, 331–43.CrossRefGoogle Scholar
Hart, D. D. (1983). The importance of competitive interactions within stream populations and communities. In Stream Ecology: Application and Testing of General Ecological Theory, eds. Barnes, J. R. and Minshall, G. W., pp. 99–136. New York: Plenum Press.CrossRefGoogle Scholar
Hartman, J. M. (1988). Recolonization of small disturbance patches in a New England salt marsh. American Journal of Botany, 75, 1625–31.CrossRefGoogle Scholar
Harvey, P. H., Colwell, R. K., Silvertown, J. W., and May, R. M. (1983). Null models in ecology. Annual Review of Ecology and Systematics, 14, 189–211.CrossRefGoogle Scholar
Haukos, D. A. and Smith, L. M. (1993). Seed-bank composition and predictive ability of field vegetation in playa lakes. Wetlands, 13, 32–40.CrossRefGoogle Scholar
Haukos, D. A. and Smith, L. M. (1994). Composition of seed banks along an elevational gradient in playa wetlands. Wetlands, 14, 301–7.CrossRefGoogle Scholar
Hayati, A. A. and Proctor, M. C. F. (1991). Limiting nutrients in acid-mire vegetation: peat and plant analyses and experiments on plant responses to added nutrients. Journal of Ecology, 79, 75–95.CrossRefGoogle Scholar
Heal, G. (2000). Valuing ecosystem services. Ecosystems, 3, 24–30.CrossRefGoogle Scholar
Heal, O. W., Latter, P. M., and Howson, G. (1978). A study of the rates of decomposition of organic matter. In Production Ecology of British Moors and Montane Grasslands, eds. Heal, O. W. and Perkins, D. F., pp. 136–59. Berlin, Germany: Springer-Verlag.CrossRefGoogle Scholar
Hellquist, C. B. and Crow, G. E. (1984). Aquatic Vascular Plants of New England, Part 7, Cabombaceae, Nymphaeaceae, Nelumbonaceae, and Ceratophyllaceae, Station Bulletin No. 527. Durham, NH: University of New Hampshire.Google Scholar
,Helsinki Commission. (2003). The Baltic Marine Environment 1999–2002, Baltic Sea Environment Proceedings No. 87. Helsinki: Helsinki Commission.
Hemphill, N. and Cooper, S. D. (1983). The effect of physical disturbance on the relative abundances of two filter-feeding insects in a small stream. Oecologia, 58, 378–82.CrossRefGoogle Scholar
Henry, H. A. L. and Jeffries, R. L. (2009). Opportunist herbivores, migratory connectivity and catastrophic shifts in arctic coastal systems. In Human Impacts on Salt Marshes: A Global Perspective, eds. Silliman, B. R., Grosholz, E. D., and Bertness, M. D., pp. 85–102. Berkeley, CA: University of California Press.Google Scholar
Herman, K. D., Masters, L. A., Penskar, M. R., Reznicek, A. A., Wilhelm, G. S., Brodovich, W. W., and Gardiner, K. P. (2001). Floristic Quality Assessment with Wetland Categories and Examples of Computer Applications for the State of Michigan, revd 2nd edn. Lansing, MI: Natural Heritage Program, Michigan Department of Natural Resources.Google Scholar
Higgs, E. S. (1997). What is good ecological restoration?Conservation Biology, 11, 338–48.CrossRefGoogle Scholar
Hik, D. S., Jefferies, R. L., and Sinclair, A. R. E. (1992). Foraging by geese, isostatic uplift and asymmetry in the development of salt-marsh plant communities. Journal of Ecology, 80, 395–406.CrossRefGoogle Scholar
Hill, N. M. and Keddy, P. A. (1992). Prediction of rarities from habitat variables: coastal plain plants on Nova Scotian lakeshores. Ecology, 73, 1852–9.CrossRefGoogle Scholar
Hill, N. M., Keddy, P. A., and Wisheu, I. C. (1998). A hydrological model for predicting the effects of dams on the shoreline vegetation of lakes and reservoirs. Environmental Management, 22, 723–36.CrossRefGoogle ScholarPubMed
Hoagland, B. W. and Collins, S. L. (1997a). Gradient models, gradient analysis, and hierarchical structure in plant communities. Oikos, 78, 23–30.CrossRefGoogle Scholar
Hoagland, B. W. and Collins, S. L. (1997b). Heterogeneity in shortgrass prairie vegetation: the role of playa lakes. Journal of Vegetation Science, 8, 277–86.CrossRefGoogle Scholar
Hochachka, P. W., Fields, J., and Mustafa, T. (1973). Animal life without oxygen: basic biochemical mechanisms. American Zoology, 13, 543–55.CrossRefGoogle Scholar
Hogenbirk, J. C. and Wein, R. W. (1991). Fire and drought experiments in northern wetlands: a climate change analogue. Canadian Journal of Botany, 69, 1991–7.CrossRefGoogle Scholar
Hogg, E. H., Lieffers, V. J., and Wein, R. W. (1992). Potential carbon losses from peat profiles: effects of temperature, drought cycles, and fire. Ecological Applications, 2, 298–306.CrossRefGoogle ScholarPubMed
Holling, , C. S. (ed.) (1978). Adaptive Environmental Assessment and Management. Chichester, UK: John Wiley.
Hook, D. D. (1984). Adaptations to flooding with fresh water. In Flooding and Plant Growth, ed. Kozlowski, T. T., pp. 265–94. Orlando, FL: Academic Press.CrossRefGoogle Scholar
Hook, , D. D., McKee, W. H., Smith, H., Gregory, J., Burrell, V. J., DeVoe, W. R., Sojka, R. E., Gilbert, S., Banks, R., Stolzy, L. H., Brooks, C., Matthews, T. D., and Shear, T. H. (eds.) (1988). The Ecology and Management of Wetlands, Vol. 1, Ecology of Wetlands. Portland, OR: Timber Press.
Hoover, J. J. and Killgore, K. J. (1998). Fish communities. In Southern Forested Wetlands: Ecology and Management, eds. Messina, M. G. and Conner, W. H., pp. 237–60. Boca Raton, FL: Lewis Publishers.Google Scholar
Horn, H. (1976). Succession. In Theoretical Ecology: Principles and Applications, ed. May, R. M., pp. 187–204. Philadelphia, PA: W.B. Saunders.Google Scholar
Hou, H.-Y. (1983). Vegetation of China with reference to its geographical distribution. Annals of the Missouri Botanical Garden, 70, 509–48.CrossRefGoogle Scholar
Houck, O. (2006). Can we save New Orleans?Tulane Environmental Law Journal, 19, 1–68.Google Scholar
Houlahan, J., Keddy, P., Makkey, K., and Findlay, C. S. (2006). The effects of adjacent land-use on wetland plant species richness and community composition. Wetlands, 26, 79–96.CrossRefGoogle Scholar
House, J. and Brovkin, V. (eds.) (2005). Climate and air quality. In Ecosystems and Human Well-Being: Current State and Trends – Findings of the Condition and Trends Working Group of the Millennium Ecosystem Assessment, eds. Hassan, R., Scholes, R., and Ash, N., pp. 355–90. Washington, DC: Island Press.Google Scholar
Howard, R. T. and Mendelssohn, I. A. (1999). Salinity as a constraint on growth of oligohaline marsh macrophytes. I. Species variation in stress tolerance. American Journal of Botany, 86, 785–94.CrossRefGoogle ScholarPubMed
Howard-Williams, C. and Thompson, K. (1985). The conservation and management of African wetlands. In The Ecology and Management of African Wetland Vegetation, ed. Denny, P., pp. 203–30. Dordrecht, the Netherlands: Dr. W. Junk Publishers.CrossRefGoogle Scholar
Howarth, R. W., Fruci, J. R., and Sherman, D. (1991). Inputs of sediment and carbon to an estuarine ecosystem: influence of land use. Ecological Applications, 1, 27–39.CrossRefGoogle Scholar
Hubbell, S. P. and Foster, R. B. (1986). Biology, chance, and the history and structure of tropical rain forest tree communities. In Community Ecology, eds. Diamond, J. and Case, T. J., pp. 314–29. New York: Harper and Row.Google Scholar
Huber, O. (1982). Significance of savanna vegetation in the Amazon Territory of Venezuela. In Biological Diversification in the Tropics, ed. Prance, G. T., pp. 221–44. New York: Columbia University Press.Google Scholar
Hughes, J. D. and Thirgood, J. V. (1982). Deforestation, erosion and forest management in ancient Greece and Rome. Journal of Forestry, 26, 60–75.Google Scholar
Hunter, M. D. and Price, P. W. (1992). Playing chutes and ladders: heterogeneity and the relative roles of bottom-up and top-down forces in natural communities. Ecology, 73, 724–32.Google Scholar
Hurlbert, S. H. (1984). Pseudoreplication and the design of ecological field experiments. Ecological Monographs, 54, 187–211.CrossRefGoogle Scholar
Hurlbert, S. H. (1990). Spatial distribution of the montane unicorn. Oikos, 58, 257–71.CrossRefGoogle Scholar
Huston, M. (1979). A general hypothesis of species diversity. The American Naturalist, 113, 81–101.CrossRefGoogle Scholar
Huston, M. (1994). Biological Diversity: The Coexistence of Species on Changing Landscapes. Cambridge, UK: Cambridge University Press.Google Scholar
Hutchinson, G. E. (1959). Homage to Santa Rosalia or why are there so many kinds of animals?The American Naturalist, 93, 145–9.CrossRefGoogle Scholar
Hutchinson, G. E. (1975). A Treatise on Limnology, Vol. 3, Limnological Botany. New York: John Wiley.Google Scholar
Ingebritsen, S. E., McVoy, C., Glaz, B., and Park, W. (1999). Florida Everglades: subsidence threatens agriculture and complicates ecosystem restoration. In Land Subsidence in the United States, U.S. Geological Survey Circular No. 1182, eds. Galloway, D., Jones, D. R., and Ingebritsen, S. E., pp. 95–106. Reston, VA: U.S. Geological Survey.Google Scholar
Ingram, H. A. P. (1982). Size and shape in raised mire ecosystems: a geophysical model. Nature, 297, 300–3.CrossRefGoogle Scholar
Ingram, H. A. P. (1983). Hydrology. In Ecosystems of the World, Vol. 4A, Mires: Swamp, Bog, Fen and Moor – General Studies, ed. Gore, A. J. P., pp. 67–158. Amsterdam, the Netherlands: Elsevier.Google Scholar
,International Joint Commission. (1980). Pollution in the Great Lakes Basin from Land Use Activities. Detroit, MI and Windsor, ON: International Joint Commission.
,International Rice Research Institute (IRRI). (2009). Rough rice consumption, by country and geographical region: USA. http://beta.irri.org/solutions/index.php? (accessed Dec 4, 2009)
Irion, G. M., Müller, J., Mello, J. N., and Junk, W. J. (1995). Quaternary geology of the Amazon lowland. Geo-Marine Letters, 15, 172–8.CrossRefGoogle Scholar
Isabelle, P. S., Fooks, L. J., Keddy, P. A., and Wilson, S. D. (1987). Effects of roadside snowmelt on wetland vegetation: an experimental study. Journal of Environmental Management, 25, 57–60.Google Scholar
,IUCN. (2008) Red List. www.iucnredlist.org
Jackson, J. B. C. (1981). Interspecific competition and species distributions: the ghosts of theories and data past. American Zoologist, 21, 889–901.CrossRefGoogle Scholar
Jackson, M. B. and Drew, M. C. (1984). Effects of flooding on growth and metabolism of herbaceous plants. In Flooding and Plant Growth, ed. Kozlowski, T. T., pp. 47–128. Orlando, FL: Academic Press.CrossRefGoogle Scholar
Janis, C. (1976). The evolutionary strategy of the Equidae and the origins of rumen and cecal digestion. Evolution, 30, 757–74.CrossRefGoogle ScholarPubMed
Janzen, D. H. and Martin, P. S. (1982). Neotropical anachronisms: the fruits the gomphotheres ate. Science, 215, 19–27.CrossRefGoogle ScholarPubMed
Jean, M. and Bouchard, A. (1991). Temporal changes in wetland landscapes of a section of the St. Lawrence River, Canada. Environmental Management, 15, 241–50.CrossRefGoogle Scholar
Jefferies, R. L. (1977). The vegetation of salt marshes at some coastal sites in arctic North America. Journal of Ecology, 65, 661–72.CrossRefGoogle Scholar
Jefferies, R. L. (1988a). Pattern and process in Arctic coastal vegetation in response to foraging by lesser snow geese. In Plant Form and Vegetation Structure, eds. Werger, M. J. A., Aart, P. J. M., During, H. J., and Verhoeven, J. T. A., pp. 281–300. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Jefferies, R. L. (1988b). Vegetational mosaics, plant-animal interactions and resources for plant growth. In Plant Evolutionary Biology, eds. Gottlieb, L. and Jain, S. K., pp. 341–69. London: Chapman and Hall.Google Scholar
Jeglum, J. K. and He, F. (1995). Pattern and vegetation–environment relationships in a boreal forested wetland in northeastern Ontario, Canadian Journal of Botany, 73, 629–37.CrossRefGoogle Scholar
Jenkins, S. H. (1975). Food selection by beavers. Oecologia, 21, 157–73.CrossRefGoogle ScholarPubMed
Jenkins, S. H. (1979). Seasonal and year to year differences in food selection by beavers. Oecologia, 44, 112–16.CrossRefGoogle ScholarPubMed
Jenkins, S. H. (1980). A size–distance relation in food selection by beavers. Ecology, 61, 740–6.CrossRefGoogle Scholar
Jochimsen, D. M. (2006). Factors influencing the road mortality of snakes on the Upper Snake River Plain, Idaho. In Proceedings of the 2005 International Conference on Ecology and Transportation, eds. Irwin, C. L., Garrett, P., and McDermott, K. P., pp. 351–65. Raleigh, NC: Center for Transportation and the Environment, North Carolina State University.Google Scholar
Johnsgard, P. A. (1980). Where have all the curlews gone?Natural History, 89(8), 30–3. Reprinted in Papers in Ornithology, http://digitalcommons.unl.edu/Gioscioenithology/23Google Scholar
Johnson, D. L., Lynch, W. E., Jr., and Morrison, T. W. (1997). Fish communities in a diked Lake Erie wetland and an adjacent undiked area. Wetlands, 17, 43–54.CrossRefGoogle Scholar
Johnson, M. G., Leach, J. H., Minns, C. K., and Oliver, C. H. (1977). Limnological characteristics of Ontario lakes in relation to associations of walleye (Stizostedion vitreum), northern pike (Esox lucius), lake trout (Salvelinus namaycush) and smallmouth bass (Micropterus dolomieui). Journal of the Fisheries Research Board of Canada, 34, 1592–601.CrossRefGoogle Scholar
Johnson, P. D. and Brown, K. M. (1998). Intraspecific life history variation in the threatened Louisiana pearlshell mussel, Margaritifera hembeli. Freshwater Biology, 40, 317–29.CrossRefGoogle Scholar
Johnson, W. B., Sasser, C. E., and Gosselink, J. G. (1985). Succession of vegetation in an evolving river delta, Atchafalaya Bay, Louisiana. Journal of Ecology, 73, 973–86.CrossRefGoogle Scholar
Johnson, W. C. (1994). Woodland expansion in the Platte River, Nebraska: patterns and causes. Ecological Monographs, 64, 45–84.CrossRefGoogle Scholar
Johnson, W. C., Burgess, R. L., and Keammerer, W. R. (1976). Forest overstory vegetation and environment on the Missouri River floodplain in North Dakota. Ecological Monographs, 46, 59–84.CrossRefGoogle Scholar
Johnston, A. J. B. (1983). The Summer of 1744: A Portrait of Life in 18th-Century Louisbourg. Hull, QC: Parks Canada.Google Scholar
Johnston, C. A. and Naiman, R. J. (1990). Aquatic patch creation in relation to beaver population trends. Ecology, 71, 1617–21.CrossRefGoogle Scholar
Johnston, J. W., Thompson, T. A., Wilcox, D. A., and Baedke, S. J. (2007). Geomorphic and sedimentologic evidence for the separation of Lake Superior from Lake Michigan and Huron. Journal of Paleolimnology, 37, 349–64.CrossRefGoogle Scholar
Jones, C. G., Lawton, J. H., and Shachak, M. (1994). Organisms as ecosystem engineers. Oikos, 69, 373–86.CrossRefGoogle Scholar
Jones, M. (2003). The Last Great Quest: Captain Scott's Antarctic Sacrifice. New York: Oxford University Press.Google Scholar
Jones, R. H., Sharitz, R. R., Dixon, P. M., Segal, D. S., and Schneider, R. L. (1994). Woody plant regeneration in four floodplain forests. Ecological Monographs, 64, 345–67.CrossRefGoogle Scholar
Jordan, W. R., I I I, Gilpin, M. E., and Aber, J. D. (1987). Restoration Ecology: Synthetic Approach to Ecological Research. Cambridge, UK: Cambridge University Press.Google Scholar
Judson, S. (1968). Erosion of the land, or what's happening to our continents?American Scientist, 56, 356–74.Google Scholar
Junk, W. J. (1983). Ecology of swamps on the Middle Amazon. In Ecosystems of the World, Vol. 4B, Mires:Swamp, Bog, Fen and Moor – Regional Studies, ed. Gore, A. J. P., pp. 98–126. Amsterdam, the Netherlands: Elsevier.Google Scholar
Junk, W. J. (1984). Ecology of the várzea, floodplain of Amazonian white-water rivers. In The Amazon: Limnology and Landscape Ecology of a Mighty Tropical River and its Basin, ed. Sioli, H., pp. 215–43. Dordrecht, the Netherlands: Dr. W. Junk Publishers.CrossRefGoogle Scholar
Junk, W. J. (1986). Aquatic plants of the Amazon system. In The Ecology of River Systems, eds. Davies, B. R. and Walker, K. F., pp. 319–37. Dordrecht, the Netherlands: Dr. W. Junk Publishers.Google Scholar
Junk, W. J. (1993). Wetlands of tropical South America. In Wetlands of the World, Vol. 1, eds. Whigham, D. F., D. Dykyjova and S. Hejny, pp. 679–739. Dordrecht, the Netherlands: Kluwer.Google Scholar
Junk, W. J. and Piedade, M. T. F. (1994). Species diversity and distribution of herbaceous plants in the floodplain of the middle Amazon. Verhandlungen Internationale Vereinigung für theoretische und angewandte Limnologie, 25, 1862–5.Google Scholar
Junk, W. J. and Piedade, M. T. F. (1997). Plant life in the floodplain with special reference to herbaceous plants. In The Central Amazon Floodplain, ed. Junk, W. J., pp. 147–85. Berlin, Germany: Springer-Verlag.CrossRefGoogle Scholar
Junk, W. J. and Welcomme, R. L. (1990). Floodplains. In Wetlands and Shallow Continental Water Bodies, Vol.1, Natural and Human Relationships, ed. Patten, B. C., pp. 491–524. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Junk, W. J., Bayley, P. B., and Sparks, R. E. (1989). The flood pulse concept in riverfloodplain systems. In Proceedings of the International Large River Symposium, ed. Dodge, D. P., pp. 110–27. Canadian Journal of Fisheries and Aquatic Sciences, Special Publication No. 106.Google Scholar
Junk, W. J., Soares, M. G. M., and Saint-Paul, U. (1997). The fish. In The Central Amazon Floodplain, ed. W. J. Junk, pp. 385–408. Berlin, Germany: Springer-Verlag.CrossRefGoogle Scholar
Junk, W. J., Brown, M., Campbell, I. C., Finlayson, M., Gopal, B., Ramberg, L., and Warner, B. G. (2006). The comparative biodiversity of seven globally important wetlands: a synthesis. Aquatic Sciences, 68, 400–14.CrossRefGoogle Scholar
Jurik, T. M., Wang, S., and Valk, A. G. (1994). Effects of sediment load on seedling emergence from wetland seed banks. Wetlands, 14, 159–65.CrossRefGoogle Scholar
Justin, S. H. F. W. and Armstrong, W. (1987). The anatomical characteristics of roots and plant response to soil flooding. New Phytologist, 106, 465–95.CrossRefGoogle Scholar
Kajak, Z. (1993). The Vistula River and its riparian zones. Hydrobiologia, 251, 149–57.CrossRefGoogle Scholar
Kalamees, K. (1982). The composition and seasonal dynamics of fungal cover on peat soils. In Peatland Ecosystems: Researches into the Plant Cover of Estonian Bogs and Their Productivity, ed. Masing, V., pp. 12–29. Tallinn, Estonia: Academy of Sciences of the Estonian S. S. R.Google Scholar
Kalliola, R., Salo, J., Puhakka, M., and Rajasilta, M. (1991). New site formation and colonizing vegetation in primary succession on the Western Amazon floodplains. Journal of Ecology, 79, 877–901.CrossRefGoogle Scholar
Kaminski, R. M. and Prince, H. H. (1981). Dabbling duck and aquatic macroinvertebrate responses to manipulated wetland habitat. Journal of Wildlife Management, 45, 1–15.CrossRefGoogle Scholar
Kaminski, R. M., Murkin, H. M., and Smith, C. E. (1985). Control of cattail and bulrush by cutting and flooding. In Coastal Wetlands, eds. Prince, H. H. and D'Itri, F. M., pp. 253–62. Chelsea, MI: Lewis Publishers.Google Scholar
Kantrud, H. A., Millar, J. B., and Valk, A. G. (1989). Vegetation of the wetlands of the prairie pothole region. In Northern Prairie Wetlands, ed. Valk, A. G., pp. 132–87. Ames, IA: Iowa State University Press.Google Scholar
Karrow, , P. F. and Calkin, P. E. (eds.) (1985). Quaternary Evolution of the Great Lakes, Special Paper No. 30. St John's, Nfld: Geological Association of Canada.
Keddy, C. J. and McCrae, T. (1989). Environmental Databases for State of the Environment Reporting, Technical Report No. 19. Ottawa, ON: State of the Environment Reporting Branch, Environment Canada.Google Scholar
Keddy, C. J. and Sharp, M. J. (1994). A protocol to identify and prioritize significant coastal plain plant assemblages for conservation. Biological Conservation, 68, 269–74.CrossRefGoogle Scholar
Keddy, P. A. (1976). Lakes as islands: the distributional ecology of two aquatic plants, Lemna minor L. and L. trisulca L. Ecology, 57, 353–9.CrossRefGoogle Scholar
Keddy, P. A. (1981). Vegetation with coastal plain affinities in Axe Lake, near Georgian Bay, Ontario. Canadian Field Naturalist, 95, 241–8.Google Scholar
Keddy, P. A. (1982). Quantifying within lake gradients of wave energy, substrate particle size and shoreline plants in Axe Lake, Ontario. Aquatic Botany, 14, 41–58.CrossRefGoogle Scholar
Keddy, P. A. (1983). Shoreline vegetation in Axe Lake, Ontario: effects of exposure on zonation patterns. Ecology, 64, 331–44.CrossRefGoogle Scholar
Keddy, P. A. (1984). Plant zonation on lakeshores in Nova Scotia: a test of the resource specialization hypothesis. Journal of Ecology, 72, 797–808.CrossRefGoogle Scholar
Keddy, P. A. (1985a). Lakeshores in the Tusket River Valley, Nova Scotia: distribution and status of some rare species, including Coreopsis rosea Nutt. and Sabtia kennedyana Fern. Rhodora, 87, 309–20.Google Scholar
Keddy, P. A. (1985b). Wave disturbance on lakeshores and the within-lake distribution of Ontario's Atlantic coastal plain flora. Canadian Journal of Botany, 63, 656–60.CrossRefGoogle Scholar
Keddy, P. A. (1989a). Competition. London: Chapman and Hall.CrossRefGoogle Scholar
Keddy, P. A. (1989b). Effects of competition from shrubs on herbaceous wetland plants: a 4-year field experiment. Canadian Journal of Botany, 67, 708–16.CrossRefGoogle Scholar
Keddy, P. A. (1990a). Competitive hierarchies and centrifugal organization in plant communities. In Perspectives on Plant Competition, eds. Grace, J. B. and Tilman, D., pp. 265–90. San Diego, CA: Academic Press.Google Scholar
Keddy, P. A. (1990b). Is mutualism really irrelevant to ecology?Bulletin of the Ecological Society of America, 71(2), 101–2.Google Scholar
Keddy, P. A. (1991a). Biological monitoring and ecological prediction: from nature reserve management to national state of environment indicators. In Biological Monitoring for Conservation, ed. Goldsmith, F. B., pp. 249–67. London: Chapman and Hall.Google Scholar
Keddy, P. A. (1991b). Water level fluctuations and wetland conservation. In Wetlands of the Great Lakes, eds. Kusler, J. and Smardon, R., pp. 79–91. Proceedings of an International Symposium, Niagara Falls, NY, May 16–18, 1990. Berne, NY: Association of State Wetland Managers.Google Scholar
Keddy, P. A. (1991c). Reviewing a festschrift: what are we doing with our scientific lives?Journal of Vegetation Science, 2, 419–24.Google Scholar
Keddy, P. A. (1992a). Assembly and response rules: two goals for predictive community ecology. Journal of Vegetation Science, 3, 157–64.CrossRefGoogle Scholar
Keddy, P. A. (1992b). A pragmatic approach to functional ecology. Functional Ecology, 6, 621–6.CrossRefGoogle Scholar
Keddy, P. A. (1994). Applications of the Hertzsprung–Russell star chart to ecology: reflections on the 21st birthday of Geographical Ecology. Trends in Ecology and Evolution, 9, 231–4.CrossRefGoogle Scholar
Keddy, P. A. (2001). Competition, 2nd edn. Dordrecht, the Netherlands: Kluwer.CrossRefGoogle Scholar
Keddy, P. A. (2007). Plants and Vegetation: Origins, Processes, Consequences. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Keddy, P. A. (2009). Thinking big: a conservation vision for the Southeastern Coastal Plain of North America. Southeastern Naturalist, 7, 213–26.CrossRefGoogle Scholar
Keddy, P. A. and Constabel, P. (1986). Germination of ten shoreline plants in relation to seed size, soil particle size and water level: an experimental study. Journal of Ecology, 74, 122–41.CrossRefGoogle Scholar
Keddy, P. A. and Fraser, L. H. (2000). Four general principles for the management and conservation of wetlands in large lakes: the role of water levels, nutrients, competitive hierarchies and centrifugal organization. Lakes and Reservoirs: Research and Management, 5, 177–85.CrossRefGoogle Scholar
Keddy, P. A. and Fraser, L. H. (2002). The management of wetlands for biological diversity: four principles. In Modern Trends in Applied Aquatic Ecology, eds. Ambasht, R. S. and Ambasht, N. K., pp. 21–42. New York: Kluwer.Google Scholar
Keddy, P. A. and MacLellan, P. (1990). Centrifugal organization in forests. Oikos, 59, 75–84.CrossRefGoogle Scholar
Keddy, P. A. and Reznicek, A. A. (1982). The role of seed banks in the persistence of Ontario's coastal plain flora. American Journal of Botany, 69, 13–22.CrossRefGoogle Scholar
Keddy, P. A. and Reznicek, A. A. (1986). Great Lakes vegetation dynamics: the role of fluctuating water levels and buried seeds. Journal of Great Lakes Research, 12, 25–36.CrossRefGoogle Scholar
Keddy, P. A. and Shipley, B. (1989). Competitive hierarchies in herbaceous plant communities. Oikos, 54, 234–41.CrossRefGoogle Scholar
Keddy, P. A. and Wisheu, I. C. (1989). Ecology, biogeography, and conservation of coastal plain plants: some general principles from the study of Nova Scotian wetlands. Rhodora, 91, 72–94.Google Scholar
Keddy, P. A., Lee, H. T., and Wisheu, I. C. (1993). Choosing indicators of ecosystem integrity: wetlands as a model system. In Ecological Integrity and the Management of Ecosystems, eds. Woodley, S., Kay, J., and Francis, G., pp. 61–79. Delray Beach, FL: St. Lucie Press.Google Scholar
Keddy, P. A., Twolan-Strutt, L., and Wisheu, I. C. (1994). Competitive effect and response rankings in 20 wetland plants: are they consistent across three environments?Journal of Ecology, 82, 635–43.CrossRefGoogle Scholar
Keddy, P. A., Fraser, L. H., and Wisheu, I. C. (1998). A comparative approach to examine competitive responses of 48 wetland plant species. Journal of Vegetation Science, 9, 777–86.CrossRefGoogle Scholar
Keddy, P. A., Campbell, D., McFalls, T., Shaffer, G., Moreau, R., Dranguet, C., and Heleniak, R. (2007). The wetlands of lakes Pontchartrain and Maurepas: past, present and future. Environmental Reviews, 15, 1–35.CrossRefGoogle Scholar
Keddy, P. A., Gough, L., Nyman, J. A., McFalls, T., Carter, J., and Siegnist, J. (2009a). Alligator hunters, pelt traders, and runaway consumption of Gulf coast marshes: a trophic cascade perspective on coastal wetland losses. In Human Impacts on Salt Marshes: A Global Perspective, eds. Silliman, B. R., Grosholz, E. D., and Bertness, M. D., pp. 115–33. Berkeley, CA: University of California Press.Google Scholar
Keddy, P. A., Fraser, L. H., Solomeshch, A. I., Junk, W. J., Campbell, D. R., Arroyo, M. T. K., and Alho, C. J. R. (2009b). Wet and wonderful: the world's largest wetlands are conservation priorities. BioScience, 59, 39–51.CrossRefGoogle Scholar
Keeley, J. E., DeMason, D. A., Gonzalez, R., and Markham, K. R. (1994). Sediment based carbon nutrition in tropical alpine Isoetes. In Tropical Alpine Environments Plant Form and Function, eds. Rundel, P. W., Smith, A. P., and Meinzer, F. C., pp. 167–94. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Keeling, C. D. and Whorf, T. P. (2005). Atmospheric CO2 records from sites in the SIO air sampling network. In Trends: A Compendium of Data on Global Change eds. T.A. Boden et al., pp. 16–26. Oak Ridge, TN: Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy.Google Scholar
Keller, E. A. and Day, J. W. (2007). Untrammeled growth as an environmental “March of Folly”. Ecological Engineering, 30, 206–14.CrossRefGoogle Scholar
Kelly, K. (1975). The artificial drainage of land in nineteenth-century southern Ontario. Canadian Geographer, 4, 279–98.CrossRefGoogle Scholar
Kendall, R. L. (1969). An ecological history of the Lake Victoria Basin. Ecological Monographs, 39, 121–76.CrossRefGoogle Scholar
Kenrick, P. and Crane, P. R. (1997). The Origin and Early Diversification of Land Plants: A Cladistic Study. Washington, DC: Smithsonian Institution Press.Google Scholar
Keogh, T. M., Keddy, P. A., and Fraser, L. H. (1998). Patterns of tree species richness in forested wetlands. Wetlands, 19, 639–47.CrossRefGoogle Scholar
Kercher, S. M., Carpenter, Q. J., and Zedler, J. B. (2004). Interrelationships of hydrologic disturbance, reed canary grass (Phalaris arundinacea L.), and native plants in Wisconsin wet meadows. Natural Areas Journal, 24, 316–25.Google Scholar
Kerr, R. A. (2006). A worrying trend of less ice, higher seas. Science, 311, 1698–701.CrossRefGoogle ScholarPubMed
Kershaw, K. A. (1962). Quantitative ecological studies from Landmannahellir, Iceland. Journal of Ecology, 50, 171–9.CrossRefGoogle Scholar
Kershner, J. L. (1997). Setting riparian/aquatic restoration objectives within a watershed context. Restoration Ecology, 5, 15–24.CrossRefGoogle Scholar
Kirby, M. X. (2004). Fishing down the coast: historical expansion and collapse of oyster fisheries along continental margins. Proceedings of the National Academy of Sciences of the USA, 101, 13 096–99.CrossRefGoogle ScholarPubMed
Kirk, K. L. and Gilbert, J. J. (1990). Suspended clay and the population dynamics of planktonic rotifers and cladocerans. Ecology, 71, 1741–55.CrossRefGoogle Scholar
Klimas, C. V. (1988). River regulation effects on floodplain hydrology and ecology. In The Ecology and Management of Wetlands, Vol. 1, Ecology of Wetlands, eds. Hook, D. D., McKee, Jr. W. H., Smith, H. K., Gregory, J., Burrell, Jr. V. G., DeVoe, M. R., Sojka, R. E., Gilbert, S., Banks, R., Stolzy, L. H., Brooks, C., Matthews, T. D., and Shear, T. H., pp. 40–9. Portland, OR: Timber Press.Google Scholar
Knight, R. L. and Kadlec, R. H. (2004). Treatment Wetlands. Boca Raton, FL: Lewis Publishers.Google Scholar
Koerselman, W. and Verhoeven, J. T. A. (1995). Eutrophication of fen ecosystems: external and internal nutrient sources and restoration strategies. In Restoration of Temperate Wetlands, eds. Wheeler, S., Shaw, S., Fojt, W., and Robertson, R., pp. 91–112. Chichester, UK: John Wiley.Google Scholar
Kozlowski, , T. T. (ed.) (1984a). Flooding and Plant Growth. Orlando, FL: Academic Press.
Kozlowski, T. T. (1984b). Responses of woody plants to flooding. In Flooding and Plant Growth, ed. Kozlowski, T. T., pp. 129–63. Orlando, FL: Academic Press.CrossRefGoogle Scholar
Kozlowski, T. T. and Pallardy, S. G. (1984). Effect of flooding on water, carbohydrate, and mineral relations. In Flooding and Plant Growth, ed. Kozlowski, T. T., pp. 165–93. Orlando, FL: Academic Press.CrossRefGoogle Scholar
Kramer, D. L., Lindsay, C. C., Moodie, G. E. E., and Stevens, E. D. (1978). The fishes and the aquatic environment of the Central Amazon basin, with particular reference to respiratory patterns. Canadian Journal of Zoology, 56, 717–29.CrossRefGoogle Scholar
Krieger, J. (2001). The Economic Value of Forest Ecosystem Services: A Review. Washington, DC: The Wilderness Society.Google Scholar
Kreutzwiser, R. D. (1981). The economic significance of the Long Point marsh, Lake Erie, as a recreational resource. Journal of Great Lakes Research, 7, 105–10.CrossRefGoogle Scholar
Kuhry, P. (1994). The role of fire in the development of Sphagnum-dominated peatlands in western boreal Canada. Journal of Ecology, 82, 899–910.CrossRefGoogle Scholar
Kuhry, P., Nicholson, B. J., Gignac, L. D., Vitt, D. H., and Bayley, S. E. (1993). Development of Sphagnum-dominated peatlands in boreal continental Canada. Canadian Journal of Botany, 71, 10–22.CrossRefGoogle Scholar
Kurihara, Y. and Kikkawa, J. (1986). Trophic relations of decomposers. In Community Ecology: Pattern and Process, eds. Kikkawa, J. and Anderson, D. J., pp. 127–60. Melbourne, Vic: Blackwell Scientific Publications.Google Scholar
Kurimo, H. (1984). Simultaneous groundwater table fluctuation in different parts of the Virgin Pine Mires. Silva Fennica, 18, 151–86.CrossRefGoogle Scholar
Kurtén, B. and Anderson, E. (1980). Pleistocene Mammals of North America. New York: Columbia University Press.Google Scholar
Kusler, , J. A. and Kentula, M. E. (eds.) (1990). Wetland Creation and Restoration: Status of the Science. Washington, DC: Island Press.
Kusler, , J. A., Willard, D. E., and Hull, H. C., Jr. (eds.) (1995). Wetlands and Watershed Management: Science Applications and Public Policy. A collection of papers from a national symposium and several workshops at Tampa, FL, Apr 23–26. Berne, NY: Association of State Wetland Managers.
LaBaugh, J. W. (1989). Chemical characteristics of water in northern prairie wetlands. In Northern Prairie Wetlands, ed. Valk, A. G., pp. 56–90. Ames, IA: Iowa State University Press.Google Scholar
Laing, H. E. (1940). Respiration of the rhizomes of Nuphar advenum and other water plants. American Journal of Botany, 27, 574–81.CrossRefGoogle Scholar
Laing, H. E. (1941). Effect of concentration of oxygen and pressure of water upon growth of rhizomes of semi-submerged water plants. Botanical Gazette, 102, 712–24.CrossRefGoogle Scholar
Lane, P. A. (1985). A food web approach to mutualism in lake communities. In The Biology of Mutualism: Ecology and Evolution, ed. Boucher, D. H., pp. 344–74. New York: Oxford University Press.Google Scholar
Larcher, W. (1995). Physiological Plant Ecology: Ecophysiology and Stress Physiology of Functional Groups, 3rd edn. New York: Springer-Verlag.CrossRefGoogle Scholar
Laroche, F. B. and Baker, G. E. (2001). Vegetation management within the Everglades protection area. In 2001 Everglades Consolidated Report, Appendix 14. Miami, FL: South Florida Water Management District.Google Scholar
Larson, D. W. (1996). Brown's Woods: an early gravel pit forest restoration project, Ontario, Canada. Restoration Ecology, 4, 11–18.CrossRefGoogle Scholar
Larson, J. S. (1988). Wetland creation and restoration: an outline of the scientific perspective. In Increasing our Wetland Resources, eds. Zelazny, J. and Feierabend, J. S., pp. 73–9. Proceedings of a conference in Washington, DC, Oct 4–7, 1987. Reston, VA: National Wildlife Federation–Corporate Conservation Council.Google Scholar
Larson, J. S. (1990). Wetland value assessment. In Wetlands and Shallow Continental Water Bodies, Vol. 1, Natural and Human Relationships, ed. Patten, B. C., pp. 389–400. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Larson, J. S., Mueller, A. J., and MacConnell, W. P. (1980). A model of natural and man-induced changes in open freshwater wetlands on the Massachusetts coastal plain. Journal of Applied Ecology, 17, 667–73.CrossRefGoogle Scholar
Latham, P. J., Pearlstine, L. G., and Kitchens, W. M. (1994). Species association changes across a gradient of freshwater, oligohaline, and mesohaline tidal marshes along the lower Savannah River. Wetlands, 14, 174–83.CrossRefGoogle Scholar
Latham, R. E. and Ricklefs, R. E. (1993). Continental comparisons of temperatezone tree species diversity. In Species Diversity in Ecological Communities: Historical and Geographical Perspectives, eds. Ricklefs, R. E. and Schluter, D., pp. 294–314. Chicago, IL: University of Chicago Press.Google Scholar
Laubhan, M. K. (1995). Effects of prescribed fire on moist-soil vegetation and soil macronutrients. Wetlands, 15, 159–66.CrossRefGoogle Scholar
Lavelle, P., Dugdale, R., and Scholes, R. (eds.) (2005). Nutrient cycling. In Ecosystems and Human Well-being: Current State and Trends – Findings of the Condition and Trends Working Group of the Millennium Ecosystem Assessment, eds. Hassan, R., Scholes, R., and Ash, N., pp. 331–53. Washington, DC: Island Press.Google Scholar
Lavoisier, A. (1789). Elements of Chemistry. In Great Books of the Western World, 2nd edn, 1990, ed. chief Adler, M. J., pp. 1–33. Chicago, IL: Encyclopaedia Britannica Inc.Google Scholar
Lawler, A. (2005). Reviving Iraq's wetlands. Science, 307, 1186–9.CrossRefGoogle ScholarPubMed
Leary, R. A. (1985). A framework for assessing and rewarding a scientist's research productivity. Scientometrics, 7, 29–38.CrossRefGoogle Scholar
Leck, M. A. and Graveline, K. J. (1979). The seed bank of a freshwater tidal marsh. American Journal of Botany, 66, 1006–15.CrossRefGoogle Scholar
Leck, , M. A., Parker, V. T., and Simpson, R. L. (eds.) (1989). Ecology of Soil Seed Banks. San Diego, CA: Academic Press.
Lee, R. (1980). Forest Hydrology. New York: Columbia University Press.Google Scholar
Legendre, L. and Legendre, P. (1983). Numerical Ecology. Amsterdam, the Netherlands: Elsevier.Google Scholar
Leitch, J. A. (1989). Politicoeconomic overview of prairie potholes. In Northern Prairie Wetlands, ed. Valk, A., pp. 2–14. Ames, IA: Iowa State University Press.Google Scholar
Leith, H. (1975). Historical survey of primary productivity research. In Primary Productivity of the Biosphere, eds. Leith, H. and Whittaker, R. H., pp. 7–16. New York: Springer-Verlag.CrossRefGoogle Scholar
Lemly, A. D. (1982). Modification of benthic insect communities in polluted streams: combined effects of sedimentation and nutrient enrichment. Hydrobiologia, 87, 229–45.CrossRefGoogle Scholar
Lent, R. M., Weiskel, P. K., Lyford, F. P., and Armstrong, D. S. (1997). Hydrologic indices for nontidal wetlands. Wetlands, 17, 19–30.CrossRefGoogle Scholar
Leonard, M. L. and Picman, J. (1986). Why are nesting marsh wrens and yellow-headed blackbirds spatially segregated?Auk, 103,135–40.Google Scholar
Leopold, A. (1949). A Sand County Almanac. New York: Oxford University Press.Google Scholar
Page, C. and Keddy, P. A. (1998). Reserves of buried seeds in beaver ponds. Wetlands, 18, 242–8.CrossRefGoogle Scholar
Lévêque, C., Balian, E. V., and Martens, K. (2005). An assessment of animal species diversity in continental waters. Hydrobiologia, 542, 39–67.CrossRefGoogle Scholar
Levin, H. L. (1992). The Earth Through Time, 4th edn. Forth Worth, TX: Saunders College Publishing.Google Scholar
Levine, J., Brewer, J. S., and Bertness, M. D. (1998). Nutrients, competition and plant zonation in a New England salt marsh. Journal of Ecology, 86, 285–92.CrossRefGoogle Scholar
Levitt, J. (1977). The nature of stress injury and resistance. In Responses of Plants to Environmental Stresses, ed. Levitt, J., pp. 11–21. New York: Academic Press.Google Scholar
Levitt, J. (1980). Responses of Plants to Environmental Stresses, Vols. 1 and 2, 2nd edn. New York: Academic Press.Google Scholar
Lewis, D. H. (1987). Evolutionary aspects of mutualistic associations between fungi and photosynthetic organisms. In Evolutionary Biology of Fungi, eds. Rayner, A. D. M., Brasier, C. M., and Moore, D., pp. 161–78. Cambridge, UK: Cambridge University Press.Google Scholar
Lewis, , , R. R. (ed.) (1982). Creation and Restoration of Coastal Plant Communities. Boca Raton, FL: CRC Press.
Lieffers, V. J. (1984). Emergent plant communities of oxbow lakes in northeastern Alberta: salinity, water-level fluctuation, and succession. Canadian Journal of Botany, 62, 310–16.CrossRefGoogle Scholar
Liu, K. and Fearn, M. L. (2000). Holocene history of catastrophic hurricane landfalls along the Gulf of Mexico coast reconstructed from coastal lake and marsh sediments. In Current Stresses and Potential Vulnerabilities: Implications of Global Change for the Gulf Coast Region of the United States, eds. Ning, Z. H. and Abdollhai, K. K., pp. 38–47. Baton Rouge, LA: Franklin Press for Gulf Coast Regional Climate Change Council.Google Scholar
Llewellyn, D. W., Shaffer, G. P., Craig, N. J., Creasman, L., Pashley, D., Swan, M., and Brown, C. (1996). A decision-support system for prioritizing restoration sites on the Mississippi River alluvial plain. Conservation Biology, 10, 1446–55.CrossRefGoogle Scholar
Lockwood, J. L and Pimm, S. L. (1999). When does restoration succeed? In Ecological Assembly Rules: Perspectives, Advances, Retreats, eds. Weiher, E. and Keddy, P., pp. 363–92. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Lodge, D. M. (1991). Herbivory on freshwater macrophytes. Aquatic Botany, 41, 195–224.CrossRefGoogle Scholar
Loffler, H. and Malkhazova, S. (1990). Impacts of wetlands on man. In Wetlands and Shallow Continental Water Bodies, Vol. 1, Natural and Human Relationships, ed. Patten, B. C., pp. 347–62. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Loope, L., Duever, M., Herndon, A., Snyder, J., and Jansen, D. (1994). Hurricane impact on uplands and freshwater swamp forest. BioScience, 44, 238–46.CrossRefGoogle Scholar
Louda, S. and Mole, S. (1991). Glucosinolates: chemistry and ecology. In Herbivores: Their Interactions with Secondary Plant Metabolites, eds. Rosenthal, G. A. and , M. R.Berenbaum, , pp. 124–64. San Diego, CA: Academic Press.Google Scholar
Louda, S. M., Keeler, K. H., and Holt, R. D. (1990). Herbivore influences on plant performance and competitive interactions. In Perspectives in Plant Competition, eds. J.B. Grace and D. Tilman, pp. 413–44. New York: Academic Press.Google Scholar
Loveless, C. M. (1959). A study of the vegetation in the Florida everglades. Ecology, 40, 1–9.CrossRefGoogle Scholar
Lowe-McConnell, R. H. (1975). Fish Communities in Tropical Freshwaters: Their Distribution, Ecology and Evolution. London: Longman.Google Scholar
Lowe-McConnell, R. H. (1987). Fish of the Amazon System. In The Ecology of River Systems, eds. Davies, B. R. and Walker, K. F., pp. 339–51. Dordrecht, the Netherlands: Dr. W. Junk Publishers.Google Scholar
Lowery, G. H. (1974). The Mammals of Louisiana and its Adjacent Waters. Baton Rouge, LA: Louisiana State University Press.Google Scholar
Lu, J. (1995). Ecological significance and classification of Chinese wetlands. Vegetatio, 118, 49–56.CrossRefGoogle Scholar
Lugo, A. E. and Brown, S. (1988). The wetlands of Caribbean islands. Acta Cientifica, 2, 48–61.Google Scholar
Lugo, A. E. and Snedaker, S. C. (1974). The ecology of mangroves. Annual Review of Ecology and Systematics, 5, 39–64.CrossRefGoogle Scholar
Lugo, A. E., Brown, S., and Brinson, M. M. (1988). Forested wetlands in freshwater and saltwater environments. Limnology and Oceanography, 33, 849–909.Google Scholar
Lugo, , A. E., Brinson, M. and Brown, S. (eds.) (1990). Forested Wetlands. Amsterdam, the Netherlands: Elsevier.
Lutman, J. (1978). The role of slugs in an Agrostis–Festuca grassland. In Production Ecology of British Moors and Montane Grasslands, eds. Heal, O. W. and Perkins, D. F., pp. 332–47. Berlin, Germany: Springer-Verlag.CrossRefGoogle Scholar
Lynch, J. A., Grimm, J. W., and Bowersox, V. C. (1995). Trends in precipitation chemistry in the United States: a national perspective, 1980–1992. Atmospheric Environment, 29, 1231–46.CrossRefGoogle Scholar
MacArthur, R. H. (1972). Geographical Ecology. New York: Harper and Row.Google Scholar
MacArthur, R. H. and MacArthur, J. (1961). On bird species diversity. Ecology, 42, 594–8.CrossRefGoogle Scholar
MacArthur, R. and Wilson, E. O. (1967). The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
MacRoberts, D. T., MacRoberts, B. R., and MacRoberts, M. H. (1997). A Floristic and Ecological Interpretation of the Freeman and Custis Red River Expedition of 1806. Shreueport, LA: Louisiana State University Press.Google Scholar
Magnuson, J. J., Regier, H. A., Christie, W. J., and Sonzongi, W. C. (1980). To rehabilitate and restore Great Lake ecosystems. In The Recovery Process in Damaged Ecosystems, ed. Cairns, Jr. J., pp. 95–112. Ann Arbor, MI: Ann Arbor Science Publishers.Google Scholar
Magnuson, J. J., Paszkowski, C. A., Rahel, F. J., and Tonn, W. M. (1989). Fish ecology in severe environments of small isolated lakes in northern Wisconsin. In Freshwater Wetlands and Wildlife, eds. Sharitz, R. and Gibbons, J. W., pp. 487–515. Conf-8603101, DOE Symposium Series No. 61. Oak Ridge, TN: Office of Scientific and Technical Information, U.S. Department of the Environment.Google Scholar
Maguire, L. A. (1991). Risk analysis for conservation biologists. Conservation Biology, 5, 123–5.CrossRefGoogle Scholar
Malmer, N. (1986). Vegetational gradients in relation to environmental conditions in northwestern European mires. Canadian Journal of Botany, 64, 375–83.CrossRefGoogle Scholar
Maltby, E. and Turner, R. E. (1983). Wetlands of the world. Geographical Magazine, 55, 12–17.Google Scholar
Maltby, E., Legg, C. J., and Proctor, C. F. (1990). The ecology of severe moorland fire on the North York Moors: effects of the 1976 fires, and subsequent surface and vegetation development. Journal of Ecology, 78, 490–518.CrossRefGoogle Scholar
Mandossian, A. and McIntosh, R. P. (1960). Vegetation zonation on the shore of a small lake. American Midland Naturalist, 64, 301–8.CrossRefGoogle Scholar
Mancil, E. (1980). Pullboat logging. Journal of Forest History, 24, 135–41.Google Scholar
Manfred, G. (1982). World Energy Supply. Berlin, Germany: Walter de Gruyter.Google Scholar
Mark, A. F., Johnson, P. N., Dickinson, K. J. M., and McGlone, M. S. (1995). Southern hemisphere pattered mires, with emphasis on southern New Zealand. Journal of the Royal Society of New Zealand, 25, 23–54.CrossRefGoogle Scholar
Marquis, R. J. (1991). Evolution of resistance in plants to herbivores. Evolutionary Trends in Plants, 5, 23–9.Google Scholar
Marschner, H. (1995). Mineral Nutrition of Higher Plants, 2nd edn. London: Academic Press.Google Scholar
Martin, P. S. and Klein, R. J. (1984). Quaternary Extinctions: A Prehistoric Revolution. Tucson, AZ: University of Arizona Press.Google Scholar
Martini, I. P. (1982). Introduction to scientific studies in Hudson and James Bay. Naturaliste Canadien, 109, 301–5.
Maseuth, J. D. (1995). Botany: An Introduction to Plant Biology, 2nd edn. Philadelphia, PA: Saunders College Publishing.Google Scholar
Matthews, E. and Fung, I. (1987). Methane emission from natural wetlands: global distribution, area, and environmental characteristics of sources. Global Biogeochemical Cycles, 1, 61–86.CrossRefGoogle Scholar
Matthews, W. J. (1998). Patterns in Freshwater Fish Ecology. New York: Chapman and Hall.CrossRefGoogle Scholar
Maun, M. A. and Lapierre, J. (1986). Effects of burial by sand on seed germination and seedling emergence of four dune species. American Journal of Botany, 73, 450–5.CrossRefGoogle Scholar
May, R. M. (1981). Patterns in multi-species communities. In Theoretical Ecology, ed. May, R. M., pp. 197–227. Oxford, UK: Blackwell Scientific Publications.Google Scholar
May, R. M. (1986). The search for patterns in the balance of nature: advances and retreats. Ecology, 67, 1115–26.CrossRefGoogle Scholar
May, R. M. (1988). How many species are there on Earth?Science, 241, 1441–9.CrossRefGoogle ScholarPubMed
Mayewski, P. A., Lyons, W. B., Spencer, M. J., Twickler, M. S., Buck, C. F., and Whitlow, S. (1990). An ice-core record of atmospheric response to anthropogenic sulphate and nitrate. Nature, 346, 554–6.CrossRefGoogle Scholar
Mayr, E. (1982). The Growth of Biological Thought: Diversity, Evolution, and Inheritance, Cambridge, MA: Belknap Press of Harvard University Press.Google Scholar
Mazzotti, F. J., Center, T. D., Dray, F. A. and Thayer, D. (1997). Ecological Consequences of Invasion by Melaleuca quinquenervia in Southern Florida Wetlands: Paradise Damaged, Not Lost. Gainesville, FL: University of Florida, Institute of Food and Agricultural Sciences.Google Scholar
McAuliffe, J. R. (1984). Competition for space, disturbance, and the structure of a benthic stream community. Ecology, 65, 894–908.CrossRefGoogle Scholar
McCanny, S. J., Keddy, P. A., Arnason, T. J., Gaudet, C. L., Moore, D. R. J., and Shipley, B. (1990). Fertility and the food quality of wetland plants: a test of the resource availability hypothesis. Oikos, 59, 373–81.CrossRefGoogle Scholar
McCarthy, K. A. (1987). Spatial and temporal distributions of species in two intermittent ponds in Atlantic County, NJ. M.Sc. thesis, Rutgers University, Rutgers, NJ.Google Scholar
McClure, J. W. (1970). Secondary constituents of aquatic angiosperms. In Phytochemical Phylogeny, ed. Harborne, J. B., pp. 233–65. New York: Academic Press.Google Scholar
McDougall, D. (2008). Global warning's front line. Guardian Weekly, Apr 11, p. 42.Google Scholar
McGeoch, M. A. and Gaston, K. J. (2002). Occupancy frequency distributions: patterns, artifacts and mechanism. Biological Reviews, 77, 311–31.CrossRefGoogle Scholar
McHarg, I. L. (1969). Design with Nature. Garden City, NJ: Natural History Press for American Museum of Natural History.Google Scholar
McIntosh, R. P. (1967). The continuum concept of vegetation. Botanical Review, 33, 130–87.CrossRefGoogle Scholar
McIntosh, R. P. (1985). The Background of Ecology: Concept and Theory. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
McIver, S. B. (2003). Death in the Everglades: The Murder of Guy Bradley, America's First Martyr to Environmentalism. Gainesville, FL: University of Florida Press.Google Scholar
McJannet, C. L., Keddy, P. A., and Pick, F. R. (1995). Nitrogen and phosphorus tissue concentrations in 41 wetland plants: a comparison across habitats and functional groups. Functional Ecology, 9, 231–8.CrossRefGoogle Scholar
McKee, K. L. and Mendelssohn, I. A. (1989). Response of a freshwater marsh plant community to increased salinity and increased water level. Aquatic Botany, 34, 301–16.CrossRefGoogle Scholar
McKenzie, D. H., Hyatt, D. E., and McDonald, V. J. (1992). Ecological Indicators, Vols. 1 and 2. London: Elsevier.Google Scholar
McMillan, M. (2006). Bog turtles make new friends: landowners and livestock. Environmental Defense Fund, Center for Conservation Incentives. www.edf.org. May 27, 2004, updated: Sep 13, 2006. (accessed July 17, 2008)
McNaughton, S. J., Russ, R. W., and Seagle, S. W. (1988). Large mammals and process dynamics in African ecosystems. BioScience, 38, 794–800.CrossRefGoogle Scholar
McPhee, J. (1989). The Control of Nature. New York: Farrar Straus Giroux.Google Scholar
McWilliams, R. G. (transl. and ed.) (1981). Iberville's Gulf Journals. Tuscaloosa, AL: University of Alabama Press.
Mead, K. (2003). Dragonflies of the North Woods. Duluth, MN: Kollath-Stensaas.Google Scholar
Meadows, D. H., Meadows, D. L., Randers, J., and Behrens, , , W. W. (1974). The Limits to Growth: A Report for the Club of Rome's Project on the Predicament of Mankind, 2nd edn. New York: New American Library.Google Scholar
Meave, J. and Kellman, M. (1994). Maintenance of rain forest diversity in riparian forests of tropical savannas: implications for species conservation during Pleistocene drought. Journal of Biogeography, 21, 121–35.CrossRefGoogle Scholar
Meave, J., Kellman, M., MacDougall, A., and Rosales, J. (1991). Riparian habitats as tropical refugia. Global Ecology and Biogeography Letters, 1, 69–76.CrossRefGoogle Scholar
Mendelssohn, I. A. and McKee, K. L. (1988). Spartina alterniflora die-back in Louisiana: time-course investigation of soil waterlogging effects. Journal of Ecology, 76, 509–21.CrossRefGoogle Scholar
Menges, E. S. and Gawler, S. C. (1986). Fourth-year changes in population size of the endemic Furbish's Lousewort: implications for endangerment and management. Natural Areas Journal, 6, 6–17.Google Scholar
Merritt, , R. W. and Cummins, K. W. (eds.) (1984). An Introduction to the Aquatic Insects of North America, 2nd edn. Dubuque, IA: Kendall/Hunt Publishing.
Messina, , M. G. and Conner, W. H. (eds.) (1998). Southern Forested Wetlands: Ecology and Management. Boca Raton, FL: Lewis Publishers.
Michener, W. K., Blood, E. R., Bildstein, K. L., Brinson, M. M., and Gardner, L. R. (1997). Climate change, hurricanes and tropical storms, and rising sea level in coastal wetlands. Ecological Applications, 7, 770–801.CrossRefGoogle Scholar
Middleton, , B. A. (ed.) (2002). Flood Pulsing in Wetlands: Restoring the Natural Hydrological Balance. New York: John Wiley.
,Millennium Ecosystem Assessment. (2005). Ecosystems and Human Well-Being: Wetlands and Water Synthesis. Washington, DC: World Resources Institute.
Miller, G. R. and Watson, A. (1978). Heather productivity and its relevance to the regulation of red grouse populations. In Production Ecology of British Moors and Montane Grasslands, eds. Heal, O. W. and Perkins, D. F., pp. 278–85. Berlin, Germany: Springer-Verlag.Google Scholar
Miller, G. R. and Watson, A. (1983). Heather moorland in northern Britain. In Conservation in Perspective, eds. Warren, A. and Goldsmith, F. B., pp. 101–17. Chichester, UK: John Wiley.Google Scholar
Miller, M. W. and Nudds, T. D. (1996). Prairie landscape change and flooding in the Mississippi River valley. Conservation Biology, 10, 847–53.CrossRefGoogle Scholar
Miller, R. M., Smith, C. I., Jastrow, J. D., and Bever, J. D. (2001). Mycorrhizal status of the genus Carex (Cyperaceae). American Journal of Botany, 86, 547–53.CrossRefGoogle Scholar
Miller, R. S. (1967). Pattern and process in competition. Advances in Ecological Research, 4, 1–74.CrossRefGoogle Scholar
Miller, R. S. (1968). Conditions of competition between redwings and yellowheaded blackbirds. Journal of Animal Ecology, 37, 43–62.CrossRefGoogle Scholar
Milliman, J. D. and Meade, R. H. (1983). World-wide delivery of river sediment to the oceans. Journal of Geology, 91, 1–21.CrossRefGoogle Scholar
Mitchell, G. F. (1965). Littleton Bog, Tipperary: an Irish vegetational record. Geological Society of America, Special Paper, 84, 1–16.CrossRefGoogle Scholar
Mitsch, W. J. and Gosselink, J. G. (1986). Wetlands. New York: Van Nostrand Reinhold.Google Scholar
Mitsch, W. J. and Wu, X. (1994). Wetlands and global change. In Advances in Soil Science: Global Carbon Sequestration, eds. Stewart, B. A., Lal, R., and Kimble, J. M., pp. 205–30. Chelsea, MI: Lewis Publishers.Google Scholar
Mitsch, W. J., Day, J. W., Gilliam, J. W., Groffman, P. M., Hey, D. L., Randall, G. W., and Wang, N. (2001). Reducing nitrogen loading to the Gulf of Mexico from the Mississippi River Basin: strategies to counter a persistent ecological problem. BioScience, 51, 373–88.CrossRefGoogle Scholar
Moeller, R. E. (1978). Carbon-uptake by the submerged hydrophyte Utricularia purpurea. Aquatic Botany, 5, 209–16.CrossRefGoogle Scholar
Monda, M. J., Ratti, J. T., and McCabe, T. R. (1994). Reproductive ecology of tundra swans on the arctic national wildlife refuge, Alaska. Journal of Wildlife Management, 58, 757–73.CrossRefGoogle Scholar
Montague, C. L. and Wiegert, R. G. (1990). Salt marshes. In Ecosystems of Florida, eds. Myers, R. L. and Ewel, J. J., pp. 481–516. Orlando, FL: University of Central Florida Press.Google Scholar
Montgomery, K. G. (1958). The Memoirs of Field-Marshal the Viscount Montgomery of Alamein. London: Collins.Google Scholar
Moore, D. R. J. (1998). The ecological component of ecological risk assessment: lessons from a field experiment. Human and Ecological Risk Assessment, 4, 1103–23.CrossRefGoogle Scholar
Moore, D. R. J. and Keddy, P. A. (1989). The relationship between species richness and standing crop in wetlands: the importance of scale. Vegetatio, 79, 99–106.CrossRefGoogle Scholar
Moore, D. R. J. and Wein, R. W. (1977). Viable seed populations by soil depth and potential site recolonization after disturbance. Canadian Journal of Botany, 55, 2408–12.CrossRefGoogle Scholar
Moore, D. R. J., Keddy, P. A., Gaudet, C. L., and Wisheu, I. C. (1989). Conservation of wetlands: do infertile wetlands deserve a higher priority?Biological Conservation, 47, 203–17.CrossRefGoogle Scholar
Moore, P. D. (1973). The influence of prehistoric cultures upon the initiation and spread of blanket bog in upland Wales. Nature, 241, 350–3.CrossRefGoogle Scholar
Moorhead, K. K. and Reddy, K. R. (1988). Oxygen transport through selected aquatic macrophytes. Journal of Environmental Quality, 17, 138–42.CrossRefGoogle Scholar
Morgan, M. D. and Philipp, K. R. (1986). The effect of agricultural and residential development on aquatic macrophytes in the New Jersey Pine Barrens. Biological Conservation, 35, 143–58.CrossRefGoogle Scholar
Morowitz, H. J. (1968). Energy Flow in Biology. New York: Academic Press.Google Scholar
Morris, J. (1973). Pax Britannica, 3 Vols. London: Faber and Faber. Reprinted 1992 by Folio Society, London.Google Scholar
Mosepele, K., Moyle, P. B., Merron, G. S., Purkey, D. R., and Mosepele, B. (2009). Fish, floods and ecosystem engineers: aquatic conservation in the Okavango Delta, Botswana. BioScience, 59, 53–64.CrossRefGoogle Scholar
Moss, B. (1983). The Norfolk Broadland: experiments in the restoration of a complex wetland. Biological Reviews of the Cambridge Philosophical Society, 58, 521–61.CrossRefGoogle Scholar
Moss, B. (1984). Medieval man-made lakes: progeny and casualties of English social history, patients of twentieth century ecology. Transactions of the Royal Society of South Africa, 45, 115–28.CrossRefGoogle Scholar
Mountford, J. O., Lakhani, K. H., and Kirkham, F. W. (1993). Experimental assessment of the effects of nitrogen addition under hay-cutting and aftermath grazing on the vegetation of meadows on a Somerset peat moor. Journal of Applied Ecology, 30, 321–32.CrossRefGoogle Scholar
Mueller-Dombois, D. and Ellenberg, H. (1974). Aims and Methods of Vegetation Ecology. New York:John Wiley.Google Scholar
Müller, J., Rosenthal, G., and Uchtmann, H. (1992). Vegetationsveränderungen und Ökologie nordwestdeutscher Feuchtgrünlandbrachen. Tuexenia, 12, 223–44.Google Scholar
Müller, J., Irion, G., Mello, J. N., and Junk, W. J. (1995). Hydrological changes of the Amazon during the last glacial–interglacial cycle in Central Amazonia (Brazil). Naturwissenschaften, 82, 232–5.CrossRefGoogle Scholar
Murkin, H. R. (1989). The basis for food chains in prairie wetlands. In Northern Prairie Wetlands, ed. Valk, A. G., pp. 316–38. Ames, IA: Iowa State University Press.Google Scholar
Mushet, D. M., Euliss, N. H., and Shaffer, T. L. (2002). Floristic quality assessment of one natural and three restored wetland complexes in North Dakota, USA. Wetlands, 22, 126–38.CrossRefGoogle Scholar
Myers, J. G. (1935). Zonation of vegetation along river courses. Journal of Ecology, 3, 356–60.CrossRefGoogle Scholar
Myers, N., Mittermeier, R. A., Mittermeier, C. G., da Fonseca, G. A. B., and Kent, J. (2000). Biodiversity hotspots for conservation priorities. Nature, 403, 853–8.CrossRefGoogle ScholarPubMed
Myers, R. K. and Lear, D. H. (1998). Hurricane–fire interactions in coastal forests of the south: a review and hypothesis. Forest Ecology and Management, 103, 265–76.CrossRefGoogle Scholar
Myers, R. L. (1983). Site susceptibility to invasion by the exotic tree Melaleuca quinquenervia in southern Florida. Journal of Applied Ecology, 20, 645–58.CrossRefGoogle Scholar
Myers, R. S., Shaffer, G. P., and Llewellyn, D. W. (1995). Baldcypress (Taxodium distichum (L.) Rich.) restoration in southeastern Louisiana: the relative effects of herbivory, flooding, competition and macronutrients. Wetlands, 15, 141–8.CrossRefGoogle Scholar
Naiman, R. J., Johnston, C. A., and Kelley, J. C. (1988). Alteration of North American streams by beaver. BioScience, 38, 753–62.CrossRefGoogle Scholar
Nanson, G. C. and Beach, H. F. (1977). Forest succession and sedimentation on a meandering-river floodplain, northeast British Columbia, Canada. Journal of Biogeography, 4, 229–51.CrossRefGoogle Scholar
Navid, D. (1988). Developments under the Ramsar Convention. In The Ecology and Management of Wetlands, Vol. 2, Management, Use and Value of Wetlands, eds. Hook, D. D., McKee, Jr. W. H., Smith, H. K., Gregory, J., Burrell, Jr. V. G., DeVoe, M. R., Sojka, R. E., Gilbert, S., Banks, R., Stolzy, L. H., Brooks, C., Matthews, T. D., and Shear, T. H., pp. 21–7. Portland, OR: Timber Press.CrossRefGoogle Scholar
Neiff, J. J. (1986). Aquatic plants of the Parana system. In The Ecology of River Systems, eds. Davies, B. R. and Walker, K. F. pp. 557–71. Dordrecht, the Netherlands: Dr. W. Junk Publishers.Google Scholar
Neill, W. T. (1950). An estivating bowfin. Copeia, 3, 240.Google Scholar
Newman, S., Grace, J. B., and Koebel, J. W. (1996). The effects of nutrients and hydroperiod on mixtures of Typha domingensis, Cladium jamaicense, and Eleocharis interstincta: implications for Everglades restoration. Ecological Applications, 6, 774–83.CrossRefGoogle Scholar
Newman, S., Schuette, J., Grace, J. B., Rutchey, K., Fontaine, T., Reddy, K. R., and Pietrucha, M. (1998). Factors influencing cattail abundance in the northern Everglades. Aquatic Botany, 60, 265–80.CrossRefGoogle Scholar
,New York Natural Heritage Program. (2008). Online Conservation Guide for Glyptemys muhlenbergii. www.acris.nynhp.org/guide.php?id=7507. (accessed July 27, 2008)
Nicholls, R. J. and Mimura, N. (1998). Regional issues raised by sea-level rise and their policy implications. Climate Research, 11, 5–18.CrossRefGoogle Scholar
Nichols, S. A. (1999). Floristic quality assessment of Wisconsin lake plant communities with example applications. Journal of Lake and Reservoir Management, 15, 133–41.CrossRefGoogle Scholar
Niering, W. A. and Warren, R. S. (1980). Vegetation patterns and processes in New England salt marshes. BioScience, 30, 301–7.CrossRefGoogle Scholar
Nilsson, C. (1981). Dynamics of the shore vegetation of a north Swedish hydroelectric reservoir during a 5-year period. Acta Phytogeographica Suecica, 69, 1–96.Google Scholar
Nilsson, C. and Jansson, R. (1995). Floristic differences between riparian corridors of regulated and free-flowing boreal rivers. Regulated Rivers: Research and Management, 11, 55–66.CrossRefGoogle Scholar
Nilsson, C. and Keddy, P. A. (1988). Predictability of change in shoreline vegetation in a hydroelectric reservoir, northern Sweden. Canadian Journal of Fisheries and Aquatic Sciences, 45, 1896–904.CrossRefGoogle Scholar
Nilsson, C. and Wilson, S. D. (1991). Convergence in plant community structure along disparate gradients: are lakeshores inverted mountainsides?The American Naturalist, 137, 774–90.CrossRefGoogle Scholar
Nilsson, C., Grelsson, G., Johansson, M., and Sperens, U. (1989). Patterns of plant species richness along riverbanks. Ecology, 70, 77–84.CrossRefGoogle Scholar
Nilsson, C., Grelsson, G., Dynesius, M., Johansson, M. E., and Sperens, U. (1991). Small rivers behave like large rivers: effects of postglacial history on plant species richness along riverbanks. Journal of Biogeography, 18, 533–41.CrossRefGoogle Scholar
Norgress, R. E. (1947). The history of the cypress lumber industry in Louisiana. Louisiana Historical Quarterly, 30, 979–1059.Google Scholar
Noss, R. (1995). Maintaining Ecological Integrity in Representative Reserve Networks, A World Wildlife Fund Canada/World Wildlife Fund United States Discussion Paper. Washington, DC: WWF.Google Scholar
Noss, R. F. and Cooperrider, A. (1994). Saving Nature's Legacy: Protecting and Restoring Biodiversity. Washington, DC: Defenders of Wildlife and Island Press.Google Scholar
Novacek, J. M. (1989). The water and the wetland resources of the Nebraska sandhills. In Northern Prairie Wetlands, ed. Valk, A. G., pp. 340–84. Ames, IA: Iowa State University Press.Google Scholar
Noy-Meir, I. (1975). Stability of grazing systems: an application of predator–prey graphs. Journal of Ecology, 63, 459–81.CrossRefGoogle Scholar
Nudds, T. D., Sjoberg, K., and Lundberg, P. (1994). Ecomorphological relationships among Palearctic dabbling ducks on Baltic coastal wetlands and a comparison with the Nearctic. Oikos, 69, 295–303.CrossRefGoogle Scholar
Nuttle, W. K., Brinson, M. M., Cahoon, D., Callaway, J. C., Christian, R. R., Chmura, G. L., Conner, W. H., Day, R. H., Ford, M., Grace, J., Lynch, J. C., Orson, R. A., Parkinson, R. W., Reed, D., Rybczyk, J. M., Smith, T. J., III, Stumpf, R. P., and Williams, K. (1997). The Working Group on Sea Level Rise and Wetland Systems: conserving coastal wetlands despite sea level rise. Eos, 78, 257–62.Google Scholar
Odum, E. P. (1971). Principles of Ecology. Philadelphia, PA: W. B. Saunders.Google Scholar
Odum, E. P. (1985). Trends expected in stressed ecosystems. BioScience, 35, 419–22.CrossRefGoogle Scholar
Odum, W. E. and McIvor, C. C. (1990). Mangroves. In Ecosystems of Florida, eds. Myers, R. L. and Ewel, J. J., pp. 517–48. Orlando, FL: University of Central Florida Press.Google Scholar
Oksanen, L. (1990). Predation, herbivory, and plant strategies along gradients of primary production. In Perspectives on Plant Competition, eds. Grace, J. B. and Tilman, D., pp. 445–74. New York: Academic Press.Google Scholar
Oksanen, L., Fretwell, S. D., Arruda, J., and Niemela, P. (1981). Exploitation ecosystems in gradients of primary productivity. The American Naturalist, 118, 240–261.CrossRefGoogle Scholar
O'Neil, T. (1949). The Muskrat in the Louisiana Coastal Marshes. New Orleans, LA: Louisiana Department of Wildlife and Fisheries.Google Scholar
,Ontario Ministry of Natural Resources. (1993). Ontario Wetland Evaluation System: Southern Manual, 3rd edn, revised 2002. Toronto, ON: Ontario Ministry of Natural Resources.Google Scholar
,Ontario Ministry of Natural Resources. (2007). Significant Wetlands and the Ontario Wetland Evaluation System. Peterborough, ON: Ontario Ministry of Natural Resources.Google Scholar
Oomes, M. J. M and Elberse, W. T. (1976). Germination of six grassland herbs in microsites with different water contents. Journal of Ecology, 64, 745–55.CrossRefGoogle Scholar
Orson, R. A., Simpson, R. L., and Good, R. E. (1990). Rates of sediment accumulation in a tidal freshwater marsh. Journal of Sedimentary Petrology, 60, 859–69.Google Scholar
Orson, R. A., Simpson, R. L., and Good, R. E. (1992). The paleoecological development of a late Holocene, tidal freshwater marsh of the Upper Delaware River estuary. Estuaries, 15, 130–46.CrossRefGoogle Scholar
Osborne, P. L. and Polunin, N. V. C. (1986). From swamp to lake: recent changes in a lowland tropical swamp. Journal of Ecology, 74, 197–210.CrossRefGoogle Scholar
Ostrofsky, M. L. and Zettler, E. R. (1986). Chemical defenses in aquatic plants. Journal of Ecology, 74, 279–87.CrossRefGoogle Scholar
Padgett, D. J. and Crow, G. E. (1993). A comparison of floristic composition and species richness within and between created and natural wetlands of southeastern New Hampshire. In Proceedings of the 20th Annual Conference on Wetlands Restoration and Creation, ed. Webb, F. J., Jr., pp. 171–86. Tampa, FL: Hillsborough Community College.Google Scholar
Padgett, D. J. and Crow, G. E. (1994). Foreign plant stock: concerns for wetland mitigation. Restoration and Management Notes, 12, 168–71.Google Scholar
Painter, S. and Keddy, P. A. (1992). Effects of Water Level Regulation on Shoreline Marshes: A Predictive Model Applied to the Great Lakes. Burlington, ON: Environment Canada, National Water Research Institute.Google Scholar
Painter, T. J. (1991). Lindow Man, Tollund Man, and other peat-bog bodies: the preservative and antimicrobial action of sphagnan, a reactive glycuronoglycan with tanning and sequestering properties. Carbohydrate Polymers, 15, 123–42.CrossRefGoogle Scholar
Palczynski, A. (1984). Natural differentiation of plant communities in relation to hydrological conditions of the Biebrza valley. Polish Ecological Studies, 10, 347–85.Google Scholar
Parmalee, P. W. and Graham, R. W. (2002). Additional records of the Giant Beaver, Castoroides, from the mid-South: Alabama, Tennessee, and South Carolina. Smithsonian Contributions to Paleobiology, 93, 65–71.Google Scholar
Partow, H. (2001). The Mesopotamian Marshlands: Demise of an Ecosystem, Early Warning and Assessment Technical Report. Nairobi, Kenya: United Nations Environment Programme.Google Scholar
Partridge, T. R. and Wilson, J. B. (1987). Salt tolerance of salt marsh plants of Otago, New Zealand. New Zealand Journal of Botany, 25, 559–66.CrossRefGoogle Scholar
Patrick, W. H. and Reddy, C. N. (1978). Chemical changes in rice soils. In Soils and Rice, pp. 361–79. Los Baños, Philippines: International Rice Research Institute.Google Scholar
Patten, , B. C. (ed.) (1990). Wetlands and Shallow Continental Water Bodies, Vol. 1, Natural and Human Relationships. The Hague, the Netherlands: SPB Academic Publishing.
Patten, D. T. (1998). Riparian ecosystems of semi-arid North America: diversity and human impacts. Wetlands, 18, 498–512.CrossRefGoogle Scholar
,Peace–Athabasca Delta Implementation Committee. (1987). Peace–Athabasca Delta Water Management Works Evaluation: Final Report. Ottawa, ON: Environment Canada, Alberta Environment and Saskatchewan Water Corporation.Google Scholar
,Peace–Athabasca Delta Project Group. (1972). The Peace–Athabasca Delta Summary Report, 1972. Ottawa, ON: Department of the Environment.Google Scholar
Peach, M. and Zedler, J. B. (2006). How tussocks structure sedge meadow vegetation. Wetlands, 26, 322–35.CrossRefGoogle Scholar
Pearce, F. (1991). The rivers that won't be tamed. New Scientist, 1764, 38–41.Google Scholar
Pearce, F. (1993). Draining life from Iraq's marshes. New Scientist, 1869,11–12.Google Scholar
Pearman, P. B. (1997). Correlates of amphibian diversity in an altered landscape of Amazonian Ecuador. Conservation Biology, 11,1211–25.CrossRefGoogle Scholar
Pearsall, W. H. (1920). The aquatic vegetation of the English Lakes. Journal of Ecology, 8, 163–201.CrossRefGoogle Scholar
Pearse, P. H., Bertrand, F. X., and MacLaren, J. W. (1985). Currents of Change, Final Report. Ottawa, ON: Inquiry on Federal Water Policy.Google Scholar
Peat, H. J. and Fitter, A. H. (1993). The distribution of arbuscular mycorrhizae in the British flora. New Phytologist, 125, 845–54.CrossRefGoogle Scholar
Pechmann, J. H. K., Scott, D. E., Gibbons, J. W., and Semlitsch, R. D. (1989). Influence of wetland hydroperiod on diversity and abundance of metamorphosing juvenile amphibians. Wetlands Ecology and Management, 1, 3–11.CrossRefGoogle Scholar
Pedersen, O., Sand-Jensen, K., and Revsbech, N. P. (1995). Diel pulses of O2 and CO2 in sandy lake sediments inhabited by Lobelia dortmanna. Ecology, 76, 1536–45.CrossRefGoogle Scholar
Peet, R. K. (1974). The measurement of species diversity. Annual Review of Ecology and Systematics, 5, 285–307.CrossRefGoogle Scholar
Peet, R. K. and Allard, D. J. (1993). Longleaf pine vegetation of the southern Atlantic and eastern Gulf Coast regions: a preliminary classification. In The Longleaf Pine Ecosystem: Ecology, Restoration and Management, ed. Hermann, S. M., pp. 45–81. Tallahassee, FL: Tall Timbers Research Station.Google Scholar
Pehek, E. L. (1995). Competition, pH, and the ecology of larval Hyla andersonii. Ecology, 76, 1786–93.CrossRefGoogle Scholar
Pemberton, R. W., Goolsby, J. A., and Wright, T. (2002). Old world climbing fern. In Biological Control of Invasive Plants in the Eastern United States, Publication No. FHTET-2002–04, eds. Driesche, R., Lyon, S., Blossey, B., Hoddle, M., and Reardon, R., pp. 139–47. Morgantown, WV: U.S. Department of Agriculture Forest Service.Google Scholar
Pengelly, J. W., Tinkler, K. J., Parkins, W. G., and McCarthy, F. M. (1997). 12,600 years of lake level changes, changing sills, ephemeral lakes and Niagara gorge erosion in the Niagara Peninsula and Eastern Lake Erie basin. Journal of Paleolimnology, 17, 377–402.CrossRefGoogle Scholar
Penland, S., Boyd, R., and Suter, J. R. (1988). The transgressive depositional systems of the Mississippi delta plain: a model for barrier shoreline and shelf sand development. Journal of Sedimentary Petrology, 58, 932–49.Google Scholar
Pennings, S. C. and Callaway, R. M. (1992). Salt marsh zonation: the relative importance of competition and physical factors. Ecology, 73, 681–90.CrossRefGoogle Scholar
Pennings, S. C., Carefoot, T. H., Siska, E. L., Chase, M. E., and Page, T. A. (1998). Feeding preferences of a generalist salt-marsh crab: relative importance of multiple plant traits. Ecology, 79, 1968–79.CrossRefGoogle Scholar
Perkins, D. F. (1978). Snowdonia grassland: introduction, vegetation and climate. In Production Ecology of British Moors and Montane Grasslands, eds. Heal, O. W. and Perkins, D. F., pp. 290–6. Berlin, Germany: Springer-Verlag.Google Scholar
Peters, R. H. (1980a). From natural history to ecology. Perspectives in Biology and Medicine, 23, 191–203.CrossRefGoogle Scholar
Peters, R. H. (1980b). Useful concepts for predictive ecology. In Conceptual Issues in Ecology, ed. Saarinen, E., pp. 63–99. Dordrecht, the Netherlands: D. Reidel.Google Scholar
Peterson, L. P., Murkin, H. R., and Wrubleski, D. A. (1989). Waterfowl predation on benthic macroinvertebrates during fall drawdown of a northern prairie marsh. In Freshwater Wetlands and Wildlife, eds. Sharitz, R. R. and Gibbons, J. W., pp. 661–96. Washington, DC: U.S. Department of Energy.Google Scholar
Petr, T. (1986). The Volta River system. In The Ecology of River Systems, eds. Davies, B. R. and Walker, K. F., pp. 163–83. Dordrecht, the Netherlands: Dr. W. Junk Publishers.CrossRefGoogle Scholar
Pezeshki, S. R., Delaune, R. D., and Patrick, W. H. (1987a). Effects of flooding and salinity on photosynthesis of Sagittaria Lancifolia. Marine Ecology Progress Series, 41, 87–91.CrossRefGoogle Scholar
Pezeshki, S. R., Delaune, R. D., and Patrick, W. H. (1987b). Response of the freshwater marsh species Panicum hemitomon Schult. to increased salinity. Freshwater Biology, 1, 195–200.CrossRefGoogle Scholar
Pfadenhauer, J. and Klotzli, F. (1996). Restoration experiments in middle European wet terrestrial ecosystems: an overview. Vegetatio, 126, 101–15.Google Scholar
Phillips, G. L., Eminson, D., and Moss, B. (1978). A mechanism to account for macrophyte decline in progressively eutrophicated fresh-waters. Aquatic Botany, 4, 103–26.CrossRefGoogle Scholar
Phipps, R. W. (1883). On the Necessity of Preserving and Replanting Forests. Toronto, ON: Blackett and Robinson.CrossRefGoogle Scholar
Pianka, E. R. (1981). Competition and niche theory. In Theoretical Ecology, ed. May, R. M., pp. 114–41. Oxford, UK: Blackwell Scientific Publications.Google Scholar
Pickett, S. T. A. (1980). Non-equilibrium coexistence of plants. Bulletin of the Torrey Botanical Club, 107, 238–48.CrossRefGoogle Scholar
Pickett, S. T. A. and White, P. S. (1985). The Ecology of Natural Disturbance and Patch Dynamics. Orlando, FL: Academic Press.Google Scholar
Picman, J. (1984). Experimental study on the role of intra- and inter-specific behaviour in marsh wrens. Canadian Journal of Zoology, 62, 2353–6.CrossRefGoogle Scholar
Pieczynska, E. (1986). Littoral communities and lake eutrophication. In Land Use Impacts on Aquatic Ecosystems, eds. Lauga, J., Decamps, H., and Holland, M. M., Proceedings of the Toulouse Workshop organized by MAB-UNESCO and PIREN-CNRS, pp. 191–201 Paris: UNESCO.Google Scholar
Pielou, E. C. (1975). Ecological Diversity. New York: John Wiley.Google Scholar
Pielou, E. C. (1977). Mathematical Ecology. New York: John Wiley.Google Scholar
Pielou, E. C. and Routledge, R. D. (1976). Salt marsh vegetation: latitudinal gradients in the zonation patterns. Oecologia, 24, 311–21.CrossRefGoogle ScholarPubMed
Pietropaolo, J. and Pietropaolo, P. (1986). Carnivorous Plants of the World. Portland, OR: Timber Press.Google Scholar
Pimental, D., Hurd, L. E., Bellotti, A. C., Forster, M. J., Oka, I., Sholes, O. D., and Whitman, W. J. (1973). Food production and the energy crisis. Science, 182, 443–9.CrossRefGoogle Scholar
Poiana, K. A. and Johnson, W. C. (1993). A spatial simulation model of hydrology and vegetation dynamics in semi-permanent prairie wetlands. Ecological Applications, 3, 279–93.CrossRefGoogle Scholar
Polunin, N. V. C. (1984). The decomposition of emergent macrophytes in fresh water. Advances in Ecological Research, 14, 115–66.CrossRefGoogle Scholar
Pomeroy, , L. R. and Wiegert, R. J. (eds.) (1981). The Ecology of a Salt Marsh. Berlin, Germany: Springer-Verlag.CrossRef
Ponnamperuma, F. N. (1972). The chemistry of submerged soils. Advances in Agronomy, 24, 29–96.CrossRefGoogle Scholar
Ponnamperuma, F. N. (1984). Effects of flooding on soils. In Flooding and Plant Growth, ed. Kozlowski, T. T., pp. 9–45. Orlando, FL: Academic Press.CrossRefGoogle Scholar
Poole, R. W. and Rathcke, B. J. (1979). Regularity, randomness, and aggregation in flowering phenologies. Science, 203, 470–1.CrossRefGoogle ScholarPubMed
Power, M. E. (1992). Top-down and bottom-up forces in food webs: do plants have primacy?Ecology, 73, 733–46.CrossRefGoogle Scholar
Prance, G. T. and Schaller, J. B. (1982). Preliminary study of some vegetation types of the Pantanal, Mato Grosso, Brazil. Brittonia, 34, 228–51.CrossRefGoogle Scholar
Pressey, R. L., Humphries, C. J., Margules, C. R., Vane-Wright, R. I., and Williams, P. H. (1993). Beyond opportunism: key principles for systematic reserve selection. Trends in Ecology and Evolution, 8, 124–8.CrossRefGoogle ScholarPubMed
Preston, F. W. (1962a). The canonical distribution of commonness and rarity: Part I. Ecology, 43, 185–215.CrossRefGoogle Scholar
Preston, F. W. (1962b). The canonical distribution of commonness and rarity: Part II. Ecology, 43, 410–32.CrossRefGoogle Scholar
Price, M. V. (1980). On the significance of test form in benthic salt-marsh foraminifera. Journal of Foraminiferal Research, 10, 129–35.CrossRefGoogle Scholar
Prince, , H. H. and D'Itri, F. M. (eds.) (1985). Coastal Wetlands. Chelsea, MI: Lewis Publishers.
Prince, H. H. and Flegel, C. S. (1995). Breeding avifauna of Lake Huron. In The Lake Huron Ecosytem: Ecology, Fisheries and Management, eds. Munawar, M., Edsall, T., and Leach, J., pp. 247–72. Amsterdam, the Netherlands: SPB Academic Publishing.Google Scholar
Prince, H. H., Padding, P. I., and Knapton, R. W. (1992). Waterfowl use of the Laurentian Great Lakes. Journal of Great Lakes Research, 18, 673–99.CrossRefGoogle Scholar
Prowse, T. D. and Culp, J. M. (2003). Ice breakup: a neglected factor in river ecology. Canadian Journal of Civil Engineering, 30, 128–44.CrossRefGoogle Scholar
Radford, A. E., Ahles, H. E., and Bell, C. R. (1968). Manual of the Vascular Flora of the Carolinas. Chapel Hill, NC:University of North Carolina Press.Google Scholar
Rapport, D. J. (1989). What constitutes ecosystem health?Perspectives in Biology and Medicine, 33, 120–32.CrossRefGoogle Scholar
Rapport, D. J., Thorpe, C., and Hutchinson, T. C. (1985). Ecosystem behaviour under stress. The American Naturalist, 125, 617–40.CrossRefGoogle Scholar
Rasker, R. and Hackman, A. (1996). Economic development and the conservation of large carnivores. Conservation Biology, 10, 991–1002.CrossRefGoogle Scholar
Raunkiaer, C. (1908). The statistics of life forms as a basis for biological plant geography. In The Life Forms of Plants and Statistical Plant Geography: Being the Collected Papers of Raunkiaer, pp. 111–47. Oxford, UK: Clarendon Press.Google Scholar
Raunkiaer, C. (1937). Plant Life Forms, translated by Gilbert-Cater, H.. Oxford, UK: Clarendon Press.Google Scholar
Raup, H. M. (1975). Species versatility in shore habitats. Journal of the Arnold Arboretum, 56, 126–63.Google Scholar
Raven, P. H., Evert, R. F., and Eichhorn, S. E. (1992). Biology of Plants, 5th edn. New York: Worth Publishers.Google Scholar
Ravera, O. (1989). Lake ecosystem degradation and recovery studied by the enclosure method. In Ecological Assessment of Environmental Degradation, Pollution and Recovery, ed. Ravera, O.. Amsterdam, the Netherlands: Elsevier.Google Scholar
Rawes, M. and Heal, O. W. (1978). The blanket bog as part of a Pennine moorland. In Production Ecology of British Moors and Montane Grasslands, eds. Heal, O. W. and Perkins, D. F., pp. 224–43. Berlin, Germany: Springer-Verlag.CrossRefGoogle Scholar
Rayamajhi, M. B., Purcell, M. F., Van, T. K., Center, T. D., Pratt, P. D., and Buckingham, G. R. (2002). Australian paperbark tree (Melaleuca). In Biological Control of Invasive Plants in the Eastern United States, Publication No. FHTET-2002–04, eds. Driesche, R., Lyon, S., Blossey, B., Hoddle, M., and Reardon, R., pp. 117–30. Morgantown, WV: U.S. Department of Agriculture Forest Service.Google Scholar
Read, D. J., Koucheki, H. K., and Hodgson, J. (1976). Vesicular–arbuscular mycorrhizae in natural vegetation systems. I. The occurrence of infection. New Phytologist, 77, 641–53.CrossRefGoogle Scholar
Read, D. J., Francis, R., and Finlay, R. D. (1985). Mycorrhizal mycelia and nutrient cycling in plant communities. In Ecological Interactions in Soil, ed. Fitter, A. H., pp. 193–217. Oxford: Blackwell Scientific Publications.Google Scholar
Reddoch, J. and Reddoch, A. (1997). The orchids in the Ottawa district. Canadian Field-Naturalist, 111, 1–185.Google Scholar
Reddy, K. R. and Patrick, W. H. (1984). Nitrogen transformations and loss in flooded soils and sediments. CRC Critical Reviews in Environmental Control, 13, 273–309.CrossRefGoogle Scholar
Reid, D. M. and Bradford, K. J. (1984). Effect of flooding on hormone relations. In Flooding and Plant Growth, ed. Kozlowski, T. W., pp. 195–219. Orlando, FL: Academic Press.CrossRefGoogle Scholar
Reid, W. V., McNeely, J. A., Tunstall, J. B., Bryant, D. A., and Winograd, M. (1993). Biodiversity Indicators for Policymakers. Washington, DC: World Resources Institute.Google Scholar
Rejmankova, E., Pope, K. O., Pohl, M. D., and Rey-Benayas, J. M. (1995). Freshwater wetland plant communities of northern Belize: implications for paleoecological studies of Maya wetland agriculture. Biotropica, 27, 28–36.CrossRefGoogle Scholar
Reuss, M. (1998). Designing the Bayous: The Control of Water in the Atchafalaya Basin 1800–1995. Alexandria, VA: U.S. Army Corps of Engineers Office of History.Google Scholar
Reynoldson, T. B. and Zarull, M. A. (1993). An approach to the development of biological sediment guidelines. In Ecological Integrity and the Management of Ecosystems, eds. Woodley, S., Kay, J., and Francis, G., pp. 177–200. Delray Beach, FL: St. Lucie Press.Google Scholar
Reznicek, A. A. and Catling, P. M. (1989). Flora of Long Point. Michigan Botanist, 28, 99–175.Google Scholar
Richardson, , C. J. (ed.) (1981). Pocosin Wetlands: An Integrated Analysis of Coastal Plain Freshwater Bogs in North Carolina. Stroudsburg, PA: Hutchinson Ross.
Richardson, C. J. (1985). Mechanisms controlling phosphorus retention capacity in freshwater wetlands. Science, 228, 1424–7.CrossRefGoogle ScholarPubMed
Richardson, C. J. (1989). Freshwater wetlands: transformers, filters, or sinks? In Freshwater Wetlands and Wildlife, eds. Sharitz, R. R. and Gibbons, J. W., pp. 25–46. Proceedings of a symposium held at Charleston, South Carolina, Mar 24–27, 1986. Washington, DC: U.S. Department of Energy.Google Scholar
Richardson, C. J. (1991). Pocosins: an ecological perspective. Wetlands, 11, 335–54.CrossRefGoogle Scholar
Richardson, C. J. (1995). Wetlands ecology. In Encyclopedia of Environmental Biology, Vol. 3, ed. W. A. Nierenberg, pp. 535–50. San Diego, CA: Academic Press.Google Scholar
Richardson, C. J. and Gibbons, J. W. (1993). Pocosins, Carolina bays and mountain bogs. In Biodiversity of the Southeastern United States, eds. Martin, W. H., Boyce, S. G., and Echternacht, A. C., pp. 257–310. New York: John Wiley.Google Scholar
Richey, J. E., Meade, R. H., Salati, E., Devol, A. H., Nordin, C. F., and dos Santos, U. (1986). Water discharge and suspended sediment concentrations in the Amazon River: 1982–1984. Water Resources Research, 23, 756–64.CrossRefGoogle Scholar
Richter, B. D., Braun, D. P., Mendelson, M. A., and Master, L. L. (1997). Threats to imperiled freshwater fauna. Conservation Biology, 11, 1081–93.CrossRefGoogle Scholar
Richter, K. O. and Azous, A. L. (1995). Amphibian occurrence and wetland characteristics in the Puget Sound Basin. Wetlands, 15, 305–12.CrossRefGoogle Scholar
Richter, S. C. and Seigel, R. A. (2002). Annual variation in the population ecology of the endangered gopher frog, Rana sevosa Goin and Netting. Copeia, 2002, 962–72.CrossRefGoogle Scholar
Richter, S. C, Young, J. E., Seigel, R. A., and Johnson, G. N. (2001). Postbreeding movements of the dark gopher frog, Rana sevosa Goin and Netting: implications for conservation and management. Journal of Herpetology, 35, 316–21.CrossRefGoogle Scholar
Richter, S. C., Young, J. E., Johnson, G. N., and Seigel, R. A. (2003). Stochastic variation in reproductive success of a rare frog, Rana sevosa: implications for conservation and for monitoring amphibian populations. Biological Conservation, 111, 171–7.CrossRefGoogle Scholar
Rickerl, D. H., Sancho, F. O., and Ananth, S. (1994). Vesicular–arbuscular endomycorrhizal colonization of wetland plants. Journal of Environmental Quality, 23, 913–16.CrossRefGoogle Scholar
Ricklefs, R. E. (1987). Community diversity: relative roles of local and regional processes. Science, 235, 167–71.CrossRefGoogle ScholarPubMed
Rigler, F. H. (1982). Recognition of the possible: an advantage of empiricism in ecology. Canadian Journal of Fisheries and Aquatic Sciences, 39, 1323–31.CrossRefGoogle Scholar
Rigler, F. H. and Peters, R. H. (1995). Science and Limnology. Oldendorf/Lutie, Germany: Ecology Institute.Google Scholar
Riley, J. L. (1982). Hudson Bay lowland floristic inventory, wetlands catalogue and conservation strategy. Naturaliste Canadien, 109, 543–55.Google Scholar
Riley, J. L. (1989). Southern Ontario bogs and fens off the Canadian Shield. In Wetlands: Inertia or Momentum? Conference Proceedings, Oct 21–22, pp. 355–67. Toronto, ON: Federation of Ontario Naturalists.Google Scholar
Riley, T. Z. and Bookhout, T. A. (1990). Responses of aquatic macroinvertebrates to early-spring drawdown in nodding smartweed marshes. Wetlands, 10, 173–85.CrossRefGoogle Scholar
Ritchie, J. C. (1987). Postglacial Vegetation of Canada. New York: Cambridge University Press.Google Scholar
Roberts, J. and Ludwig, J. A. (1991). Riparian vegetation along current-exposure gradients in floodplain wetlands of the River Murray, Australia. Journal of Ecology, 79, 117–27.CrossRefGoogle Scholar
Robertson, P. A., Weaver, G. T., and Cavanaugh, J. A. (1978). Vegetation and tree species patterns near the northern terminus of the southern floodplain forest. Ecological Monographs, 48, 249–67.CrossRefGoogle Scholar
Robertson, R. J. (1972). Optimal niche space of the redwinged blackbird (Agelaius phoeniceus). I. Nesting success in marsh and upland habitat. Canadian Journal of Zoology, 50, 247–63.CrossRefGoogle Scholar
Robins, R. H. (n.d.). Walking catfish. www.flmnh.ufl.edu/fish/Gallery/Descript/WalkingCatfish/WalkingCatfish.html (accessed June 1, 2008)
Robinson, A. R. (1973). Sediment, our greatest pollutant? In Focus on Environmental Geology, ed. Tank, R. W., pp. 186–92. London: Oxford University Press.Google Scholar
Rogers, D. R., Rogers, B. D., and Herke, W. H. (1992). Effects of a marsh management plan on fishery communities in coastal Louisiana. Wetlands, 12, 53–62.CrossRefGoogle Scholar
Rolston, H. (1994). Foreword. In An Environmental Proposal for Ethics: The Principle of Integrity, ed. Westra, L., pp. xi–xiii. Lanham, MD: Rowman and Littlefield. In Noss, R. (1995). Maintaining Ecological Integrity in Representative Reserve Networks, A World Wildlife Fund Canada/World Wildlife Fund United States Discussion Paper. Washington, DC: WWF.Google Scholar
Roni, P., Hanson, K., Beechie, T., Pess, G., Pollock, M., and Bartley, D. M. (2005). Habitat Rehabilitation for Inland Fisheries: Global Review of Effectiveness and Guidance for Rehabilitation of Freshwater Ecosystems, FAO Fisheries Technical Paper No. 484. Rome, Italy: Food and Agriculture Organization.Google Scholar
Rood, S. B. and Mahoney, J. M. (1990). Collapse of riparian poplar forests downstream from dams in western prairies: probable causes and prospects for mitigation. Environmental Management, 14, 451–64.CrossRefGoogle Scholar
Root, R. (1967). The niche exploitation pattern of the blue-grey gnatcatcher. Ecological Monographs, 37, 317–50.CrossRefGoogle Scholar
Rørslett, B. (1984). Environmental factors and aquatic macrophyte response in regulated lakes: a statistical approach. Aquatic Botany, 19, 199–220.CrossRefGoogle Scholar
Rørslett, B. (1985). Regulation impact on submerged macrophytes in the oligotrophic lakes of Setesdal, South Norway. International Association for Theoretical and Applied Limnology, 22, 2927–36.Google Scholar
Rosen, B. H., Gray, S., and Flaig, E. (1995). Implementation of Lake Okeechobee watershed management strategies to control phosphorus load. In Wetlands and Watershed Management: Science Applications and Public Policy, eds. Kusler, J. A., Willard, D. E., and Hull, Jr. H. C., pp. 199–207. A collection of papers from a national symposium and several workshops at Tampa, FL, Apr 23–26. Berne, NY: Association of State Wetland Managers.Google Scholar
Rosenberg, D. M. and Barton, D. R. (1986). The Mackenzie river system. In The Ecology of River Systems, eds. Davies, B. R. and Walker, K. F., pp. 425–33. Dordrecht, the Netherlands: Dr. W. Junk Publishers.CrossRefGoogle Scholar
Rosenberg, D. M., Bodaly, R. A., and Usher, P. J. (1995). Environmental and social impacts of large scale hydro-electric development: who is listening?Global Environmental Change, 5, 127–48.CrossRefGoogle Scholar
Rosenthal, , , G. A. and Berenbaum, M. R. (eds.) (1991). Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego, CA: Academic Press.
Rosenzweig, M. L. (1995). Species Diversity in Space and Time. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Rosgen, D. L. (1994). A classification of natural rivers. Catena, 22, 169–99.CrossRefGoogle Scholar
Rosgen, D. L. (1995). River restoration utilizing natural stability concepts. In Wetlands and Watershed Management: Science Applications and Public Policy, eds. Kusler, J. A., Willard, D. E., and Hull, Jr. H. C., pp. 55–62. A collection of papers from a national symposium and several workshops at Tampa, FL, Apr 23–26. Berne, NY: Association of State Wetland Managers.Google Scholar
Rosswall, T. (1983). The nitrogen cycle. In The Major Biogeochemical Cycles and Their Interactions, SCOPE Report No. 21, eds. Bolin, B. and Cook, R. B., pp. 46–50. Chichester, UK: John Wiley.Google Scholar
Rothhaupt, K. O. (1990). Resource competition of herbivorous zooplankton: a review of approaches and perspectives. Archives in Hydrobiology, 118, 1–29.Google Scholar
Rowe, C. L. and Dunson, W. A.. (1995). Impacts of hydroperiod on growth and survival of larval amphibians in temporary ponds of Central Pennsylvania, USA. Oecologia,102, 397–403.CrossRefGoogle ScholarPubMed
Rozan, T. F., Hunter, K. S., and Benoit, G. (1994). Industrialization as recorded in floodplain deposits of the Quinnipiac River, Connecticut. Marine Pollution Bulletin, 28, 564–9.CrossRefGoogle Scholar
Ryan, P. A. (1991). Environmental effects of sediment on New Zealand streams: a review. New Zealand Journal of Marine and Freshwater Research, 25, 207–21.CrossRefGoogle Scholar
Rybicki, N. B. and Carter, V. (1986). Effect of sediment depth and sediment type on the survival of Vallisneria americana Michx. grown from tubers. Aquatic Botany, 24, 233–40.CrossRefGoogle Scholar
Salisbury, F. B. and Ross, C. W. (1988). Plant Physiology, 3rd edn. Belmont, CA: Wadsworth.Google Scholar
Salisbury, S. E. (1970). The pioneer vegetation of exposed muds and its biological features. Philosophical Transactions of the Royal Society of London Series B, 259, 207–55.CrossRefGoogle Scholar
Salo, J., Kalliola, R., Hakkinen, I., Makinen, Y., Niemela, P., Puhakka, M., and Coley, P. D. (1986). River dynamics and the diversity of Amazon lowland forest. Nature, 322, 254–8.CrossRefGoogle Scholar
Sanders, N. K. (1972). The Epic of Gilgamesh, an English version with an introduction by N. K. Sanders, rev edn. London: Penguin Books.Google Scholar
Sand-Jensen, K. and Krause-Jensen, D. (1997). Broad-scale comparison of photosynthesis in terrestrial and aquatic plant communities. Oikos, 80, 203–8.CrossRefGoogle Scholar
Sansen, U. and Koedam, N. (1996). Use of sod cutting for restoration of wet heathlands: revegetation and establishment of typical species in relation to soil conditions. Journal of Vegetation Science, 7, 483–6.CrossRefGoogle Scholar
Santelmann, M. V. (1991). Influences on the distribution of Carex exilis: an experimental approach. Ecology, 72, 2025–37.CrossRefGoogle Scholar
Sanzone, , S. and McElroy, A. (eds.) (1998). Ecological Impacts and Evaluation Criteria for the Use of Structures in Marsh Management, EPA-SAB-EPEC-98–003. Washington, DC: U.S. Environmental Protection Agency Science Advisory Board.
Sather, J. H. and Smith, R. D. (1984). An Overview of Major Wetland Functions, FWS/OBS-84/18. Washington, DC: U.S. Fish and Wildlife Service.Google Scholar
Sather, J. H., Smith, R. D., and Larson, J. S. (1990). Natural values of wetlands. In Wetlands and Shallow Continental Water Bodies, Vol. 1, Natural and Human Relationships, ed. Patten, B. C., pp. 373–87. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Saucier, R. T. (1963). Recent Geomorphic History of the Pontchartrain Basin. Baton Rouge, LA: Louisiana State University Press.Google Scholar
Saunders, , D. A., Hobbs, R. J., and Ehrlich, P. R. (eds.) (1993). Nature Conservation 3: Reconstruction of Fragmented Ecosystems Global and Regional Perspectives. Chipping Norton, NSW: SurreyBeatty.
Savile, D. B. O. (1956). Known dispersal rates and migratory potentials as clues to the origin of the North American biota. American Midland Naturalist, 56, 434–53.CrossRefGoogle Scholar
Scagel, R. F., Bandoni, R. J., Rouse, G. E., Schofield, W. B., Stein, J. R., and Taylor, T. M. C. (1966). Plant Diversity: An Evolutionary Approach. Belmont, CA: Wadsworth.Google Scholar
Scharf, F. S., Juanes, F., and Sutherland, M. (1998). Inferring ecological relationships from the edges of scatter diagrams: comparison of regression techniques. Ecology, 79, 448–60.CrossRefGoogle Scholar
Schiemer, F., Baumgartner, C., and Tockner, K. (1999). Restoration of floodplain rivers: the ‘Danube Restoration Project’. Regulated Rivers: Research and Management, 15, 231–44.3.0.CO;2-5>CrossRefGoogle Scholar
Schindler, D. W. (1977). Evolution of phosphorus limitation in lakes. Science, 195, 260–2.CrossRefGoogle ScholarPubMed
Schindler, D. W. (1987). Detecting ecosystem responses to anthropogenic stress. Canadian Journal of Fisheries and Aquatic Sciences, 44, 6–25.CrossRefGoogle Scholar
Schneider, E., Tudor, M., and Staras, M, M. (eds.) (2008). Evolution of Babina Polder after Restoration Works: Agricultural Polder Babina, A Pilot Project of Ecological Restoration. Frankfurt am Main, Germany: WWF Germany, and Tulcea, Romania: Danube Delta National Institute for Research and Development.
Schneider, R. (1994). The role of hydrologic regime in maintaining rare plant communities of New York's coastal plain pondshores. Biological Conservation, 68, 253–60.CrossRefGoogle Scholar
Schnitzler, A. (1995). Successional status of trees in gallery forest along the river Rhine. Journal of Vegetation Science, 6, 479–86.CrossRefGoogle Scholar
Schoener, T. W. (1974). Resource partitioning in ecological communities. Science, 185, 27–39.CrossRefGoogle ScholarPubMed
Schoener, T. W. (1985). Some comments on Connell's and my reviews of field experiments on interspecific competition. The American Naturalist, 125, 730–40.CrossRefGoogle Scholar
Scholander, P. F., Hammel, H. T., Bradstreet, E. D., and Hemmingsen, E. A. (1965). Sap pressure in vascular plants. Science, 148, 339–46.CrossRefGoogle ScholarPubMed
Schröder, H. K., Andersen, H. E., Kiehl, K., and Kenkel, N. (2005). Rejecting the mean: estimating the response of fen plant species to environmental factors by non-linear quantile regression. Journal of Vegetation Science, 16, 373–82.CrossRefGoogle Scholar
Schubel, J. R., Shen, H., and Park, M. (1986). Comparative analysis of estuaries bordering the Yellow Sea. In Estuarine Variability, ed. Wolfe, D. A., pp. 43–62. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Schuyt, K. and Brander, L. (2004). Living Waters: Conserving the Source of Life – The Economic Values of the World's Wetlands. Amsterdam, the Netherlands: European Union, and Gland, Switzerland: World Wildlife Fund.Google Scholar
Scott, W. S. and Wylie, N. P. (1980). The environmental effects of snow dumping: a literature review. Journal of Environmental Management, 10, 219–40.Google Scholar
Sculthorpe, C. D. (1967). The Biology of Aquatic Vascular Plants. Reprinted 1985 Edward Arnold, by London.Google Scholar
Segers, R. (1998). Methane production and methane consumption: a review of processes underlying wetland methane fluxes. Biogeochemistry, 41, 23–51.CrossRefGoogle Scholar
Seidl, A. F. and Moraes, A. S. (2000). Global valuation of ecosystem services: application to the Pantanal da Nhecolandia, Brazil. Ecological Economics, 33, 1–6.CrossRefGoogle Scholar
Serbesoff-King, K. (2003). Melaleuca in Florida: a literature review on the taxonomy, distribution, biology, ecology, economic importance and control measures. Journal of Aquatic Plant Management, 41, 98–112.Google Scholar
Severinghaus, W. D. (1981). Guild theory development as a mechanism for assessing environmental impact. Environmental Management, 5, 187–90.CrossRefGoogle Scholar
Seward, A. C. (1931). Plant Life Through the Ages. London: Cambridge University Press.Google Scholar
Shaffer, G. P., Sasser, C. E., Gosselink, J. G., and Rejmanek, M. (1992). Vegetation dynamics in the emerging Atchafalaya Delta, Louisiana, USA. Journal of Ecology, 80, 677–87.CrossRefGoogle Scholar
Shankman, D., Keim, B. D., and Song, J. (2006). Flood frequency in China's Poyang Lake region: trends and teleconnections. International Journal of Climatology, 26, 1255–66.CrossRefGoogle Scholar
Shannon, R. D., White, J. R., Lawson, J. E., and Gilmour, B. S. (1996). Methane efflux from emergent vegetation in peatlands. Journal of Ecology, 84, 239–46.CrossRefGoogle Scholar
Sharitz, R. R. and McCormick, J. F. (1973). Population dynamics of two competing annual plant species. Ecology, 54, 723–40.CrossRefGoogle Scholar
Sharitz, , R. R. and Gibbons, J. W. (eds.) (1989). Freshwater Wetlands and Wildlife. Proceedings of a symposium held at Charleston, South Carolina, Mar 24–27, 1986. Washington, DC: U.S. Department of Energy.
Sharitz, R. R. and Mitsch, W. J. (1993). Southern floodplain forests. In Biodiversity of the Southeast United States/Lowland Terrestrial Communities, eds. Martin, W. H., , S. G.Boyce, , and Echternacht, A. C., pp. 311–71. New York: John Wiley.Google Scholar
Sharp, M. J. and Keddy, P. A. (1985). Biomass accumulation by Rhexia virginica and Triadenum fraseri along two lakeshore gradients: a field experiment. Canadian Journal of Botany, 63, 1806–10.CrossRefGoogle Scholar
Shay, J. M. and Shay, C. T. (1986). Prairie marshes in western Canada, with specific reference to the ecology of five emergent macrophytes. Canadian Journal of Botany, 64, 443–54.CrossRefGoogle Scholar
Sheail, J. and Wells, T. C. E. (1983). The Fenlands of Huntingdonshire, England: a case study in catastrophic change. In Ecosystems of the World, Vol. 4B, Mires: Swamp, Bog, Fen and Moor – Regional Studies, ed. Gore, A. J. P., pp. 375–93. Amsterdam, the Netherlands: Elsevier.Google Scholar
Sheldon, S. P. (1987). The effects of herbivorous snails on submerged macrophyte communities in Minnesota lakes. Ecology, 68, 1920–31.CrossRefGoogle ScholarPubMed
Sheldon, S. P. (1990). More on freshwater snail herbivory: a reply to Bronmark. Ecology, 71, 1215–16.CrossRefGoogle Scholar
Shimwell, D. W. (1971). The Description and Classification of Vegetation. Seattle, WA: University of Washington Press.Google Scholar
Shipley, B. (2000). Cause and Correlation in Biology. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Shipley, B. and Keddy, P. A. (1987). The individualistic and community-unit concepts as falsifiable hypotheses. Vegetatio, 69, 47–55.CrossRefGoogle Scholar
Shipley, B. and Keddy, P. A. (1994). Evaluating the evidence for competitive hierarchies in plant communities. Oikos, 69, 340–5.CrossRefGoogle Scholar
Shipley, B. and Parent, M. (1991). Germination responses of 64 wetland species in relation to seed size, minimum time to reproduction and seedling relative growth rate. Functional Ecology, 5, 111–18.CrossRefGoogle Scholar
Shipley, B. and Peters, R. H. (1990). A test of the Tilman model of plant strategies: relative growth rate and biomass partitioning. The American Naturalist, 136, 139–53.CrossRefGoogle Scholar
Shipley, B., Keddy, P. A., Moore, D. R. J., and Lemky, K. (1989). Regeneration and establishment strategies of emergent macrophytes. Journal of Ecology, 77, 1093–110.CrossRefGoogle Scholar
Shipley, B., Keddy, P. A., Gaudet, C., and Moore, D. R. J. (1991a). A model of species density in shoreline vegetation. Ecology, 72, 1658–67.CrossRefGoogle Scholar
Shipley, B., Keddy, P. A., and Lefkovitch, L. P. (1991b). Mechanisms producing plant zonation along a water depth gradient: a comparison with the exposure gradient. Canadian Journal of Botany, 69, 1420–4.CrossRefGoogle Scholar
Shrader-Frechette, K. S. and McCoy, E. D. (1993). Methods in Ecology: Strategies for Conservation. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Siegel, S. (1956). Nonparametric Statistics for the Behavioral Sciences. New York: McGraw-Hill.Google Scholar
Silander, J. A. and Antonovics, J. (1982). Analysis of interspecific interactions in a coastal plant community: a perturbation approach. Nature, 298, 557–60.CrossRefGoogle Scholar
Silliman, B. R. and Zieman, J. C. (2001). Top-down control of Spartina alterniflora production by periwinkle grazing in a Virginia salt marsh. Ecology, 82, 2830–845.CrossRefGoogle Scholar
Silliman, B. R., Grosholz, E. D., and Bertness, M. D. (eds.) (2009). Human Impacts on Salt Marshes: A Global Perspective. Berkeley, CA: University of California Press.
Silvola, J., Alm, J., Ahlholm, U., Nykanen, H., and Martikainen, P. J. (1996). CO2 fluxes from peat in boreal mires under varying temperature and moisture conditions. Journal of Ecology, 84, 219–28.CrossRefGoogle Scholar
Simberloff, D. and Dayan, T. (1991). The guild concept and the structure of ecological communities. Annual Review of Ecology and Systematics, 22, 115–43.CrossRefGoogle Scholar
Simons, M. (1997). Big, bold effort revives the Danube wetlands. The New York Times, Oct 19, 1997, pp. 1, 8.Google Scholar
Sinclair, A. R. E. (1983). The adaptations of African ungulates and their effects on community function. In Tropical Savannas, ed. Boulière, F., pp. 401–22. Amsterdam, the Netherlands: Elsevier.Google Scholar
Sinclair, A. R. E. and Fryxell, J. M. (1985). The Sahel of Africa: ecology of a disaster. Canadian Journal of Zoology, 63, 987–94.CrossRefGoogle Scholar
Sinclair, A. R. E, Hik, D. S., Schmitz, O. J., Scudder, G. G. E., Turpin, D. H., and Larter, N. C. (1995). Biodiversity and the need for habitat renewal. Ecological Applications, 5, 579–87.CrossRefGoogle Scholar
Sioli, H. (1964). General features of the limnology of Amazonia. Verhandlungen Internationale Vereinigung für theoretische und angewandte Limnologie, 15, 1053–8.Google Scholar
Sioli, H. (1986). Tropical continental aquatic habitats. In Conservation Biology: The Science of Scarcity and Diversity, ed. Soulé, M. E., pp. 383–93. Sunderland, MA: Sinauer Associates.Google Scholar
Sippel, S. J., Hamilton, S. K., Melack, J. M., and Novo, E. M. M. (1998). Passive microwave observations of inundation area and the area/stage relation in the Amazon River floodplain. International Journal of Remote Sensing, 19, 3055–74.CrossRefGoogle Scholar
Skellam, J. G. (1951). Random dispersal in theoretical populations. Biometrika, 38, 196–218.CrossRefGoogle ScholarPubMed
Sklar, , F. H. and Valk, A. G. (eds.) (2002). Tree Islands of the Everglades. Dordrecht, the Netherlands: Kluwer.CrossRef
Sklar, F. H., Chimney, M. J., Newman, S., McCormick, P., Gawlick, D., Miao, S., McVoy, C., Said, W., Newman, J., Coronado, C., Crozier, G., Korvela, M., and Rutchey, K. (2005). The ecological–societal underpinnings of Everglades restoration. Frontiers in Ecology and Environment, 3, 161–9.Google Scholar
Slack, N. G., Vitt, D. H., and Horton, D. G. (1980). Vegetation gradients of minerotrophically rich fens in western Alberta. Canadian Journal of Botany, 58, 330–50.CrossRefGoogle Scholar
Slovic, P. (1987). Perception of risk. Science, 236, 280–5.CrossRefGoogle ScholarPubMed
Smart, R. M. and Barko, J. W. (1978). Influence of sediment salinity and nutrients on the physiological ecology of selected salt marsh plants. Estuarine and Coastal Marine Science, 7, 487–95.CrossRefGoogle Scholar
Smith, D. C. and Douglas, A. E. (1987). The Biology of Symbiosis. London: Edward Arnold.Google Scholar
Smith, D. W. and Cooper, S. D. (1982). Competition among Cladocera. Ecology, 63, 1004–15.CrossRefGoogle Scholar
Smith, E. K. (2006). Bog Turtle (Clemmys muhlenbergii), Fish and Wildlife Habitat Management Leaflet No. 44. Natural Resources Conservation Service, Washington, D.C. and Wildlife Habitat Council, Silver Spring. MD. ftp-fc.sc.egov.usda.gov/WHMI/WEB/pdf/TechnicalLeaflets/bog_turtle_Oct%2023.pdf
Smith, L. M. (2003). Playas of the Great Plains. Austin, TX: University of Texas Press.Google Scholar
Smith, L. M. and Kadlec, J. A. (1983). Seed banks and their role during the drawdown of a North American marsh. Journal of Applied Ecology, 20, 673–84.CrossRefGoogle Scholar
Smith, L. M. and Kadlec, J. A. (1985a). Fire and herbivory in a Great Salt Lake marsh. Ecology, 66, 259–65.CrossRefGoogle Scholar
Smith, L. M. and Kadlec, J. A. (1985b). Comparisons of prescribed burning and cutting of Utah marsh plants. Great Basin Naturalist, 45, 463–6.Google Scholar
Smith, P. G. R., Glooschenko, V., and Hagen, D. A. (1991). Coastal wetlands of three Canadian Great Lakes: inventory, current conservation initiatives, and patterns of variation. Canadian Journal of Fisheries and Aquatic Sciences, 48, 1581–94.CrossRefGoogle Scholar
Smith, V. H. (1982). The nitrogen and phosphorus dependence of algal biomass in lakes: an empirical and theoretical analysis. Limnology and Oceanography, 27, 1101–12.CrossRefGoogle Scholar
Smith, V. H. (1983). Low nitrogen to phosphorus ratios favor dominance by bluegreen algae in lake phytoplankton. Science, 221, 669–71.CrossRefGoogle Scholar
Snodgrass, J. W., Komoroski, M. J., Bryan, A. L., and Burger, J. (2000). Relationships among isolated wetland size, hydroperiod, and amphibian species richness: implications for wetland regulations. Conservation Biology, 14, 414–19.CrossRefGoogle Scholar
Snow, A. A. and Vince, S. W. (1984). Plant zonation in an Alaskan salt marsh. II. An experimental study of the role of edaphic conditions. Journal of Ecology, 72, 669–84.CrossRefGoogle Scholar
,Society for Ecological Restoration International Science and Policy Working Group (SER). (2004). The SER International Primer on Ecological Restoration. Tucson, AZ: Society for Ecological Restoration. www.ser.orgGoogle Scholar
Sousa, W. P. (1984). The role of disturbance in natural communities. Annual Review of Ecology and Systematics, 15, 353–91.CrossRefGoogle Scholar
Southwood, T. R. E. (1977). Habitat, the templet for ecological strategies?Journal of Animal Ecology, 46, 337–65.CrossRefGoogle Scholar
Southwood, T. R. E. (1988). Tactics, strategies, and templets. Oikos, 52, 3–18.CrossRefGoogle Scholar
Specht, A. and Specht, R. L. (1993). Species richness and canopy productivity of Australian plant communities. Biodiversity and Conservation, 2, 152–67.CrossRefGoogle Scholar
Spence, D. H. N. (1964). The macrophytic vegetation of freshwater lochs, swamps and associated fens. In The Vegetation of Scotland, ed. Burnett, J. H., pp. 306–425. Edinburgh, UK: Oliver and Boyd.Google Scholar
Spence, D. H. N. (1982). The zonation of plants in freshwater lakes. Advances in Ecological Research, 12, 37–125.CrossRefGoogle Scholar
Spencer, D. F. and Ksander, G. G. (1997). Influence of anoxia on sprouting of vegetative propagules of three species of aquatic plants. Wetlands, 17, 55–64.CrossRefGoogle Scholar
Springuel, I. (1990). Riverain vegetation in the Nile valley in Upper Egypt. Journal of Vegetation Science, 1, 595–8.CrossRefGoogle Scholar
Starfield, A. M. and Bleloch, A. L. (1991). Building Models for Conservation and Wildlife Management, 2nd edn. Edina, MN: Burgers International Group.Google Scholar
Stead, I. M., Bourke, J. B., and Brothwell, D. (1986). Lindow Man: The Body in the Bog. London: British Museum Publications.Google Scholar
Steedman, R. J. (1988). Modification and assessment of an index of biotic integrity to quantify stream quality in southern Ontario. Canadian Journal of Fisheries and Aquatic Sciences, 45, 492–501.CrossRefGoogle Scholar
Stevens, P. W., Fox, S. L., and Montague, C. L. (2006). The interplay between mangroves and saltmarshes at the transition between temperate and subtropical climate in Florida. Wetlands Ecology and Management, 14, 435–44.CrossRefGoogle Scholar
Stevenson, J. C., Ward, L. G., and Kearney, M. S. (1986). Vertical accretion in marshes with varying rates of sea level rise. In Estuarine Variability, ed. Wolfe, D. A., pp. 241–59. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Stewart, R. E. and Kantrud, H. A. (1971). Classification of Natural Ponds and Lakes in the Glaciated Prairie Region, Resource Publication No. 92. Washington, DC: U.S. Fish and Wildlife Service.Google Scholar
Stewart, W. N. and Rothwell, G. W. (1993). Paleobotany and the Evolution of Plants, 2nd edn. New York: Cambridge University Press.Google Scholar
Strahler, A. N. (1971). The Earth Sciences, 2nd edn. New York: Harper and Row.Google Scholar
Street, F. A. and Grove, A. T. (1979). Global maps of lake-level fluctuations since 30,000 yrs B.P. Quaternary Research, 12, 83–118.CrossRefGoogle Scholar
Stuart, S. A., Choat, B., Martin, K. C., Holbrook, N. M., and Ball, M. C. (2007). The role of freezing in setting the latitudinal limits of mangrove forests. New Phytologist, 173, 576–83.CrossRefGoogle ScholarPubMed
Stuckey, R. L. (1975). A floristic analysis of the vascular plants of a marsh at Perry's Victory Monument, Lake Erie. Michigan Botanist, 14, 144–66.Google Scholar
Sutter, R. D. and Kral, R. (1994). The ecology, status, and conservation of two nonalluvial wetland communities in the south Atlantic and eastern Gulf coastal plain, USA. Biological Conservation, 68, 235–43.CrossRefGoogle Scholar
Swink, F. and Wilhelm, G. (1994). Plants of the Chicago Region 4th edn. Indianapolis, IN: Indiana Academy of Science.Google Scholar
Szalay, F. A. and Resh, V. H. (1997). Responses of wetland invertebrates and plants important in waterfowl diets to burning and mowing of emergent vegetation. Wetlands, 17, 149–56.CrossRefGoogle Scholar
Szczepanski, A. J. (1990). Forested wetlands of Poland. In Forested Wetlands, ed. Lugo, A. E., Brinson, M. and Brown, S., pp. 437–46. Amsterdam, the Netherlands: Elsevier.Google Scholar
Taiz, L. and Zeiger, E. (1991). Plant Physiology. Menlo Park, CA: Benjamin Cummings.Google Scholar
Talling, J. F. (1992). Environmental regulation in African shallow lakes and wetlands. Revue d'Hydrobiologie Tropicale, 25, 87–144.Google Scholar
Tallis, J. H. (1983). Changes in wetland communities. In Ecosystems of the World, Vol. 4A, Mires: Swamp, Bog, Fen and Moor – General Studies, ed. Gore, A. J. P., pp. 311–47. Amsterdam, the Netherlands: Elsevier.Google Scholar
Tans, P. (2009). Recent monthly mean CO2 at Mauna Loa. www.esol.noaa.gov/gmd/ccgg/trends (accessed May 7, 2009)
Tansley, A. G. (1939). The British Islands and Their Vegetation. Cambridge, UK: Cambridge University Press.Google Scholar
Tansley, A. G. and Adamson, R. S. (1925). Studies of the vegetation of the English chalk. III. The chalk grasslands of the Hampshire–Sussex border. Journal of Ecology, 13, 177–223.CrossRefGoogle Scholar
Tarr, T. L., Baber, M. J., and Babbitt, K. J. (2005). Invertebrate community structure across a wetland hydroperiod gradient in southern New Hampshire, USA. Wetlands Ecology and Management, 13, 321–34.CrossRefGoogle Scholar
Taylor, D. R., Aarssen, L. W., and Loehle, C. (1990). On the relationship between r/K selection and environmental carrying capacity: a new habitat templet for plant life history strategies. Oikos, 58, 239–50.CrossRefGoogle Scholar
Taylor, J. A. (1983). The peatlands of Great Britain and Ireland. In Ecosystems of the World, Vol. 4B, Mires: Swamp, Bog, Fen and Moor – Regional Studies, ed. Gore, A. J. P., pp. 1–46. Amsterdam, the Netherlands: Elsevier.Google Scholar
Taylor, K. L. and Grace, J. B. (1995). The effects of vertebrate herbivory on plant community structure in the coastal marshes of the Pearl River, Louisiana, USA. Wetlands, 15, 68–73.CrossRefGoogle Scholar
Taylor, R. B., Josenhans, H., Balcom, B. A., and Johnston, A. J. B. (2000). Louisbourg Harbour through Time, Geological Survey of Canada Open File Report 3896. Ottawa, ON: Geological Survey of Canada.CrossRefGoogle Scholar
Teller, J. T. (1988). Lake Agassiz and its contribution to flow through the Ottawa–St. Lawrence system. In The Late Quaternary Development of the Champlain Sea Basin, Geological Association of Canada Special Paper No. 35, ed. Gadd, N. R., pp. 281–9. St. John's, Nfld: Geological Association of Canada.Google Scholar
Teller, J. T. (2003). Controls, history, outbursts and impact of large late-Quaternary proglacial lakes in North America. In The Quaternary Period in the United States, eds. Gilespie, A., Porter, S., and Atwater, B., pp. 45–61. Amsterdam, the Netherlands: Elsevier.CrossRefGoogle Scholar
Terborgh, J. and Robinson, S. (1986). Guilds and their utility in ecology. In Community Ecology: Pattern and Process, eds. Kikkawa, J. and Anderson, D. J., pp. 65–90. Melbourne, Vic: Blackwell Scientific Publications.Google Scholar
Thibodeau, F. R. and Ostro, B. D. (1981). An economic analysis of wetland protection. Journal of Environmental Management, 12, 19–30.Google Scholar
Thirgood, J. V. (1981). Man and the Mediterranean Forest: A History of Resource Depletion. London: Academic Press.Google Scholar
Thomas, J. D. (1982). Chemical ecology of the snail hosts of Schistosomiasis: snail–snail and snail–plant interactions. Malacologia, 22, 81–91.Google Scholar
Thomas, , J. and Nygard, J. (eds.) (2007). The Importance of Habitat Created by Molluscan Shellfish to Managed Species along the Atlantic Coast of the United States, Habitat Management Series No. 8. Washington, DC: Atlantic States Marine Fisheries Commission.
Thompson, D. J. and Shay, J. M. (1988). First-year response of a Phragmites marsh community to seasonal burning. Canadian Journal of Botany, 67, 1448–55.CrossRefGoogle Scholar
Thompson, K. (1985). Emergent plants of the permanent and seasonally-flooded wetlands. In The Ecology and Management of African Wetland Vegetation, ed. Denny, P., pp. 43–107. Dordrecht, the Netherlands: Dr. W. Junk Publishers.Google Scholar
Thompson, K. and Hamilton, A. C. (1983). Peatlands and swamps of the African continent. In Ecosystems of the World, Vol. 4B, Mires: Swamp, Bog, Fen and Moor – Regional Studies, ed. Gore, A. J. P., pp. 331–73. Amsterdam, the Netherlands: Elsevier.Google Scholar
Thoreau, H. D. (1854). Republished 1965 as Walden and Civil Disobedience. New York: Airmont.Google Scholar
Tilman, D. (1982). Resource Competition and Community Structure. Princeton, NJ: Princeton University Press.Google ScholarPubMed
Tilman, D. (1986). Evolution and differentiation in terrestrial plant communities: the importance of the soil resource: light gradient. In Community Ecology, eds. Diamond, J. and Case, T. J., pp. 359–80. New York: Harper and Row.Google Scholar
Tilman, D. (1988). Plant Strategies and the Dynamics and Structure of Plant Communities. Princeton, NJ: Princeton University Press.Google Scholar
Tiner, R. W. (1999). Wetland Indicators: A Guide to Wetland Identification, Delineation, Classification and Mapping. Boca Raton, FL: CRC Press.CrossRefGoogle Scholar
Tinkle, W. J. (1939). Fundamentals of Zoology. Grand Rapids, MI: Zondervan.Google Scholar
Todd, T. N. and Davis, B. M. (1995). Effects of fish density and relative abundance on competition between larval lake herring and lake whitefish for zooplankton. Archiv für Hydrobiologie, Special Issues in Advanced Limnology, 46, 163–71.Google Scholar
Tomlinson, P. B. (1986). The Botany of Mangroves. Cambridge, UK: Cambridge University Press.Google Scholar
Toner, M. and Keddy, P. A. (1997). River hydrology and riparian wetlands: a predictive model for ecological assembly. Ecological Applications, 7, 236–46.CrossRefGoogle Scholar
Tonn, W. M. and Magnuson, J. J. (1982). Patterns in the species composition and richness of fish assemblages in northern Wisconsin lakes. Ecology, 63, 1149–66.CrossRefGoogle Scholar
Tonn, W. M., Magnuson, J. J., and Forbes, A. M. (1983).Community analysis in fishery management: an application with northern Wisconsin lakes. Transactions of the American Fisheries Society, 112, 368–77.2.0.CO;2>CrossRefGoogle Scholar
Toth, L. A. (1993). The ecological basis of the Kissimmee River restoration plan. Florida Scientist, 1, 25–51.
Townsend, A. R., Braswell, B. H., Holland, E. A., and Penner, J. E. (1996). Spatial and temporal patterns in terrestrial carbon storage due to deposition of fossil fuel nitrogen. Ecological Applications, 6, 806–14.CrossRefGoogle Scholar
Townsend, G. H. (1984). Simulating the Effect of Water Regime Restoration Measures on Wildlife Populations and Habitat within the Peace–Athabasca Delta, Technical report No. 13. Saskatoon, Sask.: Western and Northern Region, Canadian Wildlife Service.Google Scholar
Trombulak, S. C. and Frissell, C. A. (2000). Review of ecological effects of roads on terrestrial and aquatic communities. Conservation Biology, 14, 18–30.CrossRefGoogle Scholar
Tsuyuzaki, S. and Tsujii, T. (1990). Preliminary study on grassy marshland vegetation, western part of Sichuan Province, China, in relation to yak-grazing. Ecological Research, 5, 271–6.CrossRefGoogle Scholar
Tsuyuzaki, S., Urano, S., and Tsujii, T. (1990). Vegetation of alpine marshland and its neighboring areas, northern part of Sichuan Province, China.Vegetatio, 88, 79–86.CrossRefGoogle Scholar
Tuchman, B. (1984). The March of Folly. New York: Ballantine Books.Google Scholar
Turner, C. E., Center, T. D., Burrows, D. W., and Buckingham, G. R. (1998). Ecology and management of Melaleuca quinquenervia, an invader of wetlands in Florida, U.S.A. Wetlands Ecology and Management, 5, 165–78.CrossRefGoogle Scholar
Turner, R. E. (1977). Intertidal vegetation and commercial yields of penaeid shrimp. Transactions of the American Fisheries Society, 106, 411–16.2.0.CO;2>CrossRefGoogle Scholar
Turner, R. E.(1982). Protein yields from wetlands. In Wetlands: Ecology and Management, Proceedings of the First International Wetlands Conference, New Delhi, India, Sept 10–17, 1980.Google Scholar
Turner, R. E. and Rabelais, N. N. (2003). Linking landscape and water quality in the Mississippi River Basin for 200 years. BioScience, 53, 563–72.CrossRefGoogle Scholar
Turner, R. E. and Streever, B. (2002). Approaches to Coastal Wetland Restoration: Northern Gulf of Mexico. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Turner, R. E., Baustian, J. J., Swenson, E. M., and Spicer, J. S. (2006). Wetland sedimentation from hurricanes Katrina and Rita. Science, 314, 449–52.CrossRefGoogle ScholarPubMed
Turner, R. M. and Karpiscak, M. M. (1980). Recent Vegetation Changes along the Colorado River between Glen Canyon Dam and Lake Mead, Arizona, Geological Survey Professional Paper No. 1132. Washington, DC: U.S. Government Printing Office.Google Scholar
Tyler, G. (1971). Hydrology and salinity of Baltic sea-shore meadows: studies in the ecology of Baltic sea-shore meadows III. Oikos, 22, 1–20.CrossRefGoogle Scholar
Twolan-Strutt, L. and Keddy, P. A. (1996). Above- and below-ground competition intensity in two contrasting wetland plant communities. Ecology, 77, 259–70.CrossRefGoogle Scholar
Underwood, A. J. (1978). The detection of non-random patterns of distribution of species along a gradient. Oecologia, 36, 317–26.CrossRefGoogle ScholarPubMed
Underwood, A. J. (1986). The analysis of competition by field experiments. In Community Ecology: Pattern and Process, eds. Kikkawa, J. and Anderson, D. J., pp. 240–68. Melbourne, Vic: Blackwell Scientific Publications.Google Scholar
Urban, D. L. and Shugart, H. H. (1992). Individual based models of forest succession. In Plant Succession, eds. Glenn-Lewin, D. C., Peet, R. K., and Veblen, T. T., pp. 249–92. London: Chapman and Hall.Google Scholar
,U.S. Army Coastal Engineering Research Centre. (1977). Shore Protection Manual, Vol. 1, 3rd edn. Washington, DC: U.S. Government Printing Office.Google Scholar
,U.S. Army Corps of Engineers. (1987). Corps of Engineers Wetlands Delineation Manual, Technical Report No. Y-87–1. Vicksburg, MS: Department of the Army, Waterways Experiment Station.Google Scholar
,U.S. Army Corps of Engineers. (2004). The Mississippi River and Tributaries Project. U.S. ACE, New Orleans District. www.mvn.usace.army.mil/pao/bro/misstrib.htm (accessed Apr 7, 2009)Google Scholar
,U.S. Environmental Protection Agency. (2004). Constructed Treatment Wetlands, EPA 843-F-03–013. Washington, DC: U.S. Government Printing Office.Google Scholar
,U. S. Fish and Wildlife Service. (1989). Louisiana Pearlshell (Margaritifera hembeli) Recovery Plan. Jackson, MS: U.S. Fish and Wildlife Service.Google Scholar
,U.S. Fish and Wildlife Service. (2001). Bog Turtle (Clemmys muhlenbergii), Northern Population, Recovery Plan. Hadley, MA: U.S. Fish and Wildlife Service.Google Scholar
,U.S. Geological Survey. (1996). http://earthshots.usgs.gov/Knife/Knife. (accessed June 15, 2009)
,U.S. Geological Survey. (2000). Sea Level and Climate, U.S.G.S. Fact Sheet No. 002–00. Reston, VA: U.S. Department of the Interior.Google Scholar
Valiela, I., Foreman, K., LaMontagne, M., Hersh, D., Costa, J., D'Avanzo, C., Babione, M., Sham, C., Brawley, J., Peckol, P., DeMeo-Anderson, B., and Lajtha, K. (1992). Couplings of watersheds and coastal waters: sources and consequences of nutrient enrichment in Waquoit Bay, Massachusetts. Estuaries, 15, 443–57.CrossRefGoogle Scholar
Valentine, D. L. (2002). Biogeochemistry and microbial ecology of methane oxidation in anoxic environments: a review. Journal Antonie van Leeuwenhoek, 81, 271–82.CrossRefGoogle ScholarPubMed
Vallentyne, J. R. (1974). The Algal Bowl: Lakes and Man, Miscellaneous Special Publication No. 22. Ottawa, ON: Department of the Environment, Fisheries and Marine Service.Google Scholar
Breeman, N. (1995). How Sphagnum bogs down [sic] other plants. Trends in Ecology and Evolution, 10, 270–5.CrossRefGoogle Scholar
Kieft, C. (1991). The Low Countries. In The New Encyclopedia Britannica, 15th edn, Vol. 23, 314–25. Chicago, IL: Encyclopedia Britannica Inc.Google Scholar
Rijt, C. W. C. J., Hazelhoff, L., and Blom, C. W. P. M. (1996). Vegetation zonation in a former tidal area: a vegetation-type response model based on DCA and logistic regression using GIS. Journal of Vegetation Science, 7, 505–18.CrossRefGoogle Scholar
Leeden, F., Troise, F., and Tood, D. K. (eds.) (1990). The Water Encyclopedia, 2nd edn. Chelsea, MI: Lewis Publishers.Google Scholar
Pijl, L. (1972). Principles of Dispersal in Higher Plants. New York: Springer-Verlag.CrossRefGoogle Scholar
Toorn, J., Verhoeven, J. T. A., and Simpson, R. L. (1990). Fresh water marshes. In Wetlands and Shallow Continental Water Bodies, Vol. 1, ed. Patten, B. C., pp. 445–65. The Hague, the Netherlands: SPB Academic Publishing.Google Scholar
Valk, A. G. (1981). Succession in wetlands: a Gleasonian approach. Ecology, 62, 688–96.CrossRefGoogle Scholar
Valk, A. G. (1988). From community ecology to vegetation management: providing a scientific basis for management. In Transactions of the 53 North American Wildlife and Natural Resources Conference, pp. 463–70. Washington, DC: Wildlife Management Institute.Google Scholar
Valk, A. G. (1989). Northern Prairie Wetlands. Ames, IA: Iowa State University Press.Google Scholar
Valk, A. G. and Davis, C. B. (1976). The seed banks of prairie glacial marshes. Canadian Journal of Botany, 54, 1832–8.CrossRefGoogle Scholar
Valk, A. G. and Davis, C. B. (1978). The role of seed banks in the vegetation dynamics of prairie glacial marshes. Ecology, 59, 322–35.CrossRefGoogle Scholar
Valk, A. G., Swanson, S. D., and Nuss, R. F. (1983). The response of plant species to burial in three types of Alaskan wetlands. Canadian Journal of Botany, 61, 1150–64.CrossRefGoogle Scholar
Valk, A. G., Pederson, R. L., and Davis, C. B. (1992). Restoration and creation of freshwater wetlands using seed banks. Wetlands Ecology and Management, 1, 191–7.CrossRefGoogle Scholar
Wijck, C. and Groot, C. J. (1993). The impact of desiccation of a freshwater marsh (Garcines Nord, Camargue, France) on sediment–water–vegetation interactions. Hydrobiologia, 252, 95–103.CrossRefGoogle Scholar
Vasseur, L. and Catto, N. R. (2008). Atlantic Canada. In From Impacts to Adaptation: Canada in a Changing Climate 2007, eds. Lemmen, D. S., Warren, F. J., Lacroic, J., and Bush, E., pp. 119–70. Ottawa, ON: Government of Canada.Google Scholar
Verhoeven, J. T. A. and Liefveld, W. M. (1997). The ecological significance of organochemical compounds in Sphagnum. Acta Botanica Neerlandica, 46, 117–30.CrossRefGoogle Scholar
Verhoeven, J. T. A., Kemmers, R. H. and Koerselman, W. (1993). Nutrient enrichment of freshwater wetlands. In Landscape Ecology of a Stressed Environment, eds. Vos, C. C. and Opdam, P., pp. 33–59. London: Chapman and Hall.CrossRefGoogle Scholar
Verhoeven, J. T. A., Koerselman, W., and Meuleman, A. F. M. (1996). Nitrogen- or phosphorus-limited growth in herbaceous, wet vegetation: relations with atmospheric inputs and management regimes. Trends in Ecology and Evolution, 11, 493–7.CrossRefGoogle ScholarPubMed
Verry, E. S. 1989. Selection and management of shallow water impoundments for wildlife. In Freshwater Wetlands and Wildlife, eds. Sharitz, R. R. and Gibbons, J. W., pp. 1177–94. Washington, DC: U.S. Department of Energy.Google Scholar
Vesey-FitzGerald, D. F. (1960). Grazing succession among East African game animals. Journal of Mammalogy, 41, 161–72.CrossRefGoogle Scholar
Vijayakumar, S. P., Vasudevan, K., and Ishwar, N. M. (2001). Herpetofaunal mortality on roads in the Anamalai Hills, Southern Western Ghats. Hamadryad, 26, 265–72.Google Scholar
Vince, S. W. and Snow, A. A. (1984). Plant zonation in an Alaskan salt marsh. I. Distribution, abundance, and environmental factors. Journal of Ecology, 72, 651–67.CrossRefGoogle Scholar
Vitousek, P. M. (1982). Nutrient cycling and nitrogen use efficiency. The American Naturalist, 119, 553–72.CrossRefGoogle Scholar
Vitousek, P. M., Aber, J., Howarth, R. W., Likens, G. E., Matson, P. A., Schindler, D. W., Schlesinger, W. H., and Tilman, G. D. (1997). Human Alteration of Global Nitrogen Cycle: Causes and Consequences, Issues in Ecology No. 1, Washington, DC: Ecological Society of America.Google Scholar
Vitt, D. H. (1990). Growth and production dynamics of boreal mosses over climatic, chemical and topographic gradients. Botanical Journal of the Linnean Society, 104, 35–59.CrossRefGoogle Scholar
Vitt, D. H. (1994). An overview of factors that influence the development of Canadian peatlands. Memoirs of the Entomological Society of Canada, 169, 7–20.CrossRefGoogle Scholar
Vitt, D. H. and Chee, W. (1990). The relationships of vegetation to surface water chemistry and peat chemistry in fens of Alberta, Canada. Vegetatio, 89, 87–106.CrossRefGoogle Scholar
Vitt, D. H. and Slack, N. G. (1975). An analysis of the vegetation of Sphagnum-dominated kettle-hole bogs in relation to environmental gradients. Canadian Journal of Botany, 53, 332–59.CrossRefGoogle Scholar
Vitt, D. H. and Slack, N. G. (1984). Niche diversification of Sphagnum relative to environmental factors in northern Minnesota peatlands. Canadian Journal of Botany, 62, 1409–30.CrossRefGoogle Scholar
Vitt, D. H., Yenhung, L., and Belland, R. J. (1995). Patterns of bryophyte diversity in peatlands of continental western Canada. The Bryologist, 98, 218–27.CrossRefGoogle Scholar
Vivian-Smith, G. (1997). Microtopographic heterogeneity and floristic diversity in experimental wetland communities. Journal of Ecology, 85, 71–82.CrossRefGoogle Scholar
Vogl, R. (1969). One hundred and thirty years of plant succession in a southeastern Wisconsin lowland. Ecology, 50, 248–55.CrossRefGoogle Scholar
Vogl, R. (1973). Effects of fire on the plants and animals of a Florida wetland. American Midland Naturalist, 89, 334–47.CrossRefGoogle Scholar
Vörösmarty, C. J., Fekete, B., and Tucker, B. A. (1996). River Discharge Database, Version 1.0, vols. 0–6. Paris: UNESCO.Google Scholar
Walker, B. H. and Wehrhahn, C. F. (1971). Relationships between derived vegetation gradients and measured environmental variables in Saskatchewan wetlands. Ecology, 52, 85–95.CrossRefGoogle Scholar
Walker, D. (1970). Direction and rate in some British post-glacial hydroseres. In Studies in the Vegetational History of the British Isles, eds. Walker, D. and West, R. G., pp. 117–39. Cambridge, UK: Cambridge University Press.Google Scholar
Walters, C. (1997). Challenges in adaptive management of ripanian and coastal ecosystems. Conservation Ecology, 1(2), www.consecol.org/vol1/iss2/art1/ (accessed June 15, 2008)CrossRefGoogle Scholar
Wang, S., Jurik, T. M., and Valk, A. G. (1994). Effects of sediment load on various stages in the life and death of cattail (Typha × glauca). Wetlands, 14, 166–73.CrossRefGoogle Scholar
Ward, A. and Trimble, S. W.. (2004). Environmental Hydrology, 2nd edn. Boca Raton, FL: CRC Press.Google Scholar
Wassen, M. J., Barendregt, A., Palczynski, A., Smidt, J. T., and Mars, H. (1990). The relationship between fen vegetation gradients, groundwater flow and flooding in undrained valley mire at Biebrza, Poland. Journal of Ecology, 78, 1106–22.CrossRefGoogle Scholar
Waterkeyn, A., Grillas, P., Vanschoenwinkel, B., and Brendonck, L. (2008). Invertebrate community patterns in Mediterranean temporary wetlands along hydroperiod and salinity gradients. Freshwater Biology, 53, 1808–22.CrossRefGoogle Scholar
Waters, T. F. (1995). Sediment in Streams: Sources, Biological Effects, and Control, American Fisheries Society Monograph No. 7. Nashville, TN: American Fisheries Society.Google Scholar
Watts, W. A. and Winter, T. C. (1966). Plant macrofossils from Kirchner Marsh, Minnesota: a paleoecological study. Geological Society of America Bulletin 77, 1339–60.CrossRefGoogle Scholar
Weaver, J. E. and Clements, F. E. (1938). Plant Ecology, 2nd edn. New York: McGraw-Hill.Google Scholar
Weber, W. and Rabinowitz, A. (1996). A global perspective on large carnivore conservation. Conservation Biology, 10, 1046–54.CrossRefGoogle Scholar
Weddle, R. S. (1991). The French Thorn: Rival Explorers in the Spanish Sea, 1682–1762. College Station, TX: Texas A&M University Press.Google Scholar
Weiher, E. (1999). The combined effects of scale and productivity on species richness. Journal of Ecology, 87, 1005–11.CrossRefGoogle Scholar
Weiher, E. and Boylen, C. W. (1994). Patterns and prediction of a and b diversity of aquatic plants in Adirondack (New York) lakes. Canadian Journal of Botany, 72, 1797–804.CrossRefGoogle Scholar
Weiher, E. and Keddy, P. A. (1995). The assembly of experimental wetland plant communities. Oikos, 73, 323–35.CrossRefGoogle Scholar
Weiher, E. and Keddy, P. A. (eds.) (1999). Assembly Rules in Ecological Communities: Perspectives, Advances, Retreats. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Weiher, E., Wisheu, I. C., Keddy, P. A., and Moore, D. R. J. (1996). Establishment, persistence, and management implications of experimental wetland plant communities. Wetlands, 16, 208–18.CrossRefGoogle Scholar
Weiher, E., Clarke, G. D. P., and Keddy, P. A. (1998). Community assembly rules, morphological dispersion, and the coexistence of plant species. Oikos, 81, 309–22.CrossRefGoogle Scholar
Weiher, E., Werf, A., Thompson, K., Roderick, M., Garnier, E., and Eriksson, , O. (1999). Challenging Theophrastus: a common core list of plant traits for functional ecology. Journal of Vegetation Science, 10, 609–20.CrossRefGoogle Scholar
Wein, R. W. (1983). Fire behaviour and ecological effects in organic terrain. In The Role of Fire in Northern Circumpolar Ecosystems, eds. Wein, R. W. and Maclean, D. A., pp. 81–95. New York: John Wiley.Google Scholar
Weinberg, G. M. (1975). An Introduction to General Systems Thinking. New York: John Wiley.Google Scholar
Weisner, S. E. B. (1990). Emergent Vegetation in Eutrophic Lakes: Distributional Patterns and Ecophysiological Constraints. Lund, Sweden: Grahns Boktryckeri.Google Scholar
Welcomme, R. L. (1976). Some general and theoretical considerations on the fish yield of African rivers. Journal of Fish Biology, 8, 351–64.CrossRefGoogle Scholar
Welcomme, R. L. (1979). Fisheries Ecology of Floodplain Rivers. London: Longman.Google Scholar
Welcomme, R. L. (1986). Fish of the Niger system. In The Ecology of River Systems, eds. Davies, B. R. and Walker, K. F., pp. 25–48. Dordrecht, the Netherlands: Dr. W. Junk Publishers.Google Scholar
Weller, M. W. (1978). Management of freshwater marshes for wildlife. In Freshwater Wetlands: Ecological Processes and Management Potential, eds. Good, R. E., , D. F.Whigham, , and Simpson, R. L., pp. 267–84. New York: Academic Press.Google Scholar
Weller, M. W. (1994a). Freshwater Marshes: Ecology and Wildlife Management, 3rd edn. Minneapolis, MN: University of Minnesota Press.Google Scholar
Weller, M. W. (1994b). Bird–habitat relationships in a Texas estuarine marsh during summer. Wetlands, 14, 293–300.CrossRefGoogle Scholar
Weller, M. W. (1999). Wetland Birds: Habitat Resources and Conservation Implications. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Welty, J. C. (1982). The Life of Birds, 3rd edn. New York: Saunders College Publishing.Google Scholar
Werner, E. E. (1984). The mechanisms of species interactions and community organization in fish. In Ecological Communities: Conceptual Issues and the Evidence, eds. Strong, D. R., Jr., Simberloff, D., Abele, L. G., and Thistle, A. B., pp. 360–82. Princeton, NJ: Princeton University Press.Google Scholar
Werner, E. E. and Hall, D. J. (1976). Niche shifts in sunfishes: experimental evidence and significance. Science, 191, 404–6.CrossRefGoogle ScholarPubMed
Werner, E. E. and Hall, D. J. (1977). Competition and habitat shift in two sunfishes (Centrarchidae). Ecology, 58, 869–76.CrossRefGoogle Scholar
Werner, E. E. and Hall, D. J. (1979). Foraging efficiency and habitat switching in competing sunfishes. Ecology, 60, 256–64.CrossRefGoogle Scholar
Werner, E. E., Skelly, D. K., Relyea, R. A., and Yurewicz, K. L. (2007). Amphibian species richness across environmental gradients. Oikos, 116, 1697–712.CrossRefGoogle Scholar
Werner, K. J. and Zedler, J. B. (1997). Microtopographic heterogeneity and floristic diversity in experimental wetland communities. Journal of Ecology, 85, 71–82.Google Scholar
Western, D. (1975). Water availability and its influence on the structure and dynamics of a savannah large mammal community. African Wildlife Journal, 13, 265–86.CrossRefGoogle Scholar
Westhoff, V. and Maarel, E. (1973). The Braun–Blanquet approach. In Ordination and Classification of Communities, ed. Whittaker, R. H., pp. 617–726. The Hague, the Netherlands: Dr. W. Junk Publishers.CrossRefGoogle Scholar
Wetzel, R. G. (1975). Limnology. Philadelphia, PA: W. B. Saunders.Google Scholar
Wetzel, R. G. (1989). Wetland and littoral interfaces of lakes: productivity and nutrient regulation in the Lawrence Lake ecosystem. In Freshwater Wetlands and Wildlife, eds. Sharitz, R. R. and Gibbons, J. W., pp. 283–302. Proceedings of a Symposium held at Charleston, South Carolina, Mar 24–27, 1986. Washington, DC: U.S. Department of Energy.Google Scholar
Whalen, S. C. (2005). Biogeochemistry of methane exchange between natural wetlands and the atmosphere. Environmental Engineering Science, 22, 73–94.CrossRefGoogle Scholar
Wheeler, B. D. and Giller, K. E. (1982). Species richness of herbaceous fen vegetation in Broadland, Norfolk in relation to the quantity of above-ground plant material. Journal of Ecology, 70, 179–200.CrossRefGoogle Scholar
Wheeler, B. D. and Proctor, M. C. F. (2000). Ecological gradients, subdivisions and terminology of north-west European mires. Journal of Ecology, 88, 187–203.CrossRefGoogle Scholar
Wheeler, B. D. and Shaw, S. C. (1991). Above-ground crop mass and species richness of the principal types of herbaceous rich-fen vegetation of lowland England and Wales. Journal of Ecology, 79, 285–301.CrossRefGoogle Scholar
Whigham, , D. F., Dykyjova, D., and Hejny, S. (eds.) (1992). Wetlands of the World, Vol. 1. Dordrecht, the Netherlands: Kluwer.
White, P. S. (1979). Pattern, process and natural disturbance in vegetation. Botanical Review, 45, 229–99.CrossRefGoogle Scholar
White, P. S. (1994). Synthesis: vegetation pattern and process in the Everglades ecosystem. In Everglades: The Ecosystem and its Restoration, eds. Davis, S. and Ogden, J., pp. 445–60. DelRay Beach, FL: St. Lucie Press.Google Scholar
White, P. S., Wilds, S. P., and Thunhorst, G. A. (1998). Southeast. In Status and Trends of the Nation's Biological Resources, eds. Mac, M. J., Opler, P. A., Haecker, C. E. Puckett, and Doran, P. D., pp. 255–314. Reston, VA: U.S. Department of the Interior, U.S. Geological Survey.Google Scholar
White, T. C. R. (1993). The Inadequate Environment. Berlin, Germany: Springer-Verlag.CrossRefGoogle Scholar
Whiting, G. J. and Chanton, J. P. (1993). Primary production control of methane emission from wetlands. Nature, 364, 794–5.CrossRefGoogle Scholar
Whitney, D. M., Chalmers, A. G., Haines, E. B., Hanson, R. B., Pomeroy, L. R., and Sherr, B. (1981). The cycles of nitrogen and phosphorus. In The Ecology of a Salt Marsh, eds. Pomeroy, L. R. and Wiegert, R. G., pp. 161–78. New York: Springer-Verlag.Google Scholar
Whittaker, R. H. (1956). Vegetation of the Great Smoky Mountains. Ecological Monographs, 26, 1–80.CrossRefGoogle Scholar
Whittaker, R. H. (1962). Classification of natural communities. Botanical Review, 28, 1–160.CrossRefGoogle Scholar
Whittaker, R. H. (1967). Gradient analysis of vegetation. Biological Reviews, 42, 207–64.CrossRefGoogle ScholarPubMed
Whittaker, R. H. (1975). Communities and Ecosystems. New York: Macmillan.Google Scholar
Whittaker, R. H. and Likens, G. E. (1973). Carbon in the biota. In Carbon in the Biosphere, eds. Woodwell, G. M. and Peacan, E. R., pp. 281–302. Springfield, VA: National Technical Information Service.Google Scholar
Wickware, G. M. and Rubec, C. D. A. (1989). Ecoregions of Ontario, Ecological Land Classification Series No. 26. Ottawa, ON: Environment Canada, Sustainable Development Branch.Google Scholar
Wiegers, J. (1990). Forested wetlands in western Europe. In Forested Wetlands, eds. Lugo, A. E., Brinson, M., and Brown, S., pp. 407–36. Amsterdam, the Netherlands: Elsevier.Google Scholar
Wiegert, R. G. L., Pomeroy, R., and Wiebe, W. J. (1981). Ecology of salt marshes: an introduction. In The Ecology of a Salt Marsh, eds. Pomeroy, L. R. and Wiegert, R. G., pp. 3–20. New York: Springer-Verlag.Google Scholar
Wiens, J. A. (1965). Behavioral interactions of red-winged blackbirds and common grackles on a common breeding ground. The Auk, 82, 356–74.CrossRefGoogle Scholar
Wiens, J. A. (1983). Avian community ecology: an iconoclastic view. In Perspectives in Ornithology, essays presented for the centennial of the American Ornithologists' Union, eds. Brush, A. H. and Clark, G. A., Jr., pp. 355–403. Cambridge, UK: Cambridge University Press.CrossRefGoogle Scholar
Wikramanayake, E. D. (1990). Ecomorphology and biogeography of a tropical stream fish assemblage: evolution of assemblage structure. Ecology, 71, 1756–64.CrossRefGoogle Scholar
Wilbur, H. M. (1972). Competition, predation and the structure of the Ambystoma–Rana sylvatica community. Ecology, 53, 3–21.CrossRefGoogle Scholar
Wilbur, H. M. (1984). Complex life cycles and community organization in amphibians. In A New Ecology: Novel Approaches to Interactive Systems, eds. Price, P. W., Slobodchikoff, C. N., and Gaud, W. S., pp. 195–225. New York: John Wiley.Google Scholar
Wilcox, D. A. and Meeker, J. E. (1991). Disturbance effects on aquatic vegetation in regulated and unregulated lakes in northern Minnesota. Canadian Journal of Botany, 69, 1542–51.CrossRefGoogle Scholar
Wilcox, D. A. and Simonin, H. A. (1987). A chronosequence of aquatic macrophyte communities in dune ponds. Aquatic Botany, 28, 227–42.CrossRefGoogle Scholar
Wilcox, D. A. and Xie, Y. (2007). Predicting wetland plant responses to proposed water-level-regulation plans for Lake Ontario: GIS-based modeling. Journal of Great Lakes Research, 33, 751–73.CrossRefGoogle Scholar
Wilcox, D. A., Kowalski, K. P, Hoare, H. L., Carlson, M. L., and Morgan, H. N. (2008) Cattail invasion of sedge/grass meadows in Lake Ontario: photointerpretation analysis of sixteen wetlands over five decades. Journal of Great Lakes Research, 34, 301–23.CrossRefGoogle Scholar
,Wild Earth. (1992). The Wildlands Project, Special Issue. Richmond, VT: Wild Earth.Google Scholar
Williams, C. B. (1964). Patterns in the Balance of Nature. London: Academic Press.Google Scholar
Williams, M. (1989). The lumberman's assault on the southern forest, 1880–1920. In Americans and Their Forests: A Historical Geography, ed. Williams, M., pp. 238–88. Cambridge, UK: Cambridge University Press.Google Scholar
Williamson, G. B. (1990). Allelopathy, Koch's postulates and the neck riddle. In Perspectives on Plant Competition, eds. Grace, J. B. and Tilman, D., pp. 143–62. San Diego, CA: Academic Press.Google Scholar
Willis, A. J. (1963). Braunton Burrows: the effects on the vegetation of the addition of mineral nutrients to the dune soils. Journal of Ecology, 51, 353–74.CrossRefGoogle Scholar
Wilsey, B. J, Chabreck, R. H., and Linscombe, R. G. (1991). Variation in nutria diets in selected freshwater forested wetlands of Louisiana. Wetlands, 11, 263–78.CrossRefGoogle Scholar
Wilson, E. O. (1993). The Diversity of Life. New York: W.W. Norton.Google Scholar
Wilson, E. O. and Bossert, W. H. (1971). A Primer of Population Biology. Sunderland, MA: Sinauer Associates.Google Scholar
Wilson, J. A. (1972). Principles of Animal Physiology. New York: Macmillan.Google Scholar
Wilson, J. B., Wells, T. C. E., Trueman, I. C., Jones, G., Atkinson, M. D., Crawley, M. J., Dodds, M. E., and Silvertown, J. (1996). Are there assembly rules for plant species abundance? An investigation in relation to soil resources and successional trends. Journal of Ecology, 84, 527–38.CrossRefGoogle Scholar
Wilson, S. D. and Keddy, P. A. (1985). Plant zonation on a shoreline gradient: physiological response curves of component species. Journal of Ecology, 73, 851–60.CrossRefGoogle Scholar
Wilson, S. D. and Keddy, P. A. (1986a). Species competitive ability and position along a natural stress/disturbance gradient. Ecology, 67, 1236–42.CrossRefGoogle Scholar
Wilson, S. D. and Keddy, P. A. (1986b). Measuring diffuse competition along an environmental gradient: results from a shoreline plant community. The American Naturalist, 127, 862–9.CrossRefGoogle Scholar
Wilson, S. D. and Keddy, P. A. (1988). Species richness, survivorship, and biomass accumulation along an environmental gradient. Oikos, 53, 375–80.CrossRefGoogle Scholar
Wilson, S. D. and Keddy, P. A. (1991). Competition, survivorship and growth in macrophyte communities. Freshwater Biology, 25, 331–7.CrossRefGoogle Scholar
Winemiller, K. O. (1991). Ecomorphological diversification in lowland freshwater fish assemblages from five biotic regions. Ecological Monographs, 61, 343–65.CrossRefGoogle Scholar
Winter, T. C. and Rosenberry, D. O. (1995). The interaction of ground water with prairie pothole wetlands in the Cottonwood Lake area, east-central North Dakota, 1979–1990. Wetlands, 15, 193–211.CrossRefGoogle Scholar
Wisheu, I. C. (1998). How organisms partition habitats: different types of community organization can produce identical patterns. Oikos, 83, 246–58.CrossRefGoogle Scholar
Wisheu, I. C. and Keddy, P. A. (1989a). Species richness – standing crop relationships along four lakeshore gradients: constraints on the general model. Canadian Journal of Botany, 67, 1609–17.CrossRefGoogle Scholar
Wisheu, I. C. and Keddy, P. A. (1989b). The conservation and management of a threatened coastal plain plant community in eastern North America (Nova Scotia, Canada). Biological Conservation, 48, 229–38.CrossRefGoogle Scholar
Wisheu, I. C. and Keddy, P. A. (1991). Seed banks of a rare wetland plant community: distribution patterns and effects of human-induced disturbance. Journal of Vegetation Science, 2, 181–8.CrossRefGoogle Scholar
Wisheu, I. C. and Keddy, P. A. (1992). Competition and centrifugal organization of plant communities: theory and tests. Journal of Vegetation Science, 3, 147–56.CrossRefGoogle Scholar
Wisheu, I. C. and Keddy, P. A. (1996). Three competing models for predicting the size of species pools: a test using eastern North American wetlands. Oikos, 76, 253–8.CrossRefGoogle Scholar
Wisheu, I. C., Keddy, P. A., Moore, D. J., McCanny, S. J., and Gaudet, C. L. (1990). Effects of eutrophication on wetland vegetation. In Wetlands of the Great Lakes, eds. Kusler, J. and Smardon, R., pp. 112–21. Berne, NY: Association of State Wetland Managers.Google Scholar
Wium-Anderson, S. (1971). Photosynthetic uptake of free CO2 by the roots ofLobelia dortmanna. Plantarum, 25, 245–8.CrossRefGoogle Scholar
Wolff, W. J. (1993). Netherlands wetlands. Hydrobiologia, 265, 1–14.CrossRefGoogle Scholar
Woo, M., Rowsell, R. D., and Clark, R. G. (1993). Hydrological Classification of Canadian Prairie Wetlands and Prediction of Wetland Inundation in Response to Climatic Variability, Occasional Paper No. 79. Ottawa, ON: Canadian Wildlife Service.Google Scholar
Woodley, , S., Kay, J., and Francis, G. (eds.) (1993). Ecological Integrity and the Management of Ecosystems. Delray Beach, FL: St. Lucie Press.
Woodward, F. I. and Kelly, C. K. (1997). Plant functional types: towards a definition by environmental constraints. In Plant Functional Types, eds. Smith, T. M., Shugart, H. H., and Woodward, F. I., pp. 47–65. Cambridge, UK: Cambridge University Press.Google Scholar
Woodwell, G. M. and Whittaker, R. H. (1968). Effects of chronic gamma radiation on plant communities. Quarterly Review of Biology, 43, 42–55.CrossRefGoogle ScholarPubMed
Woodwell, G. M., Mackenzie, F. T., Houghton, R. A., Apps, A. J., Gorham, E., and Davidson, E. A. (1995).Will the warming speed the warming? In Biotic Feedbacks in the Global Climatic System, eds. Woodwell, G. M. and Mackenzie, F. T., pp. 393–411. New York: Oxford University Press.Google Scholar
Wootton, R. J. (1990). Biotic interaction. II. Competition and mutualism. In Ecology of Teleost Fishes, ed. Wootton, R. J., pp. 216–37. London: Chapman and Hall.CrossRefGoogle Scholar
,World Commission on Environment and Development. (1987). Our Common Future. Oxford, UK: Oxford University Press.Google Scholar
,World Conservation Monitoring Centre. (1992). Global Biodiversity: Status of the Earth's Living Resources. London: Chapman and Hall.Google Scholar
,World Resources Institute. (1992). World Resources 1992–1993. Oxford, UK: Oxford University Press.Google Scholar
,World Wildlife Fund (WWF). (1999). Evaluation of Wetlands and Floodplain Areas in the Danube River Basin: Final Report. Sofia, Bulgaria: WWF Danube–Carpathian Programme, and Rastatt, Germany: WWF Auen-Institut.
,World Wildlife Fund (WWF). (2003). Dikes bulldozed in Danube Delta, news release Oct 30, 2003. assets.panda.org/downloads/danube_delta_factsheet_en.pdf
Wright, H. E. and Bent, A. M. (1968). Vegetation bands around Dead Man Lake, Chuska Mountain, New Mexico. American Midland Naturalist, 79, 8–30.CrossRefGoogle Scholar
Wright, R. A. (2004). A Short History of Progress. Toronto, ON: Anansi Press.Google Scholar
Wu, Y., Rutchey, K., Wang, N., and Godin, J. (2006). The spatial pattern and dispersion of Lygodium microphyllum in the Everglades wetland ecosystem. Biological Invasions, 8, 1483–93.CrossRefGoogle Scholar
Yabe, K. (1993). Wetlands of Hokkaido. In Biodiversity and Ecology in the Northernmost Japan, eds. Higashi, S., Osawa, A., and Kanagawa, K., pp. 38–49. Hokkaido, Japan: Hokkaido University Press.Google Scholar
Yabe, K. and Numata, M. (1984). Ecological studies of the Mobawa–Yatsumi marsh: main physical and chemical factors controlling the marsh ecosystem. Japanese Journal of Ecology, 34, 173–86.Google Scholar
Yabe, K. and Onimaru, K. (1997). Key variables controlling the vegetation of a cool–temperate mire in northern Japan. Journal of Vegetation Science, 8, 29–36.CrossRefGoogle Scholar
Yang, S. L., Belkin, I. M., Belkina, A. I., Zhao, Q. Y., Zhu, J., and Ding, P. X. (2003). Delta response to decline in sediment supply from the Yangtze River: evidence of the recent four decades and expectations for the next half-century. Estuarine, Coastal and Shelf Science, 57, 689–99.CrossRefGoogle Scholar
Yodzis, P. (1986). Competition, mortality, and community structure. In Community Ecology, eds. Diamond, J. and Case, T. J., pp. 480–92. New York: Harper and Row.Google Scholar
Yodzis, P. (1989). Introduction to Theoretical Ecology. New York: Harper and Row.Google Scholar
Yu, Z., McAndrews, J. H., and Siddiqi, D. (1996). Influences of Holocene climate and water levels on vegetation dynamics of a lakeside wetland. Canadian Journal of Botany, 74, 1602–15.CrossRefGoogle Scholar
Zagwijn, W. H. (1986). Geologie van Nederland, Vol. 1, Nederland in het Holoceen. Haarlem, the Netherlands: Staatssuitgeverij, and The Hague: Rijks Geologische Dienst.Google Scholar
Zalidis, G. C., Mantzavelas, A. L., and Gourvelou, E. (1997). Environmental impacts on Greek wetlands. Wetlands, 17, 339–45.CrossRefGoogle Scholar
Zampella, R. A., Bunnell, J. F., Laidig, K. J., and Procopio, N. A. (2006). Using multiple indicators to evaluate the ecological integrity of a coastal plain stream system. Ecological Indicators, 6, 644–63.CrossRefGoogle Scholar
Zedler, J. B. (1988). Why it's so difficult to replace wetland functions. In Increasing our Wetland Resources, eds. Zelazny, J. and Feierabend, J. S.. Proceedings of a conference in Washington, DC, Oct 4–7, 1987. Reston, VA: National Wildlife Federation–Corporate Conservation Council.Google Scholar
Zedler, J. B. (1996). Ecological issues in wetland mitigation: an introduction to the forum. Ecological Applications, 6, 33–7.CrossRefGoogle Scholar
Zedler, J. B. and Beare, P. A. (1986). Temporal variability of salt marsh vegetation: the role of low-salinity gaps and environmental stress. In Estuarine Variability, ed. Wolfe, D. A., pp. 295–306. San Diego, CA: Academic Press.CrossRefGoogle Scholar
Zedler, J. B. and Kercher, S. (2004). Causes and consequences of invasive plants in wetlands: opportunities, opportunists, and outcomes. Critical Reviews in Plant Sciences, 23, 431–52.CrossRefGoogle Scholar
Zedler, J. B. and Kercher, S. (2005). Wetland resources: status, ecosystem services, degradation, and restorability. Annual Review of Environment and Resources, 30, 39–74.CrossRefGoogle Scholar
Zedler, J. B. and Onuf, C. P. (1984). Biological and physical filtering in arid-region estuaries: seasonality, extreme events, and effects of watershed modification. In The Estuary as a Filter, ed. Kennedy, V. S., pp. 415–32. New York: Academic Press.CrossRefGoogle Scholar
Zedler, J. B., Paling, E., and McComb, A. (1990). Differential responses to salinity help explain the replacement of native Juncus kraussii by Typha orientalis in Western Australian salt marshes. Australian Journal of Ecology, 15, 57–72.CrossRefGoogle Scholar
Zelazny, , J. and Feierabend, J. S. (eds.) (1988). Increasing our Wetland Resources. Proceedings of a conference in Washington, DC Oct 4–7, 1987. Reston, VA: National Wildlife Federation–Corporate Conservation Council.
Zhao, S. and Fang, J. (2004). Impact of impoldening and lake restoration on land-cover changes in Dongting Lake area, Central Yangtze. Amtio, 33, 311–15.Google ScholarPubMed
Zhulidov, A. V., Headley, J. V., Roberts, R. D., Nikanorov, A. M., and Ischenko, A. A. (1997). Atlas of Russian Wetlands, eds. Branned, M. J., translated by Flingeffman, Y.V. and Zhulidov, O.V.. Saskatoon, Sask.: Environment Canada, National Hydrology Research Institute.Google Scholar
Zobel, M. (1988). Autogenic succession in boreal mires: a review. Folia Geobotanica & Phytotaxonomica, 23, 417–45.CrossRefGoogle Scholar

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.

  • References
  • Paul A. Keddy
  • Book: Wetland Ecology
  • Online publication: 05 March 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9780511778179.017
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.

  • References
  • Paul A. Keddy
  • Book: Wetland Ecology
  • Online publication: 05 March 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9780511778179.017
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.

  • References
  • Paul A. Keddy
  • Book: Wetland Ecology
  • Online publication: 05 March 2013
  • Chapter DOI: https://doi.org/10.1017/CBO9780511778179.017
Available formats
×