Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-14T06:03:08.117Z Has data issue: false hasContentIssue false
This chapter is part of a book that is no longer available to purchase from Cambridge Core

References

Paul Keddy
Affiliation:
Southeastern Louisiana University
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
Plants and Vegetation
Origins, Processes, Consequences
, pp. 612 - 666
Publisher: Cambridge University Press
Print publication year: 2007

Access options

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

References

Aaviksoo, K., M. Ilomets, and M. Zobel. 1993. Dynamics of mire communities: a Markovian approach (Estonia). pp. 23–43. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 2. The Hague: SPB Academic Publishing.Google Scholar
Abraham, K. F. and C. J. Keddy. 2005. The Hudson Bay lowland. pp. 118–148. In Fraser, L. H. and Keddy, P. A. (eds.) The World's Largest Wetlands: Ecology and Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Abrahamson, W. G. and Gadgil, M.. 1973. Growth form and reproductive effort in goldenrods (Solidago, Compositae). The American Naturalist 107: 651–661.CrossRefGoogle Scholar
Ackerman, J. D. and Montalvo, A. M.. 1990. Short- and long-term limitations to fruit production in a tropical orchid. Ecology 71: 263–272.CrossRefGoogle Scholar
Adam, P. 1990. Saltmarsh Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Adamus, P. R. 1992. Choices in monitoring wetlands. pp. 571–592. In McKenzie, D. H., Hyatt, D. E., and McDonald, V. J. (eds.) Ecological Indicators. London: Elsevier Applied Science.Google Scholar
Aerts, R. 1996. Nutrient resorption from senescing leaves of perennials: are there general patterns. Journal of Ecology 84: 597–608.CrossRefGoogle Scholar
Agnew, A. D. Q. 1961. The ecology of Juncus effusus L. in North Wales. Journal of Ecology 49: 83–102.CrossRefGoogle Scholar
Allee, W. C. 1951. Cooperation Among Animals with Human Implications. New York: Schuman. (Revised edition of Social Life of Animals. 1938. New York: Norton.)Google Scholar
Allee, W. C., Emerson, A. E., Park, O., Park, T., and Schmidt, K. P.. 1949. Principles of Animal Ecology. Philadelphia: Saunders.Google Scholar
Allen, E. B. 1988. The Reconstruction of Disturbed Arid Ecosystems. Boulder: Westview Press.Google Scholar
Allen, E. B. and M. F. Allen. 1990. The mediation of competition by mycorrhizae in successional and patchy environments. pp. 367–389. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.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
Alvarez, W. 1998. T. rex and the Crater of Doom. New York: Vintage Books.Google Scholar
Alverson, W. S., Waller, D. M., and Solheim, S. J.. 1988. Forests to deer: edge effects in northern Wisconsin. Conservation Biology 2: 348–358.CrossRefGoogle Scholar
Anderson, R. C., Liberta, A. E., Dickman, L. A., and Katz, A. J.. 1983. Spatial variation in vesicular-arbuscular mycorrhizal spore density. Bulletin of the Torrey Botanical Club 110: 519–525.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: 111–117.CrossRefGoogle ScholarPubMed
Anderson, R. C., Fralish, J. S., and Baskin, J. M.. 1999. Savannas, Barrens, and Rock Outcrop Plant Communities of North America. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Antonovics, J. 1984. Genetic variation within populations. pp. 229–241. In Dirzo, R. and Sarukhán, J. (eds.) Perspectives on Plant Population Ecology. Sunderland: Sinauer.Google Scholar
Archer, S. 1989. Have southern Texas savannas been converted to woodlands in recent history?The American Naturalist 134: 545–561.CrossRefGoogle Scholar
Archibold, O. W. 1995. Ecology of World Vegetation. London: Chapman and Hall.CrossRefGoogle Scholar
Arthur, W. 1982. The evolutionary consequences of interspecific competition. Advances in Ecological Research 12: 127–187.CrossRefGoogle Scholar
Arthur, W. 1987. The Niche in Competition and Evolution. Chichester: Wiley.Google Scholar
Ashton, P. S. 1988. Dipterocarp biology as a window to the understanding of tropical forest structure. Annual Review of Ecology and Systematics 19: 347–370.CrossRefGoogle Scholar
Ashton, P. S., Givnish, T. J., and Appanah, S.. 1988. Staggered flowering in the Dipterocarpaceae: new insights into floral induction and the evolution of mast fruiting in the seasonal tropics. The American Naturalist 132: 44–66.CrossRefGoogle Scholar
Atwood, E. L. 1950. Life history studies of the nutria, or coypu, in coastal Louisiana. Journal of Wildlife Management 14: 249–265.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: 941–952.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. 1968. An ordination study of a chalk grassland community. Journal of Ecology 56: 739–757.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: 559–570.CrossRefGoogle Scholar
Austin, M. P. and Austin, B. O.. 1980. Behaviour of experimental plant communities along a nutrient gradient. Journal of Ecology 68: 891–918.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. Australian Journal of Ecology 21: 154–164.CrossRefGoogle Scholar
Austin, M. P., J. G. Pausas, and I. R. Noble. 1997. Modelling environmental and temporal niches of eucalypts. pp. 129–150. In Williams, J. E. and Woinarski, J. C. Z. (eds.) Eucalypt Ecology: Individuals to Ecosystems. Cambridge: Cambridge University Press.Google Scholar
Axelrod, D. I. 1970. Mesozoic paleogeography and early angiosperm history. The Botanical Review 36: 277–319.CrossRefGoogle Scholar
Axelrod, D. I. and P. H. Raven. 1972. Evolutionary biogeography viewed from plate tectonic theory. pp. 218–236. In Behnke, J. A. (ed.) Challenging Biological Problems: Directions Toward Their Solution. Oxford: Oxford University Press.Google Scholar
Bacon, P. R. 1978. Flora and Fauna of the Caribbean. Trinidad: Key Caribbean Publications.Google Scholar
Bailes, K. E. 1990. Science and Russian Culture in an Age of Revolutions. V. I. Vernadsky and his Scientific School, 1863–1945. Bloomington, IN: Indiana University Press.Google Scholar
Baker, H. 1937. Alluvial meadows: a comparative study of grazed and mown meadows. Journal of Ecology 25: 408–420.CrossRefGoogle Scholar
Baker-Brosh, K. and Peet, R. K.. 1997. The ecological significance of lobed and toothed leaves in temperate forest trees. Ecology 78: 1250–1255.Google Scholar
Bakker, R. T. 1978. Dinosaur feeding behaviour and the origin of flowering plants. Nature 274: 661–663.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: 113–120.CrossRefGoogle Scholar
Baldwin, W. K. W. 1958. Plants of the Clay Belt of Northern Ontario and Quebec. National Museum of Canada, Bulletin No. 156.
Ball, P. J. and T. D. Nudds. 1989. Mallard habitat selection: an experiment and implications for management. pp. 659–671. In R. R. Sharitz and J. W. Gibbons (eds.) Freshwater Wetlands and Wildlife. U.S. Department of Energy. Proceedings of a symposium held at Charleston, South Carolina, March 24–27, 1986.
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: 109–117.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
Barnett, V. 1994. Statistics and the long-term experiments: past achievements and future challenges. pp. 165–183. In Leigh, R. A. and Johnston, A. E. (eds.) Long-term Experiments in Agricultural and Ecological Sciences. Proceedings of a conference to celebrate the 150th anniversary of Rothamsted Experimental Station, held at Rothamsted, July 14–17, 1993. Wallingford: CAB International.Google Scholar
Barrett, S. C. H. 2002. The evolution of plant sexual diversity. Nature Reviews Genetics 3: 274–284.CrossRefGoogle ScholarPubMed
Barry, J. M. 1997. Rising Tide. The Great Mississippi Flood of 1927 and How it Changed America. New York: Simon and Schuster.Google Scholar
Barth, F. G. 1985. Insects and Flowers: The Biology of a Partnership. Princeton: Princeton University Press. Translated from 1982 German edition by M. A. Biederman-Thorson.Google Scholar
Barthlott, W., Porembski, S., Fischer, E., and Gemmel, B.. 1998. First protozoa-trapping plant found. Nature 392: 447.CrossRefGoogle Scholar
Baskin, J. M. and Baskin, C. C.. 1985. A floristic study of a cedar glade in Blue Licks Battlefield State Park, Kentucky. Castanea 50: 19–25.Google Scholar
Bauer, C. R., Kellogg, C. H., Bridgham, S. D., and Lamberti, G. A.. 2003. Mycorrhizal colonization across hydrological gradients in restored and reference freshwater wetlands. Wetlands 23: 961–968.CrossRefGoogle Scholar
Baylis, G. T. S. 1980. Mycorrhizas and the spread of beech. New Zealand Journal of Ecology 3: 151–153.Google Scholar
Beard, J. S. 1944. Climax vegetation in tropical America. Ecology 25: 127–158.CrossRefGoogle Scholar
Beard, J. S. 1949. The Natural Vegetation of the Windward and Leeward Islands. Oxford: Clarendon Press.Google Scholar
Beard, J. S. 1973. The physiognomic approach. pp. 355–386. In Whittaker, R. H. (ed.) Ordination and Classification of Communities. Part V. The Hague: W. Junk.CrossRefGoogle Scholar
Beattie, A. J. and Culver, D. C.. 1981. The guild of myrmecochores in the herbaceous flora of West Virginia forests. Ecology 62: 107–115.CrossRefGoogle Scholar
Beeftink, W. G. 1977. The coastal salt marshes of western and northern Europe: an ecological and phytosociological approach. pp. 109–155. In Chapman, V. J. (ed.) Ecosystems of the World 1: Wet Coastal Ecosystems. Amsterdam: Elsevier Scientific Publishing Company.Google Scholar
Bégin, 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. Geographie Physique et Quaternaire 43: 355–366.CrossRefGoogle Scholar
Begon, M. and Mortimer, M.. 1981. Population Ecology: A Unified Study of Animals and Plants. Oxford: Blackwell.Google Scholar
Belcher, J., Keddy, P. A., and Catling, P. M. C.. 1992. Alvar vegetation in Canada: a multivariate description at two scales. Canadian Journal of Botany 70: 1279–1291.CrossRefGoogle Scholar
Belcher, J. W., Keddy, P. A., and Twolan-Strutt, L.. 1995. Root and shoot competition intensity along a soil depth gradient. Journal of Ecology 83: 673–682.CrossRefGoogle Scholar
Bell, A. D. 1984. Dynamic morphology: a contribution to plant population ecology. pp. 48–65. In Dirzo, R. and Sarukhán, J. (eds.) Perspectives on Plant Population Ecology. Sunderland: Sinauer.Google Scholar
Bell, R. A. 1993. Cryptoendolithic algae of hot semiarid lands and deserts. Journal of Phycology 29: 133–139.CrossRefGoogle Scholar
Belsky, A. J. 1992. Effects of grazing, competition, disturbance and fire on species composition and diversity in grassland communities. Journal of Vegetation Science 3: 187–200.CrossRefGoogle Scholar
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
Benecke, P. and R. Mayer. 1971. Aspects of soil water behavior as related to beech and spruce stands: some results of water balance investigations. pp. 153–168. In Ellenburg, H. (ed.) Integrated Experimental Ecology: Methods and Results of Ecosystem Research in the German Solling Project, Vol. 2. Ecological Studies: Analysis and Synthesis. New York: Springer.CrossRefGoogle Scholar
Benson, L. 1950. The Cacti of Arizona, 2nd edn. Tucson: University of Arizona Press.Google Scholar
Benson, L. 1959. Plant Classification. Lexington: D.C. Heath and Company.Google Scholar
Benzing, D. H. 1990. Vascular Epiphytes: General Biology and Related Biota. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Berbee, M. L. and Taylor, J. W.. 1993. Dating the evolutionary radiations of the true fungi. Canadian Journal of Botany 71: 1114–1127.CrossRefGoogle Scholar
Berenbaum, M. R. 1991. Coumarins. pp. 221–249. In Rosenthal, G. A. and Berenbaum, M. R. (eds.) Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego: Academic Press.Google Scholar
Berg, R. Y. 1975. Myrmecochorous plants in Australia and their dispersal by ants. Australian Journal of Botany 23: 475–508.CrossRefGoogle Scholar
Bernard, H. A. and R. J. Leblanc. 1965. Résumé of the quaternary geology of the northwestern Gulf of Mexico province. pp. 137–185. In Wright, H. E. and Frey, D. G. (eds.) The Quaternary of the United States. Princeton: Princeton University Press.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 XIV: 519–545.Google Scholar
Bertness, M. D. and Hacker, S. D.. 1994. Physical stress and positive associations among marsh plants. The American Naturalist 144: 363–372.CrossRefGoogle Scholar
Bertness, M. D. and Yeh, S. M.. 1994. Cooperative and competitive interactions in the recruitment of marsh elders. Ecology 75: 2416–2429.CrossRefGoogle Scholar
Bessey, C. E. 1915. The phylogenetic taxonomy of flowering plants. Annals of the Missouri Botanical Garden 2: 109–164.CrossRefGoogle Scholar
Bierzychudek, P. 1980. The demographic consequences of sexuality and apomixis in Antennaria. pp 293–307. In Kawano, S. (ed.) Biological Approaches and Evolutionary Trends in Plants. London: Academic Press.Google Scholar
Billings, W. D. and Mooney, H. A.. 1968. The ecology of arctic and alpine plants. Biological Reviews 43: 481–529.CrossRefGoogle Scholar
Binford, M. W., Brenner, M., Whitmore, T. J., Higuera-Gundy, A., Deevey, E. S., and Leyden, B.. 1987. Ecosystems, paleoecology and human disturbance in subtropical and tropical America. Quaternary Science Reviews 6: 115–128.CrossRefGoogle Scholar
Björkman, E. 1960. Monotropa hypopitys L. an epiparasite on tree roots. Physiologia Plantarum 13: 308–327.CrossRefGoogle Scholar
Black, D. (ed.) 1979. Carl Linnaeus: Travels. Nature Classics. New York: Charles Scribner's Sons.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: 1095–1107.CrossRefGoogle Scholar
Blizard, D. 1993. The Normandy Landings D-Day: The Invasion of Europe 6 June 1944. London: Reed International Books.Google 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
Bohlen, P. J., Scheu, S., Hale, C. M., McLean, M. A., Migge, S., Groffman, P. M., and Parkinson, D.. 2004. Invasive earthworms as agents of change in north temperate forests. Frontiers in Ecology and the Environment 8: 427–435.CrossRefGoogle Scholar
Bolan, N. S. 1991. A critical review on the role of mychorrhizal fungi in the uptake of phosphorus by plants. Plant and Soil 134: 189–207.CrossRefGoogle Scholar
Bond, W. J. 1997. Functional types for predicting changes in biodiversity: a case study in Cape fynbos. pp. 174–194. In Smith, T. M., Shugart, H. H., and Woodward, F. I. (eds.) Plant Functional Types. Cambridge: Cambridge University Press.Google Scholar
Boot, R. G. A. 1989. The significance of size and morphology of root systems for nutrient acquisition and competition. pp. 299–311. In Lambert, H.et al. (eds.) Causes and Consequences of Variation in Growth Rate and Productivity of Higher Plants. The Hague: SPB Academic Publishing.Google Scholar
Booth, B. and D. W. Larson. 1999. Impact of history, language, and choice of system on the study of assembly rules. pp. 206–229. In Weiher, E. and Keddy, P. (eds.) Ecological Assembly Rules: Perspectives, Advances, Retreats. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Borhidi, A. 1992. The serpentine flora and vegetation of Cuba. pp. 83–95. In Baker, A. J. M., Proctor, J., and Reeves, R. D. (eds.) The Vegetation of Ultramafic (Serpentine) Soils. Andover: Intercept.Google Scholar
Bormann, B. T. and Sidle, R. C.. 1990. Changes in productivity and distribution of nutrients in a chronosequence at Glacier Bay National Park, Alaska. Journal of Ecology 78: 561–578.CrossRefGoogle Scholar
Bormann, F. H. and Likens, G. E.. 1981. Pattern and Process in a Forested Ecosystem. Second corrected printing. 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: 259–270.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: 615–622.CrossRefGoogle ScholarPubMed
Botkin, D. B. 1977. Life and death in a forest: the computer as an aid to understanding. pp. 213–33. In Hall, A. S. and Day, J. W. (eds.) Ecosystem Modelling in Theory and Practice. New York: John Wiley and Sons.Google Scholar
Botkin, D. B. 1990. Discordant Harmonies: A New Ecology for the Twenty-first Century. New York: Oxford University Press.Google Scholar
Botkin, D. B. 1993. Forest Dynamics. Oxford: Oxford University Press.Google Scholar
Boucher, D. H. 1985a. The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Boucher, D. H. 1985b. The idea of mutualism, past and future. pp. 1–28. In Boucher, D. H.. The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Boucher, D. H., James, S., and Keeler, K. H.. 1982. The ecology of mutualism. Annual Review of Ecology and Systematics 13: 315–347.CrossRefGoogle Scholar
Boutin, C. and Keddy, P. A.. 1993. A functional classification of wetland plants. Journal of Vegetation Science 4: 591–600.CrossRefGoogle Scholar
Bowers, M. D. 1991. Iridoid glycosides. pp. 297–325. In Rosenthal, G. A. and Berenbaum, M. R. (eds.) Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego: Academic Press.Google Scholar
Boyd, C. E. 1978. Chemical composition of wetland plants. pp. 155–168. In Good, R. E., Whigham, D. F., and Simpson, R. L. (eds.) Freshwater Wetlands: Ecological Processes and Management Potential. 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: 443–452.Google Scholar
Brackenridge, J. B. and Rosenberg, R. M.. 1970. The Principles of Physics and Chemistry. New York: McGraw-Hill.Google Scholar
Braun, E. L. 1950. The Deciduous Forest of Eastern North America. New York: Hafner.Google Scholar
Brewer, J. S. 1998. Effects of competition and litter on a carnivorous plant, Drosera cappilaris (Droseraceae). American Journal of Botany 85 N 11: 1592–1596.CrossRefGoogle Scholar
Brewer, J. S. 1999. Effects of fire, competition and soil disturbances on regeneration of a carnivorous plant (Drosera capillaris). The American Midland Naturalist 141: 28–42.CrossRefGoogle Scholar
Bridges, E. M., Batjes, N. H., and Nachtergaele, F. O. (eds.) 1998. World Reference Base for Soil Resources: Atlas. Leuven, Belgium: ACCO.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 WRP-DE-4. U.S. Army Corps of Engineers, Washington, D.C.Google Scholar
Brooks, R. R., R. D. Reeves, and A. J. M. Baker. 1992. The serpentine vegetation of Goiás State, Brazil. pp. 67–81. In Baker, A. J. M., Proctor, J., and Reeves, R. D. (eds.) The Vegetation of Ultramafic (Serpentine) Soils. Andover: Intercept.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: 151–158.CrossRefGoogle Scholar
Brown, J. H., D. W. Davidson, J. C. Munger, and R. S. Inouye. 1986. Experimental community ecology: the desert granivore system. pp. 41–61. In Diamond, J. and Case, T. J. (eds.) Community Ecology. New York: Harper and Row.Google Scholar
Browne, J. 1995. Charles Darwin: Voyaging. Princeton: Princeton University Press.Google Scholar
Bryant, D., Nielsen, D., and Tangley, L.. 1997. The Last Frontier Forests: Ecosystems and Economies on the Edge. Washington: World Resources Institute.Google Scholar
Burch, W. Jr. 1999. Daydreams and Nightmares – A Sociological Essay on the American Environment. Madison: Social Ecology Press.Google Scholar
Burdon, J. J. 1982. The effect of fungal pathogens on plant communities. pp. 99–112. In Newman, E. I. (ed.) The Plant Community as a Working Mechanism. Oxford: Blackwell.Google Scholar
Burger, J. C. and Louda, S. V.. 1995. Interaction of diffuse competition and insect herbivory in limiting brittle prickly pear cactus, Opuntia fragilis (Cactaceae). American Journal of Botany 82: 1558–1566.CrossRefGoogle Scholar
Burgess, R. L. and Sharpe, D. M. (eds.) 1981. Forest Island Dynamics in Man-dominated Landscapes. New York: Springer-Verlag.CrossRefGoogle Scholar
Buss, L. W. 1988. The Evolution of Individuality. Princeton: Princeton University Press.CrossRefGoogle Scholar
Cairns, J. (ed.) 1980. The Recovery Process in Damaged Ecosystems. Ann Arbor: Ann Arbor Science.Google Scholar
Cairns, J. (ed.) 1988. Rehabilitating Damaged Ecosystems. Vol. 1 and 2. Boca Raton: CRC Press.Google Scholar
Cairns, J. 1989. Restoring damaged ecosystems: is predisturbance condition a viable option?The Environmental Professional 11: 152–159.Google Scholar
Cairns-Smith, A. G. 1985. Seven Clues to the Origin of Life: A Scientific Detective Story. Canto edition 1990. Cambridge: Cambridge University Press.Google Scholar
Callaway, R. M. and King, L.. 1996. Temperature-driven variation in substrate oxygenation and the balance of competition and facilitation. Ecology 77: 1189–1195.CrossRefGoogle Scholar
Campbell, B. D., Grime, J. P., and Mackey, J. M. L.. 1991. A trade-off between scale and precision in resource foraging. Oecologia 87: 532–538.CrossRefGoogle ScholarPubMed
Campbell, B. D., Grime, J. P., and Mackey, J. M. L.. 1992. Shoot thrust and its role in plant competition. Journal of Ecology 80: 633–641.CrossRefGoogle Scholar
Canfield, R. H. 1948. Perennial grass composition as an indicator of condition of southwestern mixed grass ranges. Ecology 29: 190–204.CrossRefGoogle Scholar
Caputa, J. 1948. Untersuchungen über die Entwicklung einiger Gräser und Kleearten in Reinsaat und Mischung. Landwirtschaftliches Jahrbuch der Schweiz 62: 848–975.Google Scholar
Carleton, T. J. and MacLellan, P.. 1994. Woody vegetation responses to fire versus clear-cutting logging: a comparative survey in the central Canadian boreal forest. Ecoscience 1: 141–152.CrossRefGoogle Scholar
Carpenter, S. R., Chisholm, S. W., Krebs, C. J., Schindler, D. W., and Wright, R. F.. 1995. Ecosystem experiments. Science 269: 324–327.CrossRefGoogle ScholarPubMed
Carroll, G. 1988. Fungal endophytes in stems and leaves: from latent pathogen to mutualistic symboint. Ecology 69: 2–9.CrossRefGoogle Scholar
Carson, R. 1962. Silent Spring. Boston: Houghton Mifflin.Google Scholar
Carson, W. P. and Pickett, S. T. A.. 1990. Role of resources and disturbance in the organization of an old-field plant community. Ecology 71: 226–238.CrossRefGoogle Scholar
Catling, P. M. and Brownell, V. R.. 1995. A review of the alvars of the Great Lakes region: distribution, floristic composition, biogeography and protection. The Canadian Field-Naturalist 109: 143–171.Google Scholar
Catling, P. M. and Brownell, V. R.. 1998. Importance of fire in alvar ecosystems – evidence from the Burnt Lands, Eastern Ontario. The Canadian Field-Naturalist 112: 661–667.Google Scholar
Catling, P. M., Cruise, J. E., McIntosh, K. L., and McKay, S. M.. 1975. Alvar vegetation in southern Ontario. Ontario Field Biologist 29: 1–25.Google Scholar
Cavers, P. B. 1983. Seed demography. Canadian Journal of Botany 61: 3578–3590.CrossRefGoogle Scholar
Chaneton, E. J. and Facelli, J. M.. 1991. Disturbance effects on plant community diversity: spatial scales and dominance hierarchies. Vegetatio 93: 143–156.CrossRefGoogle Scholar
Chapin, F. S. III 1980. The mineral nutrition of wild plants. Annual Review of Ecology and Systematics 11: 233–260.CrossRefGoogle Scholar
Chapin, F. S. III, Vitousek, P. M., and Cleve, K.. 1986. The nature of nutrient limitation in plant communities. The American Naturalist 127: 48–58.CrossRefGoogle Scholar
Chapman, V. J. 1940. The functions of the pneumatophores of Avicennia nitida Jacq. Proceedings of the Linnean Society of London 152: 228–233.CrossRefGoogle Scholar
Charron, D. and Gagnon, D.. 1991. The demography of northern populations of Panax quinquefolium (American ginseng). Journal of Ecology 79: 431–445.CrossRefGoogle Scholar
Cheplick, G. P. 1992. Sibling competition in plants. Journal of Ecology 80: 567–575.CrossRefGoogle Scholar
Christensen, N. L., R. B. Burchell, A. Liggett, and E. L. Simms. 1981. The structure and development of pocosin vegetation. pp. 43–61. In Richardson, C. J. (ed.) Pocosin Wetlands: An Integrated Analysis of Coastal Plain Freshwater Bogs in North Carolina. Stroudsburg, Pennsylvania: Hutchinson Ross Publishing Company.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: 665–691.CrossRefGoogle Scholar
Clarke, 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: 213–234.CrossRefGoogle Scholar
Clay, K. 1990. The impact of parasitic and mutualistic fungi on competitive interactions among plants. pp. 391–412. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Clements, F. E. 1916. Plant Succession: An Analysis of the Development of Vegetation. Pub. 242. Washington, DC: Carnegie Institute.CrossRefGoogle Scholar
Clements, F. E. 1933. Competition in plant societies. In News Service Bulletin. Washington: Carnegie Institution of Washington, April 2, 1933.Google Scholar
Clements, F. E. 1935. Experimental ecology in the public service. Ecology 16: 324–363.CrossRefGoogle Scholar
Clements, F. E. 1936. Nature and structure of climax. Journal of Ecology 24: 254–282.CrossRefGoogle Scholar
Clements, F. E., Weaver, J. E., and Hanson, H. C.. 1929. Plant Competition. Washington, D.C.: Carnegie Institution of Washington.Google Scholar
Cloud, P. 1976. Beginnings of biospheric evolution and their biogeochemical consequences. Paleobiology 2: 351–387.CrossRefGoogle Scholar
Cloudsley-Thompson, J. L. 1996. Biotic Interactions in Arid Lands. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Clymo, R. S. and Duckett, J. G.. 1986. Regeneration of Sphagnum. New Phytologist 102: 589–614.CrossRefGoogle Scholar
Clymo, R. S. and P. M. Hayward. 1982. The ecology of Sphagnum. pp. 229–289. In Smith, A. J. E. (ed.) Bryophyte Ecology. London: Chapman and Hall.CrossRefGoogle Scholar
Cody, M. L. 1993. Do cholla cacti (Opuntia spp., Subgenus Cylindropuntia) use or need nurse plants in the Mojave Desert?Journal of Arid Environments 24: 1–16.CrossRefGoogle Scholar
Coe, M. J., D. L. Dilcher, J. O. Farlow, D. M. Jarzen, and D. A. Russel. 1987. Dinosaurs and land plants. pp. 225–258. In Friis, E., Chaloner, W. G., and Crane, P. R. (eds.) The Origins of Angiosperms and Their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Cole, L. C. 1949. The measurement of interspecific association. Ecology 30: 411–424.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
Coleman, R. G. and C. Jove. 1992. Geological origin of serpentines. pp. 1–17. In Baker, A. J. M., Proctor, J., and Reeves, R. D. (eds.) The Vegetation of Ultramafic (Serpentine) Soils. Andover: Intercept.Google Scholar
Coley, P. D. 1983. Herbivory and defensive characteristics of tree species in a lowland tropical forest. Ecological Monographs 53: 209–233.CrossRefGoogle Scholar
Colinvaux, P. 1978. Why Big Fierce Animals are Rare: An Ecologist's Perspective. Princeton: Princeton University Press.Google Scholar
Colinvaux, P. 1986. Ecology. Toronto: Wiley and Sons.Google Scholar
Colinvaux, P. 1993. Ecology 2. New York: Wiley and Sons.Google Scholar
Colinvaux, P. A., Oliveira, P. E., Moreno, J. E., Miller, M. C., and Bush, M. B.. 1996. A long pollen record from lowland Amazonia: forest and cooling in glacial times. Science 274: 85–88.CrossRefGoogle Scholar
Colinvaux, P. A., Oliveira, P. E., and Bush, M. B.. 2000. Amazonian and Neotropical plant communities on glacial time-scales: the failure of the aridity and refuge hypotheses. Quaternary Science Reviews 19: 141–169.CrossRefGoogle Scholar
Colinvaux, P.A., Irion, G., Räsänen, M. E., Bush, M. B., and Mello, J. A. S. Nunes 2001. A paradigm to be discarded: geological and paleoecological data falsify the Haffer & Prance refuge hypothesis of Amazonian speciation. Amazoniana 16: 609–646.Google Scholar
Collinson, M. E. and J. J. Hooker. 1987. Vegetational and mammalian faunal changes in the Early Tertiary of southern England. pp. 259–304. In Friis, E., Chaloner, W. G., and Crane, P. R. (eds.) The Origins of Angiosperms and Their Biological Consequences. Cambridge: Cambridge 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
Connell, J. H. 1978. Diversity in tropical rain forests and coral reefs. Science 199: 1302–1310.CrossRefGoogle ScholarPubMed
Connell, J. H. 1990. Apparent versus “real” competition in plants. pp. 9–26. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Connell, J. H. and Slatyer, R. O.. 1977. Mechanisms of succession in natural communities and their role in community stability and organization. The American Naturalist 111: 1119–1144.CrossRefGoogle Scholar
Conner, W. H. and M. A. Buford. 1998. Southern deepwater swamps. pp. 261–287. In Messina, M. G. and Conner, W. H. (eds.) Southern Forested Wetlands. Ecology and Management. Boca Raton: Lewis Publishers.Google Scholar
Connolly, J. 1986. On difficulties with replacement-series methodology in mixture experiments. Journal of Applied Ecology 23: 125–137.CrossRefGoogle Scholar
Conservation International. 2006. Biodiversity Hotspots. Tropical Andes. (www.biodiversityhotspots.org/xp/Hotspots/andes/) accessed 24 July 2006.
Corfield, T. F. 1973. Elephant mortality in Tsavo National Park, Kenya. East African Wildlife Journal 11: 339–368.CrossRefGoogle Scholar
Cowling, R. M. 1990. Diversity components in a species-rich area of the Cape Floristic Region. Journal of Vegetation Science 1: 699–710.CrossRefGoogle Scholar
Cowling, R. M. and Samways, M. J.. 1995. Predicting global patterns of endemic plant species richness. Biodiversity Letters 2: 127–131.CrossRefGoogle Scholar
Cowling, R. M., Rundel, P. W., Lamont, B. B., Arroyo, M. K., and Arianoutsou, M.. 1996. Plant diversity in mediterranean-climate regions. Trends in Ecology and Evolution 11: 362–366.CrossRefGoogle ScholarPubMed
Craighead, F. C. Sr. 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
Crandell, D. R. and H. H. Waldron. 1969. Volcanic hazards in the Cascade Range. pp. 5–18. In R. Olson and M. Wallace (eds.) Geologic Hazards and Public Problems. Conference Proceedings. U.S. Government Printing Office.
Crawford, R. M. M. 1982. Physiological response to flooding. pp. 453–477. In Lange, O. L., Nobel, P. S., Osmond, C. B., and Ziegler, H. (eds.) Physiological Plant Ecology II. Encyclopedia of Plant Physiology. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Crawley, M. J. 1983. Herbivory: The Dynamics of Animal/Plant Interactions. Oxford: Blackwell.Google Scholar
Crawley, M. J. and Harral, J. E.. 2001. Scale dependence in plant biodiversity. Science 291: 864–868.CrossRefGoogle ScholarPubMed
Crepet, W. L. and E. M. Friis. 1987. The evolution of insect pollination in angiosperms. pp. 181–201. In Friis, E., Chaloner, W. G., and Crane, P. R. (eds.) The Origins of Angiosperms and Their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Crocker, R. L. and Major, J.. 1955. Soil development in relation to vegetation and surface age at Glacier Bay, Alaska. Journal of Ecology 43: 427–448.CrossRefGoogle Scholar
Cronquist, A. 1991. Asterales. pp. 721–726. In Angiosperms: The Flowering Plants. pp. 596–765, Vol. 13. The New Encyclopaedia Britannica, 15th edn. Chicago: The University of Chicago.Google Scholar
Cronquist, A. 1993. A commentary on the general system of classification of flowering plants. pp. 272–293. In Flora of North America Editorial Committee. Flora of North America, Vol. 1. Introduction. New York: Oxford University Press.Google Scholar
Cyr, H. and Pace, M. L.. 1993. Magnitude and patterns of herbivory in aquatic and terrestrial ecosystems. Nature 361: 148–150.CrossRefGoogle Scholar
Dacey, J. W. H. 1980. Internal winds in water lilies: an adaptation for life in anaerobic sediments. Science 210: 1017–1019.CrossRefGoogle ScholarPubMed
Dacey, J. W. H. 1981. Pressurized ventilation in the yellow water lily. Ecology 62: 1137–1147.CrossRefGoogle Scholar
Dafni, A. 1992. Pollination Ecology: A Practical Approach. Oxford: Oxford University Press.Google Scholar
Dale, M. 1999. Spatial Pattern Analysis in Plant Ecology. Cambridge: Cambridge University Press.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
Darwin, C. 1871. The descent of Man and selection in relation to sex. In Adler, M. J. (ed.) Great Books of the Western World, 2nd edn, Vol. 49. Chicago: Encyclopaedia Britannica.Google Scholar
Darwin, C. 1888. Insectivorous Plants. 2nd edn. London: John Murray. Revised by F. Darwin.CrossRefGoogle Scholar
Darwin, C. R. 1881. The Formation of Vegetable Mould Through the Action of Worms, with Observations on Their Habits. London: Murray.CrossRefGoogle Scholar
Darwin, F. (ed.) 1950. Charles Darwin's Autobiography: With his Notes and Letters Depicting the Growth of the Origin of Species. New York: Henry Schuman.Google Scholar
Daubenmire, R. 1978. Plant Geography: With Special Reference to North America. Physiological Ecology. New York: Academic Press.Google Scholar
Davis, D. W. 2000. Historical perspective on crevasses, levees, and the Mississippi River. In Colten, C. E. (ed.) Transforming New Orleans and its Environs, Centuries of Change. Pittsburgh: University of Pittsburgh Press.Google Scholar
Davies, B. R. and Walker, K. F.. 1986. The Ecology of River Systems. Dordrecht: W. Junk Publishers.CrossRefGoogle Scholar
Davis, M. B. 1976. Pleistocene biogeography of temperate deciduous forests. pp. 13–26. In West, R. C. and Haag, W. G. (eds.) Geoscience and Man, Vol. 13. Ecology of the Pleistocene, a Symposium. Baton Rouge: School of Geoscience, Louisiana State University.Google Scholar
Davis, S. and Ogden, J. (eds.) 1994. Everglades: The Ecosystem and its Restoration. Del Ray Beach: St. Lucie Press.Google Scholar
Dawkins, R. 1976. The Selfish Gene. Oxford: Oxford University Press.Google Scholar
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: 1044–1054.CrossRefGoogle Scholar
Day, W. 1984. Genesis on Planet Earth, 2nd edn. New Haven: Yale University Press.Google Scholar
Dayton, P. K. 1979. Ecology: a science and a religion. pp. 3–18. In Livingston, R. J. (ed.) Ecological Processes in Coastal and Marine Systems. New York: Plenum Press.CrossRefGoogle Scholar
Calesta, D. S. 1994. Effect of white-tailed deer on songbirds within managed forests in Pennsylvania. Journal of Wildlife Management 58: 711–718.CrossRefGoogle Scholar
Deckers, J. A., Nachtergaele, F. O., and Spaargaren, O. C. (eds.) 1998. World Reference Base for Soil Resources: Introduction. Leuven, Belgium: ACCO.Google Scholar
Duve, C. 1991. Blueprint for a Cell: The Nature and Origin of Life. Burlington: Neil Patterson.Google Scholar
Groot, R. S. 1992. Functions of Nature. Groningen: Wolters-Noordhoff.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. 1983. Competition as a control mechanism in subalpine meadows. American Journal of Botany 70: 232–245.CrossRefGoogle Scholar
del Moral, R. and Bliss, L. C.. 1993. Mechanisms of primary succession: insights resulting from the eruption of Mount St. Helens. Advances in Ecological Research 24: 1–66.CrossRefGoogle Scholar
del Moral, R. and S. Y. Grishin. 1999. Volcanic disturbances and ecosystem recovery. pp. 137–160. In Walker, L. R. (ed.) Ecosystems of Disturbed Ground. Ecosystems of the World Series. Amsterdam: Elsevier Science.Google Scholar
del Moral, R. and Wood, D. M.. 1993. Early primary succession on the volcano Mount St. Helens. Journal of Vegetation Science 4: 223–234.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: 107–120.CrossRefGoogle Scholar
Denslow, J. L. 1987. Tropical rain forest gaps and tree species diversity. Annual Review of Ecology and Systematics 18: 431–451.CrossRefGoogle Scholar
Deshmukh, I. 1986. Ecology and Tropical Biology. Palo Alto: Blackwell Scientific.Google Scholar
Desmond, A. and Moore, J.. 1991. Darwin. New York: Warner Books.Google Scholar
Wit, C. T. 1960. On competition. Verslagen van Landbouwkundige Onderzoekingen 66: 1–82.Google Scholar
Diamond, J. M. 1975. Assembly of species communities. pp. 342–444. In Cody, M. L. and Diamond, J. M. (eds.) Ecology and Evolution of Communities. Cambridge: Belknap Press of Harvard University Press.Google Scholar
Diamond, J. M. 1986. Overview: laboratory experiments, field experiments, and natural experiments. pp. 3–22. In Diamond, J. M. and Case, T. J. (eds.) Community Ecology. New York: Harper and Row.Google Scholar
Diamond, J. 1994. Ecological collapses of past civilisations. Proceedings of the American Philosophical Society 138: 363–370.Google Scholar
Diamond, J. 2004. Twilight at Easter. New York Review of Books LI (5) (March 25). pp. 6–10.Google Scholar
Díaz, S., Acosta, A., and Cabido, M.. 1992. Morphological analysis of herbaceous communities under different grazing regimes. Journal of Vegetation Science 3: 689–696.CrossRefGoogle Scholar
Dickerson, R. E. 1969. Molecular Thermodynamics. New York: W. A Benjamin Inc.Google Scholar
Digby, P. G. N. and Kempton, R. A.. 1987. Multivariate Analysis of Ecological Communities. London: Chapman and Hall.Google Scholar
Dilcher, D. L. and Crane, P. R.. 1985. Archaeanthus: an early angiosperm from the Cenomanian of the western interior of North America. Annals of the Missouri Botanical Garden 71: 351–383.CrossRefGoogle Scholar
Dilcher, D. L. and Kovach, W. L.. 1986. Early angiosperm reproduction: Caloda delevoryana gen. et sp. nov., a new fructification from the Dakota Formation (Cenomanian) of Kansas. American Journal of Botany 73: 1230–1237.CrossRefGoogle Scholar
Dinerstein, E. 1991. Seed dispersal by greater one-horned rhinoceros (Rhinoceros unicornis) and the flora of Rhinoceros latrines. Mammalia 55: 355–362.CrossRefGoogle Scholar
Dinerstein, E. 1992. Effects of Rhinoceros unicornis on riverine forest structure in lowland Nepal. Ecology 73: 701–704.CrossRefGoogle Scholar
Dirzo, R., Horvitz, C. C., Quevedo, H., and López, M. A.. 1992. The effects of gap size and age on the understorey herb community of a tropical Mexican rain forest. Journal of Ecology 80: 809–822.CrossRefGoogle Scholar
Dodson, C. H. 1991. Orchidales, pp. 738–746. In Angiosperms: The Flowering Plants. pp. 596–765, Vol. 13. The New Encyclopaedia Britannica, 15th edn. Chicago: The University of Chicago.Google Scholar
Douglas, R. J. W. 1972. Geology and Economic Minerals of Canada. Ottawa: Geological Survey of Canada.Google Scholar
Dowdeswell, J. A. 2006. The Greenland Ice Sheet and global sea-level rise. Science 311: 963–964.CrossRefGoogle ScholarPubMed
Dressler, R. L. 1983. Classification of the Orchidaceae and their probable origin. Telopea 2: 413–424.CrossRefGoogle Scholar
Drew, M. C. 1975. Comparison of the effects of a localized supply of phosphate, nitrate, ammonium and potassium on the growth of the seminal root system, and the shoot, in barley. New Phytologist 75: 479–490.CrossRefGoogle Scholar
Duchesne, L. C. and Larson, D. W.. 1989. Cellulose and the evolution of plant life. Bioscience 39: 238–241.CrossRefGoogle Scholar
Dugan, P. (ed.) 1993. Wetlands in Danger. New York: Oxford University Press.Google 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: 403–416.CrossRefGoogle Scholar
Durant, W. 1944. Caesar and Christ. New York: Simon and Schuster.Google Scholar
du Rietz, G. E. 1931. Life-forms of Terrestrial Flowering Plants. Acta Phytogeographica Suecia. III. Uppsala: Almqvist and Wiksells.Google Scholar
Dynesius, M. and Nilsson, C.. 1994. Fragmentation and flow regulation of river systems in the northern third of the world. Science 266: 753–762.CrossRefGoogle Scholar
Earth Impact Database. 2006. (http://www.unb.ca/passc/ImpactDatabase) accessed 12 Sept. 2006.
Edmonds, J. (ed.) 1997. Oxford Atlas of Exploration. New York: Oxford University Press.Google Scholar
Ehrenfeld, J. G. 1983. The effects of changes in land-use on swamps of the New Jersey Pine Barrens. Biological Conservation 25: 353–375.CrossRefGoogle Scholar
Ehrlich, A. and Ehrlich, P.. 1981. Extinction: The Causes and Consequences of the Disappearance of Species. New York: Random House.Google Scholar
Ehrlich, P. and Raven, P. H. 1964. Butterflies and plants: a study in coevolution. Evolution 18: 586–608.CrossRefGoogle Scholar
Eissenstat, D. M. and Newman, E. I.. 1990. Seedling establishment near large plants: effects of vesicular-arbuscular mycorrhizae on the intensity of plant competition. Functional Ecology 4: 95–99.CrossRefGoogle Scholar
Ellenberg, H. 1985. Veränderungen der Flora Mitteleuropas unter dem Einfluß von Düngung und Immissionen. Schweizerische Zeitschrift für Forstwesen 136: 19–39.Google Scholar
Ellenberg, H. 1988a. Floristic changes due to nitrogen deposition in central Europe. In J. Nilsson and P. Grennfelt (eds.) Critical Loads for Sulphur and Nitrogen. Report from a workshop held at Skokloster, Sweden, March 19–24, 1988.
Ellenberg, H. 1988b. Vegetation Ecology of Central Europe. 4th edn. Cambridge: Cambridge University Press. Translated by G. K. Strutt.Google 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: 717–731.CrossRefGoogle Scholar
Elton, C. 1927. Animal Ecology. London: Sidgwick and Jackson Ltd.Google Scholar
Encyclopaedia Britannica. 1991a. Vol. 16. p. 500. Chicago: Encyclopaedia Britannica Inc.
Encyclopaedia Britannica. 1991b. Vol. 12. p. 41. Chicago: Encyclopaedia Britannica Inc.
Encyclopaedia Britannica. 1991c. Vol. 16. p. 481. Chicago: Encyclopaedia Britannica Inc.
Endress, P. K. 1996. Diversity and Evolutionary Biology of Tropical Flowers. Paperback edition (with corrections). Cambridge: Cambridge University Press.Google Scholar
Englert, S. 1970. Islands at the Center of the World: New Light on Easter Island. New York: Charles Scribner. Translated by W. Mulloy.Google Scholar
Environment Canada. 1976. Marine Environmental Data Service, Ocean and Aquatic Sciences. Monthly and Yearly Mean Water Levels, Vol. 1. Inland. Ottawa: Department of Environment.
Eriksson, O. 1993. The species-pool hypothesis and plant community diversity. Oikos 68: 371–374.CrossRefGoogle Scholar
Ernst, W. 1978. Discrepancy between ecological and physiological optima of plant species: a re-interpretation. Oecologia Plantarum 13: 175–188.Google Scholar
Estill, J. C. and Cruzan, M. B.. 2001. Phytogeography and rare plant species endemic to the southeastern United States. Castanea 66: 3–23.Google 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. The American Midland Naturalist 121: 125–133.CrossRefGoogle Scholar
Farjon, A. 1998. World Checklist and Bibliography of Conifers. Royal Botanical Gardens at Kew, Richmond, UK.
Farrow, E. P. 1917. On the ecology of the vegetation of Breckland. III. General effects of rabbits on the vegetation. Journal of Ecology 5: 1–18.CrossRefGoogle Scholar
Faulkner, S. P. and C. J. Richardson. 1989. Physical and chemical characteristics of freshwater wetland soils. pp. 41–72. In Hammer, D. A. (ed.) Constructed Wetlands for Wastewater Treatment. Municipal, Industrial, and Agricultural. Chelsea: Lewis Publishers.Google Scholar
Fedorov, A. V., Dekens, P. S., McCarthy, M., Ravelo, A. C., deMenocal, P. B., Barreiro, M., Pacanowski, R. C., and Philander, S. G.. 2006. The Pliocene paradox (mechanisms for a permanent El Niño). Science 312: 1485–1489.CrossRefGoogle Scholar
Feinsinger, P. 1976. Organisation of a tropical guild of nectivorous birds. Ecological Monographs 46: 257–291.CrossRefGoogle Scholar
Feinsinger, P. 1993. Coevolution and pollination. pp. 282–310. In Futuyma, D. J. and Slatkin, M. (eds.) Coevolution. Sunderland: Sinauer.Google Scholar
Fernald, M. L. 1921. The Gray Herbarium expedition to Nova Scotia 1920. Rhodora 23: 89–111, 130–171, 184–195, 233–245, 257–278, 284–300.Google Scholar
Fernald, M. L. 1922. Notes on the flora of western Nova Scotia 1921. Rhodora 24: 157–164, 165–180, 201–208.Google Scholar
Fernald, M. L. 1935. Critical plants of the upper Great Lakes region of Ontario and Michigan. Rhodora 37: 197–222, 238–262, 272–301, 324–341.Google Scholar
Fernández-Armesto, F. 1989. The Spanish Armada: The Experience of War in 1588. Oxford: Oxford University Press.Google Scholar
Ferris, T. 1988. Coming of Age in the Milky Way. Anchor Books edition 1989. New York: Doubleday.Google Scholar
Fienberg, S. E. and Hinkley, D. V. (eds.). 1980. R. A. Fisher: An Appreciation. New York: Springer-Verlag.CrossRefGoogle Scholar
Firbank, L. G. and Watkinson, A. R.. 1985. On the analysis of competition within two-species mixtures of plants. Journal of Applied Ecology 22: 503–517.CrossRefGoogle Scholar
Fischer, R., W. De Vries, W. Seidling, P. Kennedy, and M. Lorenz. 2000. Forest Condition in Europe. 2000 Executive Report. United Nations Economic Commission for Europe/European Commission, Geneva and Brussels.
Fisher, R. A. 1925. Statistical Methods for Research Workers. London: Oliver and Boyd.Google Scholar
Fitter, A. H. and Hay, R. K. M.. 1983. Environmental Physiology of Plants. London: Academic Press.Google Scholar
Flannery, T. 2001. The Eternal Frontier: An Ecological History of North America and its Peoples. Melbourne: Text Publishing.Google Scholar
Flannery, T. 2005. The Weather Makers: How Man is Changing the Climate and What it Means for Life on Earth. New York: Atlantic Monthly Press.Google Scholar
Fleischner, T. L. 1994. Ecological costs of livestock grazing in western North America. Conservation Biology 8: 629–644.CrossRefGoogle Scholar
Flint, R. F. 1971. Glacial and Quaternary Geology. New York: John Wiley and Sons.Google Scholar
Fonteyn, P. J. and Mahall, B. E.. 1978. Competition among desert perennials. Nature 275: 544–545.CrossRefGoogle Scholar
Fonteyn, P. J. and Mahall, B. E.. 1981. An experimental analysis of structure in a desert plant community. Journal of Ecology 69: 883–896.CrossRefGoogle Scholar
Forde, B. and Zhang, H.. 1998. Response: nitrate and root branching. Trends in Plant Science 3: 204–205.CrossRefGoogle Scholar
Foreman, D. 2004. Rewilding North America: A Vision for Conservation in the 21st Century. Washington, D.C.: Island Press.Google Scholar
Forman, R. T. T. 1964. Growth under controlled conditions to explain the hierarchical distributions of a moss, Tetraphis pellucida. Ecological Monographs 34: 1–25.CrossRefGoogle Scholar
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: Island Press.Google Scholar
Foster, A. S. and Gifford, E. M. Jr. 1974. Comparative Morphology of Vascular Plants. 2nd edn. San Francisco: W. H. Freeman and Company.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. Jr. 1990. Role of ecosystem development and climate change in bog formation in central Sweden. Ecology 71: 450–463.CrossRefGoogle Scholar
Fowler, N. 1981. Competition and coexistence in a North Carolina grassland. II. The effects of the experimental removal of species. Journal of Ecology 69: 843–845.CrossRefGoogle Scholar
Fox, J. F. 1977. Alternation and coexistence of tree species. The American Naturalist 111: 69–89.CrossRefGoogle Scholar
Fragoso, J. M. V. 1997. Tapir-generated seed shadows: scale-dependent patchiness in the Amazon rain forest. Journal of Ecology 85: 519–529.CrossRefGoogle Scholar
Francis, R. and Read, D. J.. 1984. Direct transfer of carbon between plants connected by vesicular-arbuscular mycorrhizal mycelium. Nature 307: 53–56.CrossRefGoogle Scholar
Franco, A. C. and Nobel, P. S.. 1989. Effect of nurse plants on the microhabit and growth of cacti. Journal of Ecology 77: 870–886.CrossRefGoogle Scholar
Fraser, L. H. and Keddy, P. 1997. The role of experimental microcosms in ecological research. Trends in Ecology and Evolution 12: 478–481.CrossRefGoogle ScholarPubMed
Fraser, L. H. and Keddy, P. A. (eds.). 2005. The World's Largest Wetlands: Ecology and Conservation.Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Freedman, B. 1995. Environmental Ecology. 2nd edn. San Diego: Academic Press.Google Scholar
Freedman, B., Zobens, W., Hutchinson, T. C., and Gizyn, W. I.. 1990. Intense, natural pollution affects arctic tundra vegetation at the Smoking Hills, Canada. Ecology 71: 492–503.CrossRefGoogle Scholar
Freemark, K. E. and Merriam, H. G.. 1986. The importance of area and habitat heterogeneity to bird assemblages in temperate forest fragments. Biological Conservation 36: 115–141.CrossRefGoogle Scholar
French, B. M. 1998. Traces of Catastrophe: A Handbook of Shock-Metamorphic Effects in Terrestrial Meteoric Impact Structures. LPI Contribution No. 954. Houston: Lunar and Planetary Institute.Google Scholar
Fretwell, S. D. 1977. The regulation of plant communities by food chains exploiting them. Perspectives in Biology and Medicine 20: 169–185.CrossRefGoogle Scholar
Frey, R. W. and P. B. Basan. 1978. Coastal salt marshes. pp. 101–169. In Davis, R. A. (ed.) Coastal Sedimentary Environments. New York: Springer-Verlag.CrossRefGoogle Scholar
Frey, T. E. 1973. The Finnish school and forest site types. pp. 403–433. In Whittaker, R. H. (ed.) Ordination and Classification of Communities. Part V. The Hague: W. Junk.CrossRefGoogle Scholar
Friedmann, E. I. 1982. Endolithic microorganisms in the antarctic cold desert. Science 215: 1045–1053.CrossRefGoogle ScholarPubMed
Friis, E. M., Pedersen, K. R., and Crane, P. R.. 2006. Cretaceous angiosperm flowers: Innovation and evolution in plant reproduction. Palaeogeography, Palaeoclimatology, Palaeoecology 23: 251–293.CrossRefGoogle Scholar
Frontier, S. 1985. Diversity and structure in aquatic ecosystems. Oceanography and Marine Biology Annual Review 23: 253–312.Google Scholar
Futuyma, D. J. and M. Slatkin. 1993. The study of coevolution. pp. 459–464. In Futuyma, D. J. and Slatkin, M. (eds.) Coevolution. Sunderland: Sinauer.Google Scholar
Galatowitsch, S. M. and Valk, A. G.. 1994. Restoring Prairie Wetlands: An Ecological Approach. Ames: Iowa State University Press.Google Scholar
Gardner, G. 1977. The reproductive capacity of Fraxinus excelsior on the Derbyshire limestone. Journal of Ecology 65: 107–118.CrossRefGoogle Scholar
Gaston, K. J. 2000. Global patterns in biodiversity. Nature 405: 220–227.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-Biological Sciences 260: 149–154.CrossRefGoogle Scholar
Gauch, H. G. Jr. 1982. Multivariate Analysis in Community Ecology. Cambridge Studies in Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Gauch, H. G. Jr. and Wentworth, T. R.. 1976. Canonical correlation analysis as an ordination technique. Vegetatio 33: 17–22.CrossRefGoogle Scholar
Gauch, H. G. Jr. and Whittaker, R. H.. 1972. Comparison of ordination techniques. Ecology 53: 868–875.CrossRefGoogle Scholar
Gauch, H. G. Jr., Whittaker, R. H., and Wentworth, T. R.. 1977. A comparative study of reciprocal averaging and other techniques. Journal of Ecology 65: 157–174.Google Scholar
Gaudet, C. L. 1993. Competition in shoreline plant communities: a comparative approach. PhD thesis. Ottawa: University of Ottawa.
Gaudet, C. L. and Keddy, P. A.. 1988. A comparative approach to predicting competitive ability from plant traits. Nature 334: 242–243.CrossRefGoogle Scholar
Gause, G. F. and Witt, A. A.. 1935. Behavior of mixed populations and the problem of natural selection. The American Naturalist 69: 596–609.CrossRefGoogle Scholar
Geis, J. W. 1985. Environmental influences on the distribution and composition of wetlands in the Great Lakes basin. pp. 15–31. In Prince, H. H. and D'Itri, F. M. (eds.) Coastal Wetlands. Chelsea: 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
Gibson, A. C. and Nobel, P. S.. 1986. The Cactus Primer. Cambridge: Harvard University Press.CrossRefGoogle Scholar
Gibson, C. W. D. and Hamilton, J.. 1983. Feeding ecology and seasonal movements of giant tortoises on Aldabra atoll. Oecologia 56: 84–92.CrossRefGoogle ScholarPubMed
Gidley, I. and Shears, R.. 1986. The Rainbow Warrior Affair. Toronto: Irwin Publishing.Google Scholar
Given, D. R. and Soper, J.. 1981. The Arctic-Alpine Element of the Vascular Flora at Lake Superior. Publications in Botany No. 10. Ottawa: National Museums of Canada.Google Scholar
Givnish, T. J. 1982. On the adaptive significance of leaf height in forest herbs. The American Naturalist 120: 353–381.CrossRefGoogle Scholar
Givnish, T. J. 1984. Leaf and canopy adaptations in tropical forests. pp. 51–84. In Medina, E., Mooney, H. A., and Vásquez-Yánes, C. (eds.) Physiological Ecology of Plants of the Wet Tropics. The Hague: Dr. Junk.Google Scholar
Givnish, T. J. 1987. Comparative studies of leaf form: assessing the relative roles of selective pressures and phylogenetic constraints. New Phytologist 106 (Suppl.): 131–160.CrossRefGoogle Scholar
Givnish, T. J. 1988. Ecology and evolution of carnivorous plants. pp. 243–290. In Abrahamson, W. B. (ed.) Plant-Animal Interactions. New York: McGraw-Hill.Google Scholar
Givnish, T. J. 1994. Does diversity beget stability?Nature 371: 113–114.CrossRefGoogle Scholar
Glaser, P. H. 1992. Raised bogs in eastern North America – regional controls for species richness and floristic assemblages. Journal of Ecology 80: 535–554.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: 1021–1048.CrossRefGoogle Scholar
Gleason, H. A. 1917. The structure and development of the plant association. Bulletin of the Torrey Botanical Club 44: 463–481.CrossRefGoogle Scholar
Gleason, H. A. 1926. The individualistic concept of the plant association. Bulletin of theTorrey Botanical Club 53: 7–26.CrossRefGoogle Scholar
Glenn-Lewin, D. C., Peet, R. K., and Veblen, T. T. (eds.) 1992. Plant Succession: Theory and Prediction. Population and Community Biology. No. 11. London: Chapman and Hall.Google Scholar
Glooschenko, W. A. 1980. Coastal salt marshes in Canada. pp. 39–47. In C. D. A. Rubec and F. C. Pollet (eds.) Proceedings of the Workshop on Canadian Wetlands. Saskatoon, Saskatchewan. Environment Canada, Lands Directorate, Ecological Land Class. Series No. 12.
Gnanadesikan, R. 1997. Methods for Statistical Data Analysis of Multivariate Observations. 2nd edn. New York: Wiley.CrossRefGoogle Scholar
Goebel, K. 1905. Wilhelm Hofmeister. The Plant World 8: 291–298.Google Scholar
Goldberg, D. E. 1982a. The distribution of evergreen and deciduous trees relative to soil type: an example from the Sierra Madre, Mexico, and a general model. Ecology 63: 942–951.CrossRefGoogle Scholar
Goldberg, D. E. 1982b. Comparison of factors determining growth rates of deciduous vs. broad-leaf evergreen trees. The American Midland Naturalist 108: 133–143.CrossRefGoogle Scholar
Goldberg, D. E. 1990. Components of resource competition in plant communities. pp. 27–49. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.
Goldberg, D. E. and Landa, K.. 1991. Competitive effect and response: hierarchies and correlated traits in the early stages of competition. Journal of Ecology 79: 1013–1030.CrossRefGoogle Scholar
Goldberg, D. E. and Werner, P. A.. 1983. Equivalence of competitors in plant communities: a null hypothesis and a field experimental approach. American Journal of Botany 70: 1098–1104.CrossRefGoogle Scholar
Goldsmith, F. B. 1973a. The vegetation of exposed sea cliffs at South Stack, Anglesey: I. The multivariate approach. Journal of Ecology 61: 787–818.CrossRefGoogle Scholar
Goldsmith, F. B. 1973b. The vegetation of exposed sea cliffs at South Stack, Anglesey: II. Experimental studies. Journal of Ecology 61: 819–829.CrossRefGoogle Scholar
Goldsmith, F. B. 1978. Interaction (competition) studies as a step towards the synthesis of sea-cliff vegetation. Journal of Ecology 66: 921–931.CrossRefGoogle Scholar
Goldsmith, F. B. and C. M. Harrison. 1976. Description and analysis of vegetation. pp. 85–155. In Chapman, S. B. (ed.) Methods in Plant Ecology. Oxford: Blackwell Scientific.Google Scholar
Gopal, B. 1990. Nutrient dynamics of aquatic plant communities. pp. 177–197. In Gopal, B. (ed.) Ecology and Management of Aquatic Vegetation in the Indian Subcontinent. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Gopal, B. and Goel, U.. 1993. Competition allelopathy in aquatic plant communities. The Botanical Review 59: 155–210.CrossRefGoogle Scholar
Gopal, B., J. Kvet, H. Loffler, V. Masing, and B. C. Patten. 1990. Definition and classification. pp. 9–15. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 1. Natural and Human Relationships. The Hague: SPB Academic Publishing.Google Scholar
Gore, A. J. P. 1983. Introduction. In Gore, A. J. P. (ed.) Ecosystems of the World 4A. Mires: Swamp, Bog, Fen and Moor. Amsterdam: Elsevier Scientific Publishing Company.Google Scholar
Gore, A. 2006. An Inconvenient Truth. The Planetary Emergency of Global Warming and What We Can Do About It. New York: Melcher Media/Rodale.Google Scholar
Goremykin, V. V., Hirsch-Ernst, K. I., Wölfl, S., and Hellwig, F. H.. 2003. Analysis of the Amborella trichopoda chloroplast genome sequence suggests that Amborella is not a basal angiosperm. Molecular Biology and Evolution 20: 1499–1505.CrossRefGoogle Scholar
Gorham, E. 1953. Some early ideas concerning the nature, origin and development of peat lands. Journal of Ecology 41: 257–274.CrossRefGoogle Scholar
Gorham, E. 1957. The development of peat lands. The Quarterly Review of Biology 32: 145–166.CrossRefGoogle Scholar
Gorham, E. 1979. Shoot height, weight and standing crop in relation to density of nonspecific plant stands. Nature 279: 148–150.CrossRefGoogle Scholar
Gorham, E. 1990. Biotic impoverishment in northern peatlands. pp. 65–98. In Woodwell, G. M. (ed.) The Earth in Transition. Cambridge: Cambridge University Press.Google Scholar
Gorham, E. 1991. Northern peatlands role in the carbon cycle and probable responses to climatic warming. Ecological Applications 1: 182–195.CrossRefGoogle ScholarPubMed
Gosselink, J. G. and R. E. Turner. 1978. The role of hydrology in freshwater wetland ecosystems. pp. 63–78. In Good, R. E., Whigham, D. F., and Simpson, R. L. (eds.) Freshwater Wetlands: Ecological Processes and Management Potential. New York: Academic Press.Google Scholar
Gosselink, J. G., J. M. Coleman, and R. E. Stewart, Jr. 1998. Coastal Louisiana. pp. 385–436. In Mac, M. J., Opler, P. A., Haecker, C. E. Puckett, and Doran, P. D. (eds.) 1998. Status and Trends of the Nation's Biological Resources, 2 Vols. Reston: U.S. Department of the Interior, U.S. Geological Survey.Google Scholar
Gotelli, N. J. and Graves, G. R.. 1996. Null Models in Ecology. Washington, D.C.: Smithsonian Institution Press.Google 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 Manchester4: 1–19. In D. Walker. 1970. Direction and rate in some British post-glacial hydoseres. pp. 117–139. In Walker, D. and West, R. G. (eds.) Studies in the Vegetational History of the British Isles. Cambridge: Cambridge University Press.Google Scholar
Gough, L., Grace, J. B., and Taylor, K. L.. 1994. The relationship between species richness and community biomass: the importance of environmental variables. Oikos 70: 271–279.CrossRefGoogle Scholar
Gould, S. J. 1977. Ever Since Darwin: Reflections in Natural History. New York: W. W. Norton and Company.Google Scholar
Grace, J. B. 1993. The effects of habitat productivity on competition intensity. Trends in Ecology and Evolution 8: 229–230.CrossRefGoogle ScholarPubMed
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. 2001. The roles of community biomass and species pools in the regulation of plant diversity. Oikos 92: 193–207.CrossRefGoogle Scholar
Grace, J. B. 2006. Structural Equation Modeling and Natural Systems. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Grace, J. B. and Pugesek, B. H.. 1997. A structural equation model of plant species richness and its application to a coastal wetland. The American Naturalist 149: 436–460.CrossRefGoogle Scholar
Grace, J. B. and Tilman, D. (eds.) 1990. Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Grant, M. C. 1993. The trembling giant. Discover 4(10): 82–89.Google Scholar
Green, P. T., O'Dowd, D. J., and Lake, P. S.. 1997. Control of seedling recruitment by land crabs in rain forest on a remote oceanic island. Ecology 78: 2472–2486.CrossRefGoogle Scholar
Greenslade, P. J. M. 1983. Adversity selection and the habitat templet. Nature 242: 344–347.Google Scholar
Greig-Smith, P. 1952. Use of random and contiguous quadrats in the study of the structure of plant communities. Annals of Botany 16: 293–316.CrossRefGoogle Scholar
Greig-Smith, P. 1957. Quantitative Plant Ecology. London: Butterworths.Google Scholar
Grime, J. P. 1973a. Control of species density in herbaceous vegetation. Journal of Environmental Management 1: 151–167.Google Scholar
Grime, J. P. 1973b. Competitive exclusion in herbaceous vegetation. Nature 242: 344–347.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: 1169–1194.CrossRefGoogle Scholar
Grime, J. P. 1979. Plant Strategies and Vegetation Processes. Chichester: John Wiley.Google Scholar
Grime, J. P. 1994. The role of plasticity in exploiting environmental heterogeneity. pp. 1–19. In Caldwell, M. M. and Percy, R. W. (eds.) Exploitation of Environmental Heterogeneity by Plants: Ecophysical Processes Above- and Belowground. San Diego: Academic Press.Google Scholar
Grime, J. P. 1997. The humped-back model: a response to Oksanen. Journal of Ecology 85: 97–98.CrossRefGoogle Scholar
Grime, J. P. 2002. Declining plant diversity: empty niches or functional shifts?Journal of Vegetation Science 13: 457–460.CrossRefGoogle 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. and Jeffrey, D. W.. 1965. Seedling establishment in vertical gradients of sunlight. Journal of Ecology 53: 621–642.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: 1017–1059.CrossRefGoogle Scholar
Grime, J. P., Mackey, J. M. L., Hillier, S. H., and Read, D. J.. 1987. Floristic diversity in a model system using experimental microcosms. Nature 328: 420–422.CrossRefGoogle 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: 129–153.CrossRefGoogle Scholar
Groombridge, B. (ed.). 1992. Global Biodiversity: Status of the Earth's Living Resources. London: Chapman and Hall.CrossRefGoogle Scholar
Grover, A. M. and Baldassarre, G. A.. 1995. Bird species richness within beaver ponds in south-central New York. Wetlands 15: 108–118.CrossRefGoogle Scholar
Grubb, P. J. 1977. The maintenance of species-richness in plant communities: the importance of the regeneration niche. Biological Reviews 52: 107–145.CrossRefGoogle Scholar
Grubb, P. J. 1987. Global trends in species-richness in terrestrial vegetation: a view from the Northern Hemisphere. pp. 99–118. In Gee, J. H. R. and Giller, P. S. (eds.) Organization of Communities Past and Present. Oxford: Blackwell Scientific Publications.Google Scholar
Grumbine, R. E. 1997. Reflections on “What is ecosystem management?”Conservation Biology 11: 41–47.CrossRefGoogle Scholar
Guariguata, M. R. 1990. Landslide disturbance and forest regeneration in the Upper Luquillo mountains of Puerto Rico. Journal of Ecology 78: 814–832.CrossRefGoogle Scholar
Guerlac, H. 1975. Antoine-Laurent Lavoisier, Chemist and Revolutionary. New York: Charles Scribner's Sons.Google Scholar
Gurevitch, J. and Unnasch, R. S.. 1989. Experimental removal of a dominant species at two levels of soil fertility. Canadian Journal of Botany 67: 3470–3477.CrossRefGoogle Scholar
Haber, L. F. 1986. The Poisonous Cloud. Chemical Warfare in the First World War. Oxford: Clarendon Press.Google Scholar
Haffer, J. 1969. Speciation in Amazonian forest birds. Science 165: 131–137.CrossRefGoogle ScholarPubMed
Hairston, N. G., Smith, F. E., and Slobodkin, L. B.. 1960. Community structure, population control, and competition. The American Naturalist XCIV: 421–425.CrossRefGoogle Scholar
Hamann, O. 1979. Regeneration of vegetation on Santa Fe and Pinta Islands, Galápagos, after the eradication of goats. Biological Conservation 15: 215–236.CrossRefGoogle Scholar
Hamann, O. 1993. On vegetation recovery, goats and giant tortoises on Pinta Island, Galápagos, Ecuador. Biodiversity and Conservation 2: 138–151.CrossRefGoogle Scholar
Harper, J. L. 1965. The nature and consequence of interference amongst plants. Genetics Today 2: 465–482.Google Scholar
Harper, J. L. 1967. A Darwinian approach to plant ecology. Journal of Ecology 55: 247–270.CrossRefGoogle Scholar
Harper, J. L. 1977. Population Biology of Plants. London: Academic Press.Google Scholar
Harper, J. L. 1982. After description. pp. 11–25. In Newman, E. I. (ed.) The Plant Community as a Working Mechanism. Oxford: Blackwell.Google Scholar
Harper, J. L. and Ogden, J.. 1970. The reproductive strategy of higher plants. I. The concept of strategy with special reference to Senecio vulgaris L. Journal of Ecology 58: 681–698.CrossRefGoogle Scholar
Harper, J. L. and White, J.. 1974. The demography of plants. Annual Review of Ecology and Systematics 5: 419–463.CrossRefGoogle Scholar
Harper, J. L., Rosen, B. R., and White, J.. 1986. The Growth and Form of Modular Organisms. London: The Royal Society.Google Scholar
Harris, L. D. 1984. The Fragmented Forest: Island Biogeography Theory and the Preservation of Biotic Diversity. Chicago: University of Chicago Press.Google Scholar
Hartman, J. M. 1988. Recolonization of small disturbance patches in a New England salt marsh. American Journal of Botany 75: 1625–1631.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
Hatch, A. B. 1937. The physical basis of mycotrophy in Pinus. The Black Rock Forest Bulletin, No. 6. 17 pp.Google Scholar
Hawksworth, D. L. 1988. Coevolution of fungi with algae and cyanobacteria in lichen symbioses. pp. 125–148. In Pirozynski, K. A. and Hawksworth, D. L. (eds.) Coevolution of Fungi with Plants and Animals. London: Academic Press.Google Scholar
Hawksworth, D. L. 1990. The fungal dimension of biodiversity: magnitude, significance, and conservation. Mycological Research 95: 641–655.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
Heady, H. F. (ed.) 1988. The Vale Rangeland Rehabilitation Program: An Evaluation. USDA Forest Service, Resource Bulletin PNW-RB-157, 151 pp.
Heady, H. F. and J. Bartolome. 1977. The Vale Rangeland Rehabilitation Program: The Desert Repaired in Southeastern Oregon. USDA Forest Service, Resource Bulletin PHW-70, 139 pp.
Heckman, D. S., Geiser, D. M., Eidell, B. R., Stauffer, R. L., Kardos, N. L., and Hedges, S. B.. 2001. Molecular evidence for the early colonization of land by fungi and plants. Science 293: 1129–1133.CrossRefGoogle ScholarPubMed
Heinselman, M. L. 1973. Fire in the virgin forests of the Boundary Waters Canoe Area, Minnesota. Quaternary Research 3: 329–382.CrossRefGoogle Scholar
Heinselman, M. L. 1981. Fire and succession in the conifer forests of northern North America. pp. 374–405. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Applications. New York: Springer-Verlag.CrossRefGoogle Scholar
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: 378–382.CrossRefGoogle Scholar
Higgs, E. S. 1997. What is good ecological restoration?Conservation Biology 11: 338–348.CrossRefGoogle Scholar
Hill, N. M. and Keddy, P. A.. 1992. Predicting numbers of rarities from habitat variables: coastal plain plants of Nova Scotian lakeshores. Ecology 73: 1852–1859.CrossRefGoogle Scholar
Hills, G. A. 1961. The Ecological Basis for Land-Use Planning. Report No. 46. Ontario: Ontario Department of Lands and Forests, Research Branch.
Hoagland, B. W. and Collins, S. L.. 1997. Gradient models, gradient analysis, and hierarchical structure in plant communities. Oikos 78: 23–30.CrossRefGoogle Scholar
Hoffman, T. M., Midgley, G. F., and Cowling, R. M.. 1994. Plant richness is negatively related to energy availability in semi-arid southern Africa. Biodiversity Letters 2: 35–38.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: 1991–1997.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
Holechek, J. L., Vavra, M., and Pieper, R. D.. 1982. Botanical composition determination of herbivore diets: a review. Journal of Range Management 31: 309–315.CrossRefGoogle Scholar
Holling, C. S. 1959. The components of predation as revealed by a study of small-mammal predation of the European pine sawfly. Canadian Entomologist 91: 293–320.CrossRefGoogle Scholar
Holling, C. S. (ed.) 1978a. Adaptive Environmental Assessment and Management. New York: John Wiley and Sons.Google Scholar
Holling, C. S. 1978b. The spruce-budworm/forest-management problem. pp. 143–182. In Holling, C. S. (ed.) Adaptive Environmental Assessment and Management. New York: John Wiley and Sons.Google Scholar
Holt, R. D. and Lawton, J. H.. 1993. Apparent competition and enemy-free space in insect host-parasitoid communities. The American Naturalist 142: 623–645.CrossRefGoogle ScholarPubMed
Holt, R. D. and Lawton, J. H.. 1994. The ecological consequences of shared natural enemies. Annual Review of Ecology and Systematics 25: 495–520.CrossRefGoogle Scholar
Hook, D. D. 1984. Adaptations to flooding with fresh water. pp. 265–294. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Hopkins, D. M. (ed.) 1967. The Bering Land Bridge. Stanford: Stanford University Press.Google Scholar
Horn, H. S. 1971. The Adaptive Geometry of Trees. Princeton: Princeton University Press.Google Scholar
Horn, H. 1976. Succession. pp. 187–204. In May, R. M. (ed.) Theoretical Ecology: Principles and Applications. Philadelphia: W. B. Saunders.
Horn, H. S. 1981. Some causes of variety of patterns of secondary succession. pp. 24–35. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
Horn, H. S. and MacArthur, R. H.. 1972. Competition among fugitive species in a harlequin environment. Ecology 53: 749–752.CrossRefGoogle Scholar
Houck, O. 2006. Can we save New Orleans?Tulane Environmental Law Journal 19: 1–68.Google Scholar
Hubbell, S. P. and R. B. Foster. 1986. Biology, chance and history and the structure of the tropical rain forest tree communities. pp. 314–329. In Diamond, J. and Case, T. J. (eds.) Community Ecology. New York: Harper and Row.Google Scholar
Huber, H., Lukács, S., and Watson, M. A.. 1999. Spatial structure of stoloniferous herbs: an interplay between structural blue-print, ontogeny and phenotypic plasticity. Plant Ecology 141: 107–115.CrossRefGoogle Scholar
Hughes, J. D. 1982. Deforestation, erosion, and forest management in ancient Greece and Rome. Journal of Forest History 26: 60–75.Google Scholar
Humphries, C. J. 1981. Biogeographical methods and the southern beeches. pp. 283–297. In Forey, P. L. (ed.) The Evolving Biosphere. British Museum (Natural History). Cambridge: London and Cambridge University Press.Google Scholar
Hunt, R., Hand, D. W., Hannah, M. A., and Neal, A. M.. 1991. Response to CO2 enrichment in 27 herbaceous species. Functional Ecology 5: 410–421.CrossRefGoogle 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: 724–732.Google Scholar
Huntley, B. 1990. European post-glacial forests: compositional changes in response to climatic change. Journal of Vegetation Science 1: 507–518.CrossRefGoogle 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: 257–271.CrossRefGoogle Scholar
Huston, M. A. 1979. A general hypothesis of species diversity. The American Naturalist 113: 81–101.CrossRefGoogle Scholar
Huston, M. A. 1994. Biological Diversity. The Coexistence of Species on Changing Landscapes. Cambridge: Cambridge University Press.Google Scholar
Huston, M. A. 1997. Hidden treatments in ecological experiments: re-evaluating the ecosystem function of biodiversity. Oecologia 110: 449–460.CrossRefGoogle ScholarPubMed
Hutchinson, G. E. 1959. Homage to Santa Rosalia; or, why are there so many kinds of animals?The American Naturalist 93: 145–159.CrossRefGoogle Scholar
Hutchinson, G. E. 1970. The biosphere. pp. 194–203. In Wilson, E. O. (ed.) 1974. Ecology, Evolution, and Population Biology. Readings from Scientific American. San Francisco: W.H. Freeman and Company.Google Scholar
Hutchinson, G. E. 1975. A Treatise on Limnology, Vol. 3. Limnological Botany. New York: John Wiley and Sons.Google Scholar
Huxley, C. R. 1980. Symbiosis between plants and epiphytes. Biological Reviews 55: 321–340.CrossRefGoogle Scholar
Imbrie, J., et al. 1992. On the structure and origin of major glaciation cycles 2. The 100,000-year cycle. Paleoceanography 8: 699–736.CrossRefGoogle Scholar
International Joint Commission. 1980. Pollution in the Great Lakes Basin from Land Use Activities. Washington, D.C.: International Joint Commission.
Irion, G. M., Müller, J., Mello, J. N., and Junk, W. J.. 1995. Quaternary geology of the Amazon lowland. Geo-Marine Letters 15: 172–178.CrossRefGoogle Scholar
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, J. B. C., Buss, L. W., and Cook, R. E.. 1985. Population Biology and Evolution of Clonal Organisms. New Haven: Yale University Press.Google Scholar
Jackson, M. B. and M. C. Drew. 1984. Effects of flooding on growth and metabolism of herbaceous plants. pp. 47–128. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Jaksic, F. M. and Fuentes, E. R.. 1980. Why are native herbs in the Chilean matorral more abundant beneath bushes: microclimate or grazing?Journal of Ecology 68: 665–669.CrossRefGoogle Scholar
James, W. 1907. Pragmatism. Reprinted pp. xv–xvii, 1–89. In Adler, M. J. (ed.) 1990. Great Books of the Western World. Vol. 55. Chicago: Encyclopaedia Britannica.Google Scholar
Janis, C. 1976. The evolutionary strategy of the Equidae and the origins of rumen and cecal digestion. Evolution 30: 757–774.CrossRefGoogle ScholarPubMed
Janssens, F., Peeters, A., Tallowin, J. R. B., Bakker, J. P., Fillat, F., and Oomes, M. J. M.. 1998. Relationship between soil chemical factors and grassland diversity. Plant and Soil 202: 69–78.CrossRefGoogle Scholar
Janzen, D. H. 1966. Coevolution of mutualism between ants and acacias in Central America. Evolution 20: 249–275.CrossRefGoogle ScholarPubMed
Janzen, D. H. 1967. Interaction of the bull's-horn acacia (Acacia cornigera L.) with an ant inhabitant (Pseudomyrmex ferruginea F. Smith) in eastern Mexico. The University of Kansas Science Bulletin XLVII: 315–558.Google Scholar
Janzen, D. H. 1971. Seed predation by animals. Annual Review of Ecology and Systematics 2: 465–492.CrossRefGoogle Scholar
Janzen, D. H. 1974. Epiphytic myrmecophytes in Sarawak: mutualism through the feeding of plants by ants. Biotropica 6: 237–259.CrossRefGoogle Scholar
Janzen, D. H. 1976. Why bamboos wait so long to flower. Annual Review of Ecology and Systematics 7: 347–391.CrossRefGoogle Scholar
Janzen, D. H. 1983. Dispersal of seeds by vertebrate guts. pp. 232–262. In Futuyma, D. J. and Slatkin, M. (eds.) Coevolution. Sunderland: Sinauer.Google Scholar
Janzen, D. H. 1985. The natural history of mutualisms. pp. 40–99. In Boucher, D. H. (ed.) The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Janzen, D. H. and Martin, P. S.. 1982. Neotropical anachronisms: the fruits the gomphotheres ate. Science 215: 19–27.CrossRefGoogle ScholarPubMed
Janzen, D. H., Miller, G. A., Hackforth-Jones, J., Pond, C. M., Hooper, K., and Janos, D. P.. 1976. Two Costa Rican bat-generated seed shadows of Andira inermis (Leguminosae). Ecology 57: 1068–1075.CrossRefGoogle Scholar
Jarvis, P. G. 1964. Interference by Deschampsia flexuosa (L.) Trin. Oikos 15: 56–78.CrossRefGoogle Scholar
Jefferies, R. L. 1977. The vegetation of salt marshes at some coastal sites in arctic North America. Journal of Ecology 65: 661–672.CrossRefGoogle Scholar
Jeglum, J. K. 1983. Changes in tree species composition in naturally regenerating strip clearcuts in shallow-soil upland black spruce. pp. 180–193. In Wein, R. W., Riewe, R. R., and Methven, I. R. (eds.) Resources and Dynamics of the Boreal Zone. Ottawa: Association of Canadian Universities for Northern Studies.
Jensen, T. S. 1985. Seed–seed predator interactions of European beech, Fagus silvatica and forest rodents, Clethrionomys glareolus and Apodemus flavicollis. Oikos 44: 149–156.CrossRefGoogle Scholar
Jickells, T. D., Dorling, S., Deuser, W. G., Church, T. M., Arimoto, R., and Propsero, J. M.. 1998. Air-borne dust fluxes to a deep water sediment trap in the Sargasso Sea. Global Biogeochemical Cycles 12: 311–320.CrossRefGoogle Scholar
Johansson, M. E. and Keddy, P. A.. 1991. Intensity and asymmetry of competition between plant pairs of different degrees of similarity: an experimental study on two guilds of wetland plants. Oikos 60: 27–34.CrossRefGoogle Scholar
Johnson, P. L. and Billings, W. D.. 1962. The alpine vegetation of the Beartooth Plateau in relation to cryopedogenic processes and patterns. Ecological Monographs 32: 105–135.CrossRefGoogle Scholar
Johnston, A. E. 1994. The Rothamsted classical experiments. pp. 9–35. In R. A. Leigh and A. E. Johnston (eds.) Long-term Experiments in Agricultural and Ecological Sciences. Proceedings of a conference to celebrate the 150th anniversary of Rothamsted Experimental Station, held at Rothamsted, July 14–17, 1993. Wallingford: CAB International.
Johnston, C. A. and Naiman, R. J.. 1990. Aquatic patch creation in relation to beaver population trends. Ecology 71: 1617–1621.CrossRefGoogle Scholar
Jones, C. G., Lawton, J. H., and Shachak, M.. 1994. Organisms as ecosystem engineers. Oikos 69: 373–386.CrossRefGoogle Scholar
Jones, R. K., G. Pierpoint, G. M. Wickware, and J. K. Jeglum. 1983a. A classification and ordination of forest ecosystems in the Great Claybelt of northeastern Ontario. pp. 83–96. In R. W. Wein, R. R. Riewe, and I. R. Methven (eds.) Resources and Dynamics of the Boreal Zone. Proceedings of a Conference held at Thunder Bay, Ontario, August 1982. Ottawa: Association of Canadian Universities for Northern Studies.
Jones, R. K., Pierpoint, G., Wickware, G. M., Jeglum, J. K., Arnup, R. W., and Bowles, J. M.. 1983b. Field Guide to Forest Classification for the Clay Belt, Site Region 3E. Toronto: Queen's Printer for Ontario.Google Scholar
Jones, W. G., Hill, K. D., and Allen, J. M.. 1995. Wollemia nobilis, a new living Australian genus and species in the Araucariaceae. Telopea 6: 173–176.CrossRefGoogle Scholar
Jordan, C. F., Golley, F. B., Hall, J. D., and Hall, J.. 1980. Nutrient scavenging of rainfall by the canopy of an Amazonian rain forest. Biotropica 12: 61–66.CrossRefGoogle Scholar
Jordan, W. R. III, Gilpin, M. E., and Aber, J. D.. 1987. Restoration Ecology: A Synthetic Approach to Ecological Research. Cambridge: Cambridge University Press.Google Scholar
Judd, W. S., Campbell, C. S., Kellogg, E. A., Stevens, P. F., and Donoghue, M. J.. 2002. Plant Systematics: A Phylogenetic Approach. 2nd edn. Sunderland: Sinauer.Google Scholar
Judson, S. 1968. Erosion of the land, or what's happening to our continents?American Scientist 56: 356–374.Google Scholar
Junk, W. J. 1983. Ecology of swamps on the Middle Amazon. pp. 269–294. In Gore, A. J. P.. (ed.) Ecosystems of the World 4B: Mires: Swamp, Bog, Fen, and Moor. Amsterdam: Elsevier Science.Google Scholar
Jutila, H. M. and Grace, J. B.. 2002. Effects of disturbance on germination and seedling establishment in a coastal prairie grassland: a test of the competitive release hypothesis. Journal of Ecology 90: 291–302.CrossRefGoogle Scholar
Kalamees, K. 1982. The composition and seasonal dynamics of fungal cover on peat soils. pp. 12–29. In Masing, V. (ed.) Peatland Ecosystems: Researches into the Plant Cover of Estonian Bogs and Their Productivity. Tallinn: Academy of Sciences of the Estonian S.S.R.
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
Kaplan, D. R. and Cooke, T. J.. 1996. The genius of Wilhelm Hofmeister: The origin of causal-analytical research in plant development. American Journal of Botany 83: 1647–1660.CrossRefGoogle Scholar
Kastner, T. P., and Goñi, M. A.. 2003. Constancy in the vegetation of the Amazon Basin during the late Pleistocene: evidence from the organic matter composition of Amazon deep sea fan sediments. Geology 31: 291–294.2.0.CO;2>CrossRefGoogle Scholar
Kay, S. 1993. Factors affecting severity of deer browsing damage within coppiced woodlands in the south of England. Biological Conservation 63: 524–532.CrossRefGoogle Scholar
Kearney, T. H. and Shantz, H. L.. 1912. The water economy of dry-land crops. pp. 351–362. Yearbook of the United States Department of Agriculture-1911. Washington: Department of Agriculture.Google Scholar
Keddy, P. A. 1980. Population ecology in an environmental mosaic: Cakile edentula on a gravel bar. Canadian Journal of Botany 58: 1095–1100.CrossRefGoogle Scholar
Keddy, P. A. 1981a. Why gametophytes and sporophytes are different: form and function in a terrestrial environment. The American Naturalist 118: 452–454.CrossRefGoogle Scholar
Keddy, P. A. 1981b. Vegetation with Atlantic coastal plain affinities in Axe Lake, near Georgian Bay, Ontario. The Canadian Field Naturalist 95: 241–248.Google Scholar
Keddy, P. A. 1981c. Experimental demography of the sand dune annual, Cakile edentula, growing along an environmental gradient in Nova Scotia. Journal of Ecology 69: 615–630.CrossRefGoogle Scholar
Keddy, P. A. 1982. Population ecology on an environmental gradient: Cakile edentula on a sand dune. Oecologia 52: 348–355.CrossRefGoogle ScholarPubMed
Keddy, P. A. 1983. Shoreline vegetation in Axe Lake, Ontario: effects of exposure on zonation patterns. Ecology 64: 331–344.CrossRefGoogle Scholar
Keddy, P. A. 1987. Beyond reductionism and scholasticism in plant community ecology. Vegetatio 69: 209–211.CrossRefGoogle Scholar
Keddy, P. A. 1989. Competition.London: Chapman and Hall.CrossRefGoogle Scholar
Keddy, P. A. 1990a. The use of functional as opposed to phylogenetic systematics: a first step in predictive community ecology. pp. 387–406. In Kawano, S. (ed.) Biological Approaches and Evolutionary Trends in Plants. London: Academic Press.Google Scholar
Keddy, P. A. 1990b. Competitive hierarchies and centrifugal organization in plant communities. pp. 265–289. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Keddy, P. A. 1991. Biological monitoring and ecological prediction: from nature reserve management to national state of environment indicators. pp. 249–267. In Goldsmith, F. B. (ed.) Biological Monitoring for Conservation. London: Chapman and Hall.Google Scholar
Keddy, P. A. 1992. Assembly and response rules: two goals for predictive community ecology. Journal of Vegetation Science 3: 157–164.CrossRefGoogle Scholar
Keddy, P. 1994. Reflections on the 21st birthday of MacArthur's Geographical Ecology – applications of the Hertzprung-Russel star diagram to ecology. Trends in Ecology and Evolution 9: 231–234.CrossRefGoogle Scholar
Keddy, P. 1998. Review of Null Models in Ecology (N. J. Gotelli and G. R. Graves, 1996, Smithsonian Institution Press, Washington). The Canadian Field-Naturalist112: 752–754.
Keddy, P. A. 2000. Wetland Ecology: Principles and Conservation. Cambridge: Cambridge University Press.Google Scholar
Keddy, P. A. 2001. Competition. 2nd edn. Dordrecht: Kluwer.CrossRefGoogle Scholar
Keddy, P. A. 2004. Plants matter. Review of The Ecology Of Plants (J. Gurevitch, S. Scheiner and G. A. Fox. 2002. Sinauer Associates, Sunderland, Massachusetts). The Quarterly Review of Biology 79: 55–59.CrossRefGoogle Scholar
Keddy, P. A. 2005a. Milestones in ecological thought – a canon for plant ecology. Journal of Vegetation Science 16: 145–150.Google Scholar
Keddy, P. A. 2005b. Putting the plants back into plant ecology: six pragmatic models for understanding, conserving and restoring plant diversity. Annals of Botany 96: 177–189.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: 122–141.CrossRefGoogle Scholar
Keddy, P. A. and Drummond, C. G.. 1996. Ecological properties for the evaluation, management, and restoration of temperate deciduous forest ecosystems. Ecological Applications 6: 748–762.CrossRefGoogle Scholar
Keddy, P. A. and L. H. Fraser. 2005. Introduction: big is beautiful. pp. 1–10. In Fraser, L. H. and Keddy, P. A. (eds.) The World's Largest Wetlands: Ecology and Conservation. Cambridge: Cambridge University Press.CrossRefGoogle 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 plant communities. Oikos 49: 234–241.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., H. T. Lee, and I. C. Wisheu. 1993. Choosing indicators of ecosystem integrity: wetlands as a model system. pp. 61–79. In Woodley, S., Kay, J., and Francis, G. (eds.) Ecological Integrity and the Management of Ecosystems. Ottawa: 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 enivronments?Journal of Ecology 82: 635–643.CrossRefGoogle Scholar
Keddy, P. A., Nielsen, K., Weiher, E., and Lawson, L. R.. 2002. Relative competitive performance of 63 species of terrestrial herbaceous plants. Journal of Vegetation Science 13: 5–16.CrossRefGoogle Scholar
Keeler, K. H. 1985. Cost:benefit models of mutualism. pp. 100–127. In Boucher, D. H. (ed.) The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Keeley, J. E. 1998. CAM photosynthesis in submerged aquatic plants. The Botanical Review 64: 121–175.CrossRefGoogle Scholar
Keeley, J. E. and Rundel, P. W.. 2003. Evolution of CAM and C4 carbon-concentrating mechanisms. International Journal of Plant Science 164 (Supplement): S55–S77.CrossRefGoogle Scholar
Keeley, J. E., D. A. DeMason, R. Gonzalez, and K. R. Markham. 1994. Sediment-based carbon nutrition in tropical alpine Isoetes. pp. 167–194. In Rundel, P. W., Smith, A. P., and Meinzer, F. C. (eds.) Tropical Alpine Environments Plant Form and Function. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Keeling, C. D. and T. P. Whorf. 2005. Atmospheric CO2 records from sites in the SIO air sampling network. In Trends: A Compendium of Data on Global Change. Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, TN.
Keller, G., Adatte, T., Stinnesbeck, W. et al. 2004. Chicxulub impact predates the K-T boundary mass extinction. Proceedings of the National Academy of Sciences (of the United States of America) 101: 3753–3758.CrossRefGoogle ScholarPubMed
Kellman, M. 1985. Forest seedling establishment in Neotropical savannas: transplant experiments with Xylopia frutescens and Calophyllum brasiliense. Journal of Biogeography 12: 373–379.CrossRefGoogle Scholar
Kellman, M. and Delfosse, B.. 1993. Effect of the red land crab (Gecarcinus lateralis) on leaf litter in a tropical dry forest in Vera Cruz, Mexico. Journal of Tropical Ecology 9: 55–65.CrossRefGoogle Scholar
Kellman, M. and Kading, M.. 1992. Facilitation of tree seedling establishment in a sand dune succession. Journal of Vegetation Science 3: 679–688.CrossRefGoogle Scholar
Kellman, M. and Roulet, N.. 1990. Nutrient flux and retention in a tropical sand-dune succession. Journal of Ecology 78: 664–676.CrossRefGoogle Scholar
Kellner, L. 1963. Alexander von Humboldt. London: Oxford University Press.Google Scholar
Kenrick, P. and Crane, P. R.. 1997. The origin and early evolution of plants on land. Nature 389: 33–39.CrossRefGoogle Scholar
Kershaw, K. A. 1962. Quantitative ecological studies from Landmannahellir, Iceland. Journal of Ecology 50: 171–179.CrossRefGoogle Scholar
Kershaw, K. A. 1973. Quantitative and Dynamic Plant Ecology. 2nd edn. London: Edward Arnold.Google Scholar
Kershaw, K. A. and Looney, J. H. H.. 1985. Quantitative and Dynamic Plant Ecology. 3rd edn. Victoria: Edward Arnold.Google Scholar
Kevan, P. G. 1975. Sun-tracking solar furnaces in high arctic flowers: significance for pollination and insects. Science 189: 723–726.CrossRefGoogle ScholarPubMed
Kidston, R. and Lang, W. H.. 1921. Transactions of the Royal Society Edinburgh LII(IV): 855–902.CrossRef
Killingbeck, K. T. 1996. Nutrients in senesced leaves: keys to the search for potential resorption and resorption efficiency. Ecology 77: 1716–1727.CrossRefGoogle Scholar
King, J. 1997. Reaching for the Sun: How Plants Work. New York: Cambridge University Press.Google Scholar
Kinzig, A. P., Pacala, S., and Tilman, G. D. (eds.) 2002. The Functional Consequences of Biodiversity: Empirical Progress and Theoretical Extensions. Princeton: Princeton University Press.Google Scholar
Knoll, A. H. 1992. The early evolution of eukaryotes: a geological perspective. Science 256: 622–627.CrossRefGoogle ScholarPubMed
Koerselman, W. and Meulman, A. F. M.. 1996. The vegetation N:P ratio: a new tool to detect the nature of nutrient limitation. Journal of Applied Ecology 33: 1441–1450.CrossRefGoogle Scholar
Koyama, H. and Kira, T.. 1956. Intraspecific competition among higher plants. VIII. Frequency distributions of individual plant weight as affected by the interaction between plants. Journal of the Institute of Polytechnics, Osaka City University Series D 7: 73–94.Google Scholar
Kozlowski, T. T. (ed.) 1984. Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Kozlowski, T. T. and S. G. Pallardy. 1984. Effect of flooding on water, carbohydrate, and mineral relations. pp. 165–193. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Kramer, P. J. 1983. Water Relations of Plants. Orlando: Academic Press.Google Scholar
Krebs, C. J. 1978. Ecology: The Experimental Analysis of Distribution and Abundance. New York: Harper and Row.Google Scholar
Krebs, C. J. 1989. Ecological Methodology. New York: Harper and Row.Google Scholar
Kruckeberg, A. R. 1954. The ecology of serpentine soils. III. Plant species in relation to serpentine soils. Ecology 35: 267–274.Google Scholar
Küchler, A. W. 1949. A physiognomic classification of vegetation. Annals of the Association of American Geographers 39: 201–210.CrossRefGoogle Scholar
Küchler, A. W. 1966. Analyzing the physiognomy and structure of vegetation. Annals of the Association of American Geographers 56: 112–127.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
Kuijt, J. 1969. The Biology of Parasitic Flowering Plants. Berkeley: University of California Press.Google Scholar
Kyte, F. T. 1998. A meteorite from the Cretaceous/Tertiary boundary. Nature 396: 237–239.CrossRefGoogle Scholar
Lack, D. 1947. Darwin's Finches: An Essay on the General Biological Theory of Evolution. New York: Harper and Row.Google Scholar
Laing, H. E. 1940. Respiration of the rhizomes of Nuphar advenum and other water plants. The American Journal of Botany 27: 574–581.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: 712–724.CrossRefGoogle Scholar
Langer, P. 1974. Stomach evolution in the artiodactyla. Mammalia 38: 295–314.CrossRefGoogle Scholar
Larcher, W. 1995. Physiological Plant Ecology: Ecophysiology and Stress Physiology of Functional Groups. 3rd edn. New York: Springer-Verlag.CrossRefGoogle Scholar
Larcher, W. 2003. Physiological Plant Ecology: Ecophysiology and Stress Physiology of Functional Groups. 4th edn. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Larcher, W. and H. Bauer. 1981. Ecological significance of resistance to low temperature. pp. 403–437. In Lange, O. L., Nobel, P. S., Osmond, C. B., and Ziegler, H. (eds.) Physiological Plant Ecology I: Responses to the Physical Environment. Encyclopedia of Plant Physiology: New Series, Vol. 12A. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Larson, D. W. 1980. Patterns of species distribution in an Umbilicaria-dominated community. Canadian Journal of Botany 58: 1269–1279.CrossRefGoogle Scholar
Larson, D. W. 1982. Environmental stress and Umbilicaria lichens: the effect of subzero temperature pretreatments. Oecologia 55: 268–278.CrossRefGoogle ScholarPubMed
Larson, D. W. 1989. The impact of ten years at − 20 °C on gas exchange in five lichen species. Oecologia 78: 87–92.CrossRefGoogle ScholarPubMed
Larson, D. W. 1996. Brown's Woods: an early gravel pit forest restoration project, Ontario, Canada. Society for Ecological Restoration 4: 11–18.CrossRefGoogle Scholar
Larson, D. W. 2001. The paradox of great longevity in a short-lived tree species. Experimental Gerontology 36: 651–673.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 Savanna River. Wetlands 14: 174–183.CrossRefGoogle Scholar
Latham, R. E. and Ricklefs, R. E.. 1993a. Global patterns of tree species richness in moist forests: energy-diversity theory does not account for variation in species richness. Oikos 67: 325–333.CrossRefGoogle Scholar
Latham, R. E. and R. E. Ricklefs. 1993b. Continental comparisons of temperate-zone tree species diversity. pp. 294–314. In Ricklefs, R. E. and Schluter, D. (eds.) Species Diversity in Ecological Communities: Historical and Geographical Perspectives. Chicago: The University of Chicago Press.Google Scholar
Latham, R. E., Beyea, J., Benner, M., Dunn, C. A., Fajvan, M. A., Freed, R. R., Grund, M., Horsley, S. B., Rhoads, A. F., and Shissler, B. P.. 2005. Managing White-tailed Deer in Forest Habitat from an Ecosystem Perspective: Pennsylvania Case Study. Harrisburg: Audubon Pennsylvania and Pennsylvania Habitat Alliance.Google Scholar
Lavoisier, A. L. 1789. Elements of Chemistry. Translated by R. Kerr and reprinted in xi, xii and pp. 1–60. In Adler, M. J. (ed.) 1990. Great Books of the Western World. 2nd edn., Vol. 42. Chicago: Encyclopaedia Britannica.Google Scholar
Lechowicz, M. J. 1981. The effects of climatic pattern on lichen productivity: Cetraria cucullata (Bell.) Ach. in the arctic tundra of northern Alaska. Oecologia 50: 210–216.CrossRefGoogle ScholarPubMed
Leck, M. A. and Graveline, K. J.. 1979. The seed bank of a freshwater tidal marsh. American Journal of Botany 66: 1006–1015.CrossRefGoogle Scholar
Leck, M. A., Parker, V. T., and Simpson, R. L. (eds.) 1989 Ecology of Soil Seed Banks. San Diego: Academic Press.Google Scholar
Lee, K. E. 1985. Earthworms: Their Ecology and Relationships with Soils and Land Use. Sydney: Academic Press.Google Scholar
Legendre, L. and Legendre, P.. 1983. Numerical Ecology. Amsterdam: Elsevier.Google Scholar
Leopold, A. 1949. A Sand County Almanac. London: Oxford University Press.Google Scholar
Page, C. and Keddy, P. A.. 1988. Reserves of buried seeds in beaver ponds. Wetlands 18: 242–248.CrossRefGoogle Scholar
Page, C. and Keddy, P. A.. 1998. Reserves of buried seeds in beaver ponds. Wetlands 18: 242–248.CrossRefGoogle Scholar
Levin, H. L. 1994. The Earth Through Time. 4th edn., updated. Fort Worth: Saunders College Publishing; Harcourt Brace College Publishers.Google Scholar
Levins, R. 1968. Evolution in Changing Environments. Princeton: Princeton University Press.Google Scholar
Levitt, J. 1977. The nature of stress injury and resistance. pp. 11–21. In Levitt, J. (ed.) Responses of Plants to Environmental Stress. New York: Academic Press.Google Scholar
Levitt, J. 1980. Responses of Plants to Environmental Stresses, Vols. I and II. 2nd edn. New York: Academic Press.Google Scholar
Lewis, D. H. 1987. Evolutionary aspects of mutualistic associations between fungi and photosynthetic organisms. pp. 161–178. In Rayner, A. D. M., Brasier, C. M., and Moore, D. (eds.) Evolutionary Biology of the Fungi. Symposium of the British Mycological Society, held at the University of Bristol, April 1986. Cambridge: Cambridge University Press.Google Scholar
Lewis, R. R. III (ed.) 1982. Creation and Restoration of Coastal Plant Communities. Boca Raton: CRC Press.Google Scholar
Leyser, O. and Fitter, A. 1998. Roots are branching out in patches. Trends in Plant Science 3: 203–204.CrossRefGoogle Scholar
Li, X.-L., George, E., and Marschner, H.. 1991. Phosphorus depletion and pH decrease at the root-soil and hyphae-soil interfaces of VA mycorrhizal white clover fertilized with ammonium. New Phytologist 119: 397–404.CrossRefGoogle Scholar
Lieth, H. 1975. Historical survey of primary productivity research. pp. 7–16. In Leith, H. and Whittaker, R. H. (eds.) Primary Productivity of the Biosphere. New York: Springer-Verlag.CrossRefGoogle Scholar
Likens, G. E., Bormann, F. H., Pierce, R. S., Eaton, J. S., and Johnson, N. M.. 1977. Biogeochemistry of a Forested Ecosystem. New York: Springer-Verlag.CrossRefGoogle Scholar
Little, C. E. 1995. The Dying of the Trees: The Pandemic in America's Forests. New York: Penguin Books.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: 1446–1455.CrossRefGoogle Scholar
Lloyd, D. G. and Barrett, S. C. H. (eds.) 1996. Floral Biology: Studies on Floral Evolution in Animal-Pollinated Plants. London: Chapman and Hall.CrossRefGoogle Scholar
Lodge, D. M. 1991. Herbivory on freshwater macrophytes. Aquatic Botany 41: 195–224.CrossRefGoogle Scholar
Loehle, C. 1998a. Height growth rate tradeoffs determine northern and southern range limits for trees. Journal of Biogeography 25: 735–742.CrossRefGoogle Scholar
Loehle, C. 1988b. Problems with the triangular model for representing plant strategies. Ecology 69: 284–286.CrossRefGoogle Scholar
Loehle, C. 1995. Anomalous responses of plants to CO2 enrichment. Oikos 73: 181–187.CrossRefGoogle Scholar
Lopoukhine, N. 1983. Parks Canada in the boreal forest ecosystem (a pilgrim's progress). pp. 167–179. In Wein, R. W., Riewe, R. R., and Methven, I. R. (eds.) Resources and Dynamics of the Boreal Zone. Proceedings of a Conference held at Thunder Bay, Ontario, August 1982. Ottawa: Association of Canadian Universities for Northern Studies.Google Scholar
Loreau, M. L. 2000. Biodiversity and ecosystem functioning: recent theoretical advances. Oikos 91: 3–17.CrossRefGoogle Scholar
Louda, S. M. and S. Mole. 1991. Glucosinolates: chemistry and ecology. pp. 124–164. In Rosenthal, G. A. and Berenbaum, M. R. (eds.) Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego: Academic Press.Google Scholar
Louda, S. M., K. H. Keller, and R. D. Holt. 1990. Herbivore influence on plant performance and competitive interactions. pp. 413–444. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competiton. San Diego: Academic Press.Google Scholar
Lovejoy, T. E., R. O. Bierregaard, Jr., A. B. Rylands, J. R. Malcolm, C. E. Quintela, L. H. Harper, K. S. Brown, Jr., A. H. Powell, G. V. N. Powell, H. O. R. Schubart, and M. B. Hays. 1986. Edge and other effects of isolation on Amazon forest fragments. pp. 257–285. In Soulé, M. E. (ed.) Conservation Biology; the Science of Scarcity and Diversity. Sunderland: Sinauer Associates.Google Scholar
Loveless, C. M. 1959. A study of the vegetation in the Florida Everglades. Ecology 40: 1–9.CrossRefGoogle Scholar
Lowman, M. D. 1992. Leaf growth dynamics and herbivory in five species of Australian rain-forest canopy trees. Journal of Ecology 80: 433–447.CrossRefGoogle Scholar
Lowman, M. D. and Rinker, H. B. (eds). 2004. Forest Canopies. 2nd edn. Burlington: Elsevier Academic Press.Google Scholar
Ludwig, D., Jones, D. D., and Holling, C. S.. 1978. Qualitative analysis of insect outbreak systems: the spruce budworm and forest. Journal of Animal Ecology 47: 315–332.CrossRefGoogle Scholar
Lugo, A. E. and Snedaker, S. C.. 1974. The ecology of mangroves. Annual Review of Ecology and Systematics 5: 39–64.CrossRefGoogle Scholar
Lutman, J. 1978. The role of slugs in an Agrostis–Festuca grassland. pp. 332–347. In Heal, O. W. and Perkins, D. F. (eds.) Production Ecology of British Moors and Montane Grasslands. Ecological Studies, Vol. 27. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Mabry, C. M. 2004. The number and size of seeds in common versus restricted woodland herbaceous species in central Iowa, USA. Oikos 107: 497–504.CrossRefGoogle Scholar
MacArthur, R. H. 1957. On the relative abundance of bird species. Proceedings of the National Academy of Sciences of the USA 43: 293–295.CrossRefGoogle ScholarPubMed
MacArthur, R. H. 1972. Geographical Ecology. New York: Harper and Row.Google Scholar
MacArthur, R. H. and Wilson, E. O.. 1967. The Theory of Island Biogeography. Monographs in Population Biology, No. 1. Princeton: Princeton University Press.Google Scholar
MacDonald, P. (ed.) 1989. The Solar System. The World of Science, Vol. 7. Oxford: Equinox (Oxford) Ltd.Google Scholar
MacFarland, C. G., Villa, J., and Toro, B.. 1974. The Galápagos giant tortoises (Geochelone elephantopus). Part I: Status of the surviving populations. Biological Conservation 6: 118–133.CrossRefGoogle Scholar
MacGillivray, C. W., Grime, J. P. and the Integrated Screening Programme (ISP) team. 1995. Testing predictions of the resistance and resilience of vegetation subjected to extreme events. Functional Ecology 9: 640–649.CrossRefGoogle Scholar
MacMahon, J. A. 1981. Successional processes: comparisons among biomes with special reference to probable roles of and influence on animals. pp. 277–305. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
Magnuson, J. J., H. A. Regier, W. J. Christie, and W. C. Sonzongi. 1980. To rehabilitate and restore Great Lakes ecosystems. pp. 95–122. In Cairns, J. Jr. (ed.) The Recovery Process in Damaged Ecosystmens. Ann Arbour: Ann Arbour Science Publishers.Google Scholar
Magnusson, M. (ed.) 1990. Chambers Biographical Dictionary. 5th edn. Edinburgh: W & R Chambers.Google Scholar
Mains, G. 1972. The Oxygen Revolution. London: David and Charles.Google Scholar
Major, J. 1988. Endemism: a botanical perspective. pp. 117–146. In Myers, A. A. and Giller, P. S. (eds.) Analytical Biogeography. London: Chapman and Hall.CrossRefGoogle Scholar
Margulis, L. 1970. Origin of Eukaryotic Cells. New Haven: Yale University Press.Google Scholar
Margulis, L. 1993. Symbiosis in Cell Evolution. 2nd edn. New York: W. H. Freeman.Google Scholar
Margulis, L. and Sagan, D.. 1986. Microcosmos: Four Billion Years of Evolution from Our Microbial Ancestors. Reprinted in 1997 in paperback. Berkeley: University of California Press.Google Scholar
Marquis, R. 1991. Evolution of resistance in plants to herbivores. Evolutionary Trends in Plants 5: 23–29.Google Scholar
Marquis, R. and Whelan, C.. 1994. Insectivorous birds increase growth of white oak through consumption of leaf-chewing insects. Ecology 75: 2007–2014.CrossRefGoogle Scholar
Marschner, H. 1995. Mineral Nutrition of Higher Plants. 2nd edn. London: Academic Press.Google Scholar
Martin, J. H.et al. 1994. Testing the iron hypothesis in ecosystems of the equatorial Pacific Ocean. Nature 371: 123–129.CrossRefGoogle Scholar
Martin, P. S. and Klein, R. J.. 1984. Quaternary Extinctions: A Prehistoric Revolution. Tucson: The University of Arizona Press.Google Scholar
Marx, K. 1867. In Engles, F. (ed.) Capital. Translated from 3rd German edition by S. Moore and E. Aveling. Revised from 4th edition by M. Sachey and H. Lamm. pp. 1–411. In Adler, M. J. (ed.) Great Books of the Western World. 2nd edn. 1990. Vol. 50. Chicago: Encyclopaedia Britannica.Google Scholar
Matson, P. A., Lohse, K., and Hall, S.. 2002. The globalization of nitrogen deposition: consequences for terrestrial ecosystems. Ambio 31: 113–119.CrossRefGoogle ScholarPubMed
Matthes-Sears, U., Gerrath, J., and Larson, D.. 1997. Abundance, biomass, and productivity of endolithic and epilithic lower plants on the temperate-zone cliffs of the Niagara Escarpment, Canada. International Journal of Plant Science 158: 451–460.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: 450–455.CrossRefGoogle Scholar
May, E. 1982. Budworm Battles: The Fight to Stop the Aerial Insecticide Spraying of the Forests of Eastern Canada. Halifax: Four East Publications Ltd.Google Scholar
May, R. M. 1973. Stability and Complexity in Model Ecosystems. Princeton: Princeton University Press.Google ScholarPubMed
May, R. M. 1977. Thresholds and breakpoints in ecosystems with a multiplicity of stable states. Nature 269: 471–477.CrossRefGoogle Scholar
May, R. M. 1981. Patterns in multi-species communities. pp. 197–227. In May, R. M. (ed.) Theoretical Ecology. Oxford: Blackwell.Google Scholar
May, R. M. 1988. How many species are there on Earth?Science 241: 1441–1449.CrossRefGoogle ScholarPubMed
Smith, Maynard J. 1978. The Evolution of Sex. Cambridge: Cambridge University Press.Google Scholar
Mayr, E. 1982. The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Cambridge: Belknap Press of Harvard University Press.Google 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: 373–381.CrossRefGoogle Scholar
McCarthy, K. A. 1987. Spatial and temporal distributions of species in two intermittent-tent ponds in Atlantic county, N. J. MSc thesis. New Brunswick: Rutgers University.
McClaran, M. P. 2003. A century of vegetation change on the Santa Rita Experimental Range. pp. 16–33. In USDA Forest Service Proceedings RMRS-P-30. US Department of Agriculture.
McClure, J. W. 1970. Secondary constituents of aquatic angiosperms. pp. 233–265. In Harborne, J. B. (ed.) Phytochemical Phylogeny. New York: Academic Press.Google Scholar
McComb, W. C. and Muller, R. N.. 1983. Snag management in old growth and second-growth Appalachian forests. Journal of Wildlife Management 47: 376–382.CrossRefGoogle Scholar
McGraw, J. B. 2001. Evidence for decline in stature of American ginseng plants from herbarium specimens. Biological Conservation 98: 25–32.CrossRefGoogle Scholar
McGraw, J. B. and Furedi, M. A.. 2005. Deer browsing and population viability of a forest understory plant. Science 307: 920–922.CrossRefGoogle ScholarPubMed
McIntosh, R. P. 1967. The continuum concept of vegetation. The Botanical Review 33: 130–187.CrossRefGoogle Scholar
McIntosh, R. P. 1981. Succession and ecological theory. pp. 10–23. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
McIntosh, R. P. 1985. The Background of Ecology: Concept and Theory. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
McKenzie, D. H., Hyatt, D. E., and McDonald, V. J.. 1992. Ecological Indicators, Vols. 1 and 2. London: Elsevier.Google Scholar
McNaughton, S. J. 1985. Ecology of a grazing ecosystem: the Serengeti. Ecological Monographs 55: 259–294.CrossRefGoogle Scholar
McNaughton, S. J., Ruess, R. W., and Seagle, S. W.. 1988. Large mammal and process dynamics in African ecosystems. Bioscience 38: 794–800.CrossRefGoogle Scholar
McNeill, J. R. and Winiwarter, V.. 2004. Breaking the sod: humankind, history and soil. Science 304: 1627–1629.CrossRefGoogle Scholar
McShea, W. J. and Rappole, J. H.. 2000. Managing the abundance and diversity of breeding bird populations through manipulation of deer populations. Conservation Biology 14: 1161–1170.CrossRefGoogle Scholar
McVaugh, R. 1943. The vegetation of the granitic flat-rocks of the southeastern United States. Ecological Monographs 13: 119–166.CrossRefGoogle Scholar
Meadows, D. H., Meadows, D. L., Randers, J., and Behrens, W. W. III. 1974. The Limits to Growth: A Report for the Club of Rome's Project on the Predicament of Mankind. 2nd edn. New York: The New American Library.Google Scholar
Meave, J. and Kellman, M.. 1994. Maintenance of rain forest diversity in riparian forest of tropical savannas: implications for species conservation during Pleistocene drought. Journal of Biogeography 21: 121–135.CrossRefGoogle Scholar
Merrens, E. J. and Peart, D. R.. 1992. Effects of hurricane damage on individual growth and stand structure in a hardwood forest in New Hampshire, USA. Journal of Ecology 80: 787–795.CrossRefGoogle Scholar
Milchunas, D. G. and Lauenroth, W. K.. 1993. Quantitative effects of grazing on vegetation and soils over a global range of environments. Ecological Monographs 63: 327–366.CrossRefGoogle Scholar
Milchunas, D. G., Laurenroth, W. K., and Burk, I. C.. 1998. Livestock grazing: animal and plant biodiversity of shortgrass steppe and the relationship to ecosystem function. Oikos 83: 65–74.CrossRefGoogle Scholar
Miller, G. R. and A. Watson. 1978. Heather productivity and its relevance to the regulation of red grouse populations. pp. 277–285. In Heal, O. W., Perkins, D. F. and Brown, W. M. (eds.) Production Ecology of British Moors and Montaine Grasslands. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Miller, G. R. and A. Watson. 1983. Heather moorland in northern Britain. pp. 101–117. In Warren, A. and Goldsmith, F. B. (eds.) Conservation in Perspective. Chichester: John Wiley and Sons Ltd.Google Scholar
Miller, K. G., Kominz, M. A., Browning, J. V., Wright, J. D., Mountain, G. S., Katz, M. E., Sugarman, P. J., Cramer, B. S., Christie-Blick, N., and Pekar, S. F.. 2005. The Phanerozoic record of global sea-level changes. Science 310: 1293–1298.CrossRefGoogle Scholar
Miller, R. S. 1967. Pattern and process in competition. Advances in Ecological Research 4: 1–74.CrossRefGoogle Scholar
Miller, S. L. 1953. A production of amino acids under possible primitive earth conditions. Science 117: 528–529.CrossRefGoogle ScholarPubMed
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
Milton, S. J., Dean, W. R. J., Plessis, M. A., and Siegfried, W. R.. 1994. A conceptual model of arid rangeland degradation. Bioscience 44: 70–76.CrossRefGoogle Scholar
Mitchley, J. 1988. Control of relative abundance of perennials in chalk grassland in southern England. II. Vertical canopy structure. Journal of Ecology 76: 341–350.CrossRefGoogle Scholar
Mitchley, J. and Grubb, P. J.. 1986. Control of relative abundance of perennials in chalk grassland in southern England. I. Constancy of rank order and results of pot- and field-experiments on the role of interference. Journal of Ecology 74: 1139–1166.CrossRefGoogle Scholar
Mitsch, W. J. and Gosselink, J. G.. 1986. Wetlands. New York: Van Nostrand Reinhold.Google Scholar
Mitsch, W. J. and Gosselink, J. G.. 2000. Wetlands. 3rd edn. New York: John Wiley & Sons.Google Scholar
Mitsch, W. J. and S. E. Jørgensen. 1990. Modelling and management. pp. 727–744. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 1. The Hague: SPB Academic Publishing.Google Scholar
Moffett, M. W. 1994. The High Frontier: Exploring the Tropical Rainforest Canopy. Cambridge: Harvard University Press.Google Scholar
Moles, A. T., Ackerly, D. D., Webb, C. O., Tweddle, J. C., Dickie, J. B., and Westoby, M.. 2005. A brief history of seed size. Science 307: 576–580.CrossRefGoogle ScholarPubMed
Molisch, H. 1937. Der Einfluss einer Pflanze auf die andere. Allelopathie. Jena: Gustav Fischer.Google Scholar
Moolman, H. J. and Cowling, R. M.. 1994. The impact of elephant and goat grazing on the endemic flora of South African succulent thicket. Biological Conservation 68: 53–61.CrossRefGoogle Scholar
Mooney, H. A. and Dunn, E. L.. 1970. Convergent evolution of mediterranean-climate evergreen sclerophyll shrubs. Evolution 24: 292–303.CrossRefGoogle ScholarPubMed
Mooney, H. A., Drake, B. G., Luxmoore, R. J., Oechel, W. C., and Pitelka, L. F.. 1991. Predicting ecosystem responses to elevated CO2 concentrations. Bioscience 41: 96–104.CrossRefGoogle Scholar
Moore, B. and Bolin, B.. 1987. The oceans, carbon dioxide and global climate change. Oceanus 29: 9–15.Google Scholar
Moore, D. R. J. 1990. Pattern and process in wetlands of varying standing crop: the importance of scale. PhD thesis. Ottawa: University of Ottawa.
Moore, D. R. J. 1998. The ecological component of ecological risk assessment: lessons from a field experiment. Human and Ecological Risk Assessment 4: 1103–1123.CrossRefGoogle Scholar
Moore, D. R. J. and Keddy, P. A.. 1989a. 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 P. A. Keddy. 1989b. Infertile wetlands: conservation priorities and management. pp. 391–397. In Bardecki, M. J. and Patterson, N. (eds.) Wetlands: Inertia or Momentum. Proceedings of a Conference held in Toronto, Ontario, October 21–22, 1988. Federation of Ontario Naturalists.Google 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: 2408–2412.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: 203–217.CrossRefGoogle Scholar
Moore, P. D., Webb, J. A., and Collinson, M. E.. 1991. Pollen Analysis. London: Blackwell Scientific.Google Scholar
Moreno, M. T., Cubero, J. I, Berner, D., Joel, D., Musselman, L. J., and Parker, C. (eds.) 1996. Advances in Parasitic Plant Research. Cordoba: Junta de Andalucia, Dirección General de Investigación Agraria.Google Scholar
Morgan, R. and Whitaker, B. 1986. Rainbow Warrior: The French Attempt to Sink Greenpeace. London: Arrow Books Ltd.Google Scholar
Morowitz, H. J. 1968. Energy Flow in Biology: Biological Organization as a Problem in Thermal Physics. New York: Academic Press.Google Scholar
Morris, E. C. and Myerscough, P. J.. 1991. Self-thinning and competition intensity over a gradient of nutrient availability. Journal of Ecology 79: 903–923.CrossRefGoogle Scholar
Moss, B. 1983. The Norfolk Broadland: experiments in the restoration of a complex wetland. Biological Reviews of the Cambridge Philosophical Society 58: 521–561.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: 115–128.CrossRefGoogle Scholar
Mueller-Dombois, D. 1987. Natural dieback in forests. Bioscience 37: 575–583.CrossRefGoogle Scholar
Mueller-Dombois, D. and Ellenberg, H.. 1974. Aims and Methods of Vegetation Ecology. New York: John Wiley and Sons.Google Scholar
Muller, C. H. 1966. The role of chemical inhibition (allelopathy) in vegetational composition. Bulletin of the Torrey Botanical Club 93: 332–351.CrossRefGoogle Scholar
Muller, C. H. 1969. Allelopathy as a factor in ecological process. Vegetatio 18: 348–357.CrossRefGoogle 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: 232–235.CrossRefGoogle Scholar
Muller, R. A. and MacDonald, G. J.. 1997. Glacial cycles and astronomical forcing. Science 277: 215–218.CrossRefGoogle Scholar
Musselman, L. J. and Mann, W. F. Jr. 1978. Root Parasites of Southern Forests. Southern Forest Experiment Station, Forest Service, U.S. Department of Agriculture.Google Scholar
Myers, C. W. and Donnelly, M. A.. 1997. A tepui herpetofauna on a granitic mountain (Tamacuari) in the borderland between Venezuela and Brazil: report from the Phipps Tapirapecó expedition. American Museum Novitates 3213: 1–71.Google Scholar
Myers, N. 1988. Threatened biotas: “hotspots” in tropical forests. Environmentalist 8: 1–20.CrossRefGoogle Scholar
Myers, N., Mittermeier, R. A., Mittermeier, C. G., Fonseca, G. A. B., and Kent, J.. 2000. Biodiversity hotspots for conservation priorities. Nature 403: 853–858.CrossRefGoogle ScholarPubMed
Myers, R., J. O'Brien, and S. Morrison. 2006. Fire Management Overview of the Caribbean Pine (Pinus caribaea) Savannas of the Mosquitia, Honduras. GFI Technical Report 2006-1b. Arlington: The Nature Conservancy.
Nabokov, P. 1993. Long threads. pp. 301–383. In Ballantine, B. and Ballantine, I. (eds.) The Native Americans: An Illustrated History. Atlanta: Turner Publishing Inc.Google Scholar
Naeem, S. 2002. Ecosystem consequences of biodiversity loss: the evolution of a paradigm. Ecology 83: 1537–1552.CrossRefGoogle Scholar
Naeem, S. L., Thompson, J., Lawler, S. P., Lawton, J. H., and Woodfin, R. M.. 1994. Declining biodiversity can alter the performance of ecosystems. Nature 368: 734–737.CrossRefGoogle Scholar
Naiman, R. J., Johnston, C. A., and Kelley, J. C.. 1988. Alteration of North American streams by beaver. Bioscience 38: 753–762.CrossRefGoogle Scholar
Nakamura, R. P. 1980. Plant kin selection. Evolutionary Theory 5: 113–117.Google 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: 229–251.CrossRefGoogle Scholar
Nantel, P. and Neuman, P.. 1992. Ecology of ectomycorrhizal-basidiomycete communities on a local vegetation gradient. Ecology 73: 99–117.CrossRefGoogle Scholar
Nantel, P., Gagnon, D., and Nault, A.. 1996. Population viability analysis of American ginseng and wild leek harvested in stochastic environments. Conservation Biology 10: 608–621.CrossRefGoogle Scholar
National Parks and Wildlife Service. 1998. Wollemi Pine Recovery Plan. Sydney: NPWS.
Newman, E. I. 1988. Mycorrhizal links between plants: their functioning and ecological significance. Advances in Ecological Research 18: 243–270.CrossRefGoogle Scholar
Newman, E. I. 1993. Applied Ecology. Oxford: Blackwell Scientific Publications.Google Scholar
Newman, S., Grace, J. B., and Koebel, J. W.. 1996. Effects of nutrients and hydroperiod on Typha, Cladium and Eleocharis: implications of Everglades restoration. Ecological Applications 6: 774–783.CrossRefGoogle Scholar
Nicholson, A. and Keddy, P. A.. 1983. The depth profile of a shoreline seed bank in Matchedash Lake, Ontario. Canadian Journal of Botany 61: 3293–3296.CrossRefGoogle Scholar
Nickrent, D. L. 2006. The parasitic plant connection. (http://www.parasiticplants.siu/index.htm) accessed 10 Nov. 2006.
Nicolson, M. and McIntosh, R. P.. 2002. H. A. Gleason and the individualistic hypothesis revisited. Bulletin of the Ecological Society of America (April 2002): 133–142.Google Scholar
Niering, W. A. and Warren, R. S.. 1980. Vegetation patterns and processes in New England salt marshes. Bioscience 30: 301–307.CrossRefGoogle Scholar
Niklas, K. J. 1994. Predicting the height of fossil plant remains: an allometric approach to an old problem. American Journal of Botany 81: 1235–1242.CrossRefGoogle Scholar
Niklas, K. J., Tiffney, B. H., and Knoll, A. H.. 1983. Patterns in vascular land plant diversification. Nature 303: 614–616.CrossRefGoogle Scholar
Niklas, K. J., B. H. Tiffney, and A. H. Knoll. 1985. Patterns in vascular plant diversification: an analysis at the species level. pp. 97–128. In Valentine, J. W. (ed.) Phanerozoic Diversity Pattern: Profiles in Macroevolution. Princeton: Princeton University Press.Google Scholar
Nilsson, L. A. 1992. Orchid pollination biology. Trends in Ecology and Evolution 7: 255–259.CrossRefGoogle Scholar
Nobel, P. S. 1976. Water relations and photosynthesis of a desert CAM plant, Agave deserti. Plant Physiology 58: 576–582.CrossRefGoogle ScholarPubMed
Nobel, P. S. 1977. Water relations and photosynthesis of a barrel cactus, Ferocactus acanthodes, in the Colorado Desert. Oecologia 27: 117–133.CrossRefGoogle ScholarPubMed
Nobel, P. S. 1985. Desert succulents. pp. 181–197. In Chabot, B. F. and Mooney, H. A. (eds.) Physiological Ecology of North American Plant Communities. London: Chapman and Hall.CrossRefGoogle Scholar
Noble, I. R. and Slatyer, R. O.. 1980. The use of vital attributes to predict successional changes in plant communities subject to recurrent disturbances. Vegetatio 43: 5–21.CrossRefGoogle Scholar
Noss, R. 1995. Maintaining Ecological Integrity in Representative Reserve Networks. A World Wildlife Fund Canada/United States Discussion Paper, WWF.
Noss, R. F. and Cooperrider, A. Y.. 1994. Saving Nature's Legacy. Washington, D.C.: Island Press.Google Scholar
Noy-Meir, I. 1973. Desert ecosystems: environment and producers. Annual Review of Ecology and Systematics 4: 25–51.CrossRefGoogle Scholar
Noy-Meir, L. 1975. Stability of grazing systems: an application of predator–prey graphs. Journal of Ecology 63: 459–481.CrossRefGoogle Scholar
Oakes, E. H. 2002. Notable Scientists. A to Z of Chemists. New York: Facts on File.Google Scholar
Ocampo, J. A. 1986. Vesicular-arbuscular mycorrhizal infection of “host” and “non-host” plants: effect on the growth responses of the plants and competition between them. Soil Biology and Biochemistry 18: 607–610.CrossRefGoogle Scholar
Ochoa, J. G., Molina, C., and Giner, S.. 1993. Inventario y estudio comunitario de los mamiferos del Parque National Canaima, con una lista de las especies registradas para la Guayana Venezolana. Ecologia Acta Cientifíca Venezolana 44: 245–262.Google Scholar
Okihana, H. and Ponnamperuma, C.. 1982. A protective function of the coacervates against UV light on the primitive Earth. Nature 299: 347–349.CrossRefGoogle Scholar
Oksanen, L. 1990. Predation, herbivory, and plant strategies along gradients of primary production. pp. 445–474. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Oksanen, L., Fretwell, S. D., Arruda, J., and Niemelä, P.. 1981. Exploitation ecosystems in gradients of primary productivity. The American Naturalist 118: 240–261.CrossRefGoogle Scholar
Oksanen, L., M. Aunapuu, T. Oksanen, M. Schneider, P. Ekerholm, P. A. Lundberg, T. Armulik, V. Aruoja, and L. Bondestad. 1997. Outlines of food webs in a low arctic tundra landscape in relation to three theories on trophic dynamics. pp. 351–373. In Gange, A. C. and Brown, V. K. (eds.) Multitrophic Interactions in Terrestrial Systems. The 36th Symposium of The British Ecological Society. Oxford: Blackwell Science.Google Scholar
Olson, D. M.et al. 2001. Terrestrial ecoregions of the world: a new map of life on Earth. Bioscience 51: 933–938.CrossRefGoogle Scholar
Ondok, J. P. 1990. Modelling ecological processes. pp. 659–89. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 1. The Hague: SPB Academic Publishing.Google Scholar
Oosting, H. J. 1956. The Study of Plant Communities, 2nd edn. San Francisco: W. H. Freeman.Google Scholar
Oparin, A. I. 1938. The Origin of Life. New York: The Macmillan Company. Translated by S. Morgulis.Google Scholar
Orloci, L. 1978. Multivariate Analysis in Vegetation Research. 2nd edn. The Hague: Junk.Google 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: 859–869.Google Scholar
Ostrofsky, M. L. and Zettler, E. R.. 1986. Chemical defenses in aquatic plants. Journal of Ecology 74: 279–287.CrossRefGoogle Scholar
Oxford Atlas of the World. 1997. New York: Oxford University Press.
Parkinson, C. L., Adams, K. L., and Palmer, J. D.. 1999. Multigene analyses identify the three earliest lineages of extant flowering plants. Current Biology 9: 1485–1491.CrossRefGoogle ScholarPubMed
Parks, J. C. and Werth, C. R.. 1993. A study of spatial features of clones in a population of bracken fern, Pteridium aquilinum (Dennstaedtiaceae). American Journal of Botany 80: 537–544.CrossRefGoogle Scholar
Pärtel, M., Zobel, M., Zobel, K., and Maarel, E.. 1996. The species pool and its relationship to species richness: evidence from Estonia plant communities. Oikos 75: 111–117.CrossRefGoogle Scholar
Pearce, C. M. 1993. Coping with forest fragmentation in southwestern Ontario. pp. 100–113. In Poser, S. F., Crins, W. J., and Beechey, T. J. (eds.) Size and Integrity Standards for Natural Heritage Areas in Ontario. Proceedings of a Seminar. Parks and Natural Heritage Policy Branch, Ontario Ministry of Natural Resources, Queen's Printer, Toronto.Google Scholar
Pearsall, W. H. 1920. The aquatic vegetation of the English Lakes. Journal of Ecology 8: 163–201.CrossRefGoogle Scholar
Peat, H. J. and Fitter, A. H.. 1993. The distribution of arbuscular mycorrhizas in the British flora. New Phytologist 125: 845–854.CrossRefGoogle Scholar
Peattie, D. C. 1922. The Atlantic coastal plain element in the flora of the Great Lakes. Rhodora 24: 50–70, 80–88.Google 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: 1536–1545.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. 1978. Forest vegetation of the Colorado Front Range: patterns of species diversity. Vegetatio 37: 65–78.CrossRefGoogle Scholar
Peet, R. K. 1992. Community structure and ecosystem function. pp. 103–151. In Glenn-Lewin, D. C., Peet, R. K., and Veblen, T. T. (eds.) Plant Succession: Theory and Prediction. Population and Community Biology, Vol. 11. London: Chapman and Hall.Google Scholar
Peet, R. K. and D. J Allard. 1993. Longleaf pine vegetation of the southern Atlantic and eastern Gulf coast regions: a preliminary classification. pp. 45–81. In Hermann, S. M. (ed.) Proceedings of the Tall Timbers Fire Ecology Conference. No. 18. The Longleaf Pine Ecosystem: Ecology, Restoration, and Management. Florida: Tall Timbers Research Station.Google Scholar
Pennings, S. C. and Callaway, R. M.. 1996. Impact of a parasitic plant on the structure and dynamics of salt marsh vegetation. Ecology 77: 1410–1419.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: 1968–1979.CrossRefGoogle Scholar
Percival, M. S. 1965. Floral Diversity. Oxford: Pergamon Press.Google Scholar
Perry, D. 1986. Life Above the Jungle Floor. San José: Don Perro Press.Google Scholar
Peters, R. H. 1992. A Critique for Ecology. Cambridge: Cambridge University Press.Google Scholar
Petit, J. R. et al. 1999. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature 399: 429–436.CrossRefGoogle Scholar
Petterson, B. 1965. Gotland and Öland: two limestone islands compared. Acta Phytogeographic Suecica 50: 131–140.Google Scholar
Phillips, D. L. and Shure, D. J.. 1990. Patch-size effects on early succession in southern Appalachian forest. Ecology 71: 204–212.CrossRefGoogle Scholar
Phipps, R. W. 1883. On the Necessity of Preserving and Replanting Forests. Toronto: Blackett and Robinson.CrossRefGoogle Scholar
Pianka, E. R. 1981. Competition and niche theory. pp. 167–196. In May, R. M. (ed.) Theoretical Ecology. Oxford: Blackwell.Google Scholar
Pianka, E. R. 1983. Evolutionary Ecology. 3rd edn. New York: Harper and Row.Google Scholar
Pickett, S. T. A. 1980. Non-equilibrium coexistence of plants. Bulletin of the Torrey Botanical Club 107: 238–248.CrossRefGoogle Scholar
Pickett, S. T. A. and White, P. S.. 1985. The Ecology of Natural Disturbance and Patch Dynamics. Orlando: Academic Press.Google Scholar
Pielou, E. C. 1975. Ecological Diversity. New York: John Wiley and Sons.Google Scholar
Pielou, E. C. 1977. Mathematical Ecology. New York: John Wiley and Sons.Google Scholar
Pielou, E. C. 1979. Biogeography. New York: John Wiley and Sons.Google Scholar
Pielou, E. C. 1984. The Interpretation of Ecological Data: A Primer on Classification and Ordination. New York: Wiley.Google Scholar
Pielou, E. C. and Routledge, R. D.. 1976. Salt marsh vegetation: latitudinal gradients in the zonation patterns. Oecologia 24: 311–321.CrossRefGoogle ScholarPubMed
Pietropaolo, J. and Pietropaolo, P.. 1986. Carnivorous Plants of the World. Portland: Timber Press.Google Scholar
Pimm, S. L. 2001. The World According to Pimm: A Scientist Audits the Earth. New York: McGraw-Hill.Google Scholar
Pirozynski, D. W. and D. W. Malloch. 1988. Seeds, spores and stomachs: coevolution in seed dispersal mutualisms. pp. 228–244. In Pirozynski, K. A. and Hawksworth, D. L. (eds.) Coevolution of Fungi with Plants and Animals. London: Academic Press.Google Scholar
Pirozynski, K. A. and Dalpé, Y.. 1989. Geological history of the Glomaceae with particular reference to mycorrhizal symbiosis. Symbiosis 7: 1–36.Google Scholar
Pitman, N. C. A. and Jørgensen, P. M.. 2002. Estimating the size of the world's threatened flora. Science 298: 989.CrossRefGoogle ScholarPubMed
Platt, W. J. 1999. Southeastern pine savannas. pp. 23–51. In Anderson, R. C., Fralish, J. S., and Baskin, J. M. (eds.) Savannas, Barrens and Rock Outcrop Communities of North America. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Poljakoff-Mayber, A. and Gale, J.. (eds.) 1975. Plants in Saline Environments. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Ponnamperuma, F. N. 1984. Effects of flooding on soils. pp. 9–45. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Poole, R. W. and Rathcke, B. J.. 1979. Regularity, randomness, and aggregation in flowering phenologies. Science 203: 470–471.CrossRefGoogle ScholarPubMed
Porter, C. L. 1967. Taxonomy of Flowering Plants. 2nd edn. San Francisco: W. H. Freeman.Google Scholar
Porter, H. 1993. Interspecific variation in the growth response of plants to an elevated ambient CO2 concentration. Vegetatio 104/105: 77–97.CrossRefGoogle Scholar
Postel, S. L., Daily, G. C., and Ehrlich, P. R.. 1996. Human appropriation of renewable fresh water. Science 271: 785–788.CrossRefGoogle Scholar
Pound, R. 1893. Symbiosis and mutualism. The American Naturalist 27: 509–520.CrossRefGoogle Scholar
Power, M. E. 1992. Top-down and bottom-up forces in food webs: do plants have primacy?Ecology 73: 733–746.CrossRefGoogle Scholar
Prance, G. T. and Johnson, D. M.. 1991. Plant collections from the plateau of Serra do Aracá (Amazonas, Brazil) and their phytogeographic affinities. Kew Bulletin 47: 1–24.CrossRefGoogle Scholar
Press, M. C. and Graves, J. D. (eds.) 1995. Parasitic Plants. London: Chapman and Hall.Google 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: 124–128.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: 410–432.CrossRefGoogle Scholar
Price, M. V. 1983. Ecological consequences of body size: a model for patch choice in desert rodents. Oecologia 59: 384–392.CrossRefGoogle ScholarPubMed
Price, M. V. 1984. Alternative paradigms in community ecology. pp. 354–383. In Price, P. W., Slobodchikoff, C. N., and Gaud, W. S. A. (eds.) A New Ecology: Novel Approaches to Interactive Systems. New York: John Wiley and Sons.Google Scholar
Price, M. V. and Reichman, O. J.. 1987. Distribution of seeds in Sonoran Desert soils: implications for heteromyid rodent foraging. Ecology 68: 1797–1811.CrossRefGoogle ScholarPubMed
Pringle, H. 1996. In Search of Ancient North America. New York: John Wiley and Sons.Google Scholar
Pritchard, P. C. H. 1996. The Galápagos Tortoises: Nomenclatural and Survival Status. Chelonian Research Monographs No. 1. Lunenbug: Chelonian Research Foundation.Google Scholar
Puckett, L. J. 1994. Nonpoint and Point Sources of Nitrogen in Major Watersheds of the United States. U.S. Geological Survey Water-Resources Investigations Report 94–4001.
Puerto, A., Rico, M., Matias, M. D., and García, J. A.. 1990. Variation in structure and diversity in Mediterranean grasslands related to trophic status and grazing intensity. Journal of Vegetation Science 1: 445–452.CrossRefGoogle Scholar
Pusey, A. and Wolf, M.. 1996. Inbreeding avoidance in animals. Trends in Ecology and Evolution 11: 201–206.CrossRefGoogle ScholarPubMed
Putwain, P. D. and Harper, J. L.. 1970. Studies in the dynamics of plant populations. III. The influence of associated species on populations of Rumex acetosa L. and R. acetosella L. in grassland. Journal of Ecology 58: 251–264.CrossRefGoogle Scholar
Putz, F. E. and Canham, C. D.. 1992. Mechanisms of arrested succession in shrublands: root and shoot competition between shrubs and tree seedlings. Forest Ecology and Management 49: 267–275.CrossRefGoogle Scholar
Quinn, C. J. and R. A. Price. 2003. Phylogeny of the southern hemisphere conifers. pp. 129–136. In R. R. Mill (ed.) Proceedings of the 4th International Conifer Conference. ISHS Acta Horticulturae 615.
Radford, A. E., Ahles, H. E. and Bell, C. R.. 1968. Manual of the Vascular Flora of the Carolinas. Chapel Hill: The University of North Carolina Press.Google Scholar
Rapport, D. J. 1989. What constitutes ecosystem health?Perspectives in Biology and Medicine 33: 120–132.CrossRefGoogle Scholar
Rapport, D. J., Thorpe, C., and Hutchinson, T. C.. 1985. Ecosystem behaviour under stress. The American Naturalist 125: 617–640.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. 1907. The life-forms of plants and their bearing on geography. Translated from Danish and republished in 1934. In The Life Forms of Plants and Statistical Plant Geography. pp. 2–104. Oxford: Clarendon Press.Google Scholar
Raunkiaer, C. 1908. The statistics of life-forms as a basis for biological plant geography. Translated from Danish and republished in 1934. In The Life Forms of Plants and Statistical Plant Geography. pp. 111–147. Oxford: Clarendon Press.Google Scholar
Raunkiaer, C. 1918. On the biological normal spectrum. Translated from German and republished in 1934 in The Life Forms of Plants and Statistical Plant Geography. pp. 425–434. Oxford: Clarendon Press.Google Scholar
Raunkiaer, C. 1934. The Life Forms of Plants and Statistical Plant Geography: Being the Collected Papers of Raunkiaer. Translated from the Danish, French and German. Preface by A. G. Tansley. Oxford: Clarendon Press.Google Scholar
Raven, P. H., Evert, R. F., and Eichhorn, S. E.. 1999. Biology of Plants. 6th edn. New York: W. H. Freeman and Company/Worth Publishers.Google Scholar
Raven, P. H., Evert, R. F. and Eichhorn, S. E.. 2005. Biology of Plants. 7th edn. New York: W. H. Freeman and Company Publishers.Google Scholar
Ravera, O. 1989. Lake ecosystem degradation and recovery studied by the enclosure method. pp. 217–243. In Ravera, O. (ed.) Ecological Assessment of Environmental Degradation, Pollution and Recovery. Amsterdam: Elsevier Science Publishers.Google Scholar
Rawes, M. and O. W. Heal. 1978. The blanket bog as part of a Pennine moorland. pp. 224–243. In Heal, O. W. and Perkins, D. F. (eds.) Production Ecology of British Moors and Montane Grasslands. Ecological Studies, Vol. 27. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Read, D. J., Koucheki, H. K., and Hodgson, J.. 1976. Vesicular-arbuscular mycorrhizae in natural vegetation systems. 1. The occurrence of infection. New Phytologist 77: 641–653.CrossRefGoogle Scholar
Reader, R. J. and Best, B. J.. 1989. Variation in competition along an environmental gradient: Hieracium floribundum in an abandoned pasture. Journal of Ecology 77: 673–684.CrossRefGoogle Scholar
Rees, W. E. and M. Wackernagel. 1994. Ecological footprints and appropriated carrying capacity: measuring the natural capital requirements of the human economy. pp. 362–390. In Jansson, A., Hammer, M., Folke, C., and Costanza, R. (eds.) Investing in Natural Capital: The Ecological Economics Approach to Sustainability. Washington, D.C.: Island Press.Google Scholar
Regal, P. J. 1977. Ecology and evolution of flowering plant dominance. Science 196: 622–629.CrossRefGoogle ScholarPubMed
Reid, D. M. and K. J. Bradford. 1984. Effect of flooding on hormone relations. pp. 195–219. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Reid, W. V., McNeely, J. A., Tunstall, D. B., Bryant, D. A., and Winograd, M.. 1993. Biodiversity Indicators for Policymakers. Washington, D.C.: World Resources Institute.Google Scholar
Richards, P. W. 1952. The Tropical Rain Forest: An Ecological Study. Paperback edition 1979. Cambridge: Cambridge University Press.Google Scholar
Richardson, C. J. (ed.) 1981. Pocosin Wetlands: An Integrated Analysis of Coastal Plain Freshwater Bogs in North Carolina. Stroudsburg: Hutchinson Ross Publishing Company.Google Scholar
Richardson, S. J., Peltzer, D. A., Allen, R. B., and McGlone, M. S.. 2005. Resorption proficiency along a chronosequence: responses among communities and within species. Ecology 80: 20–25.CrossRefGoogle Scholar
Rickerl, D. H., Sancho, F. O., and Ananth, S.. 1994. Vesicular-arbuscular endomycorrhizal colonization of wetland plants. Journal of Environmental Quality 23: 913–916.CrossRefGoogle Scholar
Rigler, F. H. 1982. Recognition of the possible: an advantage of empiricism in ecology. Canadian Journal of Fisheries and Aquatic Sciences 39: 1323–1331.CrossRefGoogle Scholar
Rigler, F. H. and Peters, R. H.. 1995. Science and Limnology. Oldendorf/Lutie: Ecology Institute.Google Scholar
Ritchie, J. C. 1987. Postglacial Vegetation of Canada. New York: Cambridge University Press.Google Scholar
Roberts, B. A. 1992. The serpentinized areas of Newfoundland, Canada: a brief review of their soils and vegetation. pp. 53–66. In Baker, A. J. M., Proctor, J., and Reeves, R. D. (eds.) The Vegetation of Ultramafic (Serpentine) Soils. Andover: Intercept.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: 117–127.CrossRefGoogle Scholar
Robinson, A. R. 1973. Sediment, our greatest pollutant? In Tank, R. W. (ed.) Focus on Environmental Geology. London: University Press.Google Scholar
Robinson, D. 1996. Resource capture by localised root proliferation: why do plants bother?Annals of Botany 77: 179–185.CrossRefGoogle Scholar
Robinson, J. M. 1990. Lignin, land plants, and fungi: biological evolution affecting phanerozoic oxygen balance. Geology 18: 607–610.2.3.CO;2>CrossRefGoogle Scholar
Rodman, J. E. 1974. Systematics and evolution of the genus Cakile (Cruciferae). Contributions from the Gray Herbarium 205: 3–146.Google Scholar
Rodrigues, A. S. L.et al. 2004. Effectiveness of the global protected areas network in representing species diversity. Nature 428: 640–643.CrossRefGoogle ScholarPubMed
Roger, A. J. 1999. Reconstructing early events in eukaryotic evolution. The American Naturalist 154 (Suppl.): S146–S163.CrossRefGoogle ScholarPubMed
Rohde, K. 1997. The larger area of tropics does not explain latitudinal gradients in species diversity. Oikos 79: 169–172.CrossRefGoogle Scholar
Rolston, H. 1994. Foreword. pp. xi–xiii. In L. Westra (ed.) An Environmental Proposal for Ethics: The Principle of Integrity. Lanham: Rowman and Littlefield. In R. Noss (ed.) 1995. Maintaining Ecological Integrity in Representative Reserve Networks. A World Wildlife Fund Canada/World Wildlife Fund/United States Discussion Paper, WWF.
Rose, R. 1924. Man and the Galapagos Islands. pp. 332–417. In Beebe, W. (ed.) Galapagos: World's End. New York: Putnam's Sons.Google Scholar
Rosenthal, G. A. and Berenbaum, M. R.. (eds.) 1991. Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego: Academic Press.Google Scholar
Rosenzweig, M. L. 1995. Species Diversity in Space and Time. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Rosgen, D. L. 1995. River restoration utilizing natural stability concepts. pp. 55–62. In Kusler, J. A., Willard, D. E., and Hull, H. C. Jr. (eds.) Wetlands and Watershed Management: Science Applications and Public Policy. A collection of papers from a national symposium and several workshops at Tampa, Florida, April 23–26. New York: The Association of State Wetland Managers.Google Scholar
Rowe, J. S. and Scotter, G. W.. 1973. Fire in the boreal forest. Quaternary Research 3: 444–464.CrossRefGoogle Scholar
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: 564–569.CrossRefGoogle Scholar
Rundel, P. W., Cowling, R. M., Esler, K. J., Mustart, P. M., Jaarsveld, E., and Bezuidenhout, H.. 1995. Winter growth phenology and leaf orientation in Pachypodium namaquanum (Apocynaceae) in the succulent karoo of the Richtersveld, South Africa. Oecologia 101: 472–477.CrossRefGoogle ScholarPubMed
Russell, F. L., Zippin, D. B., and Fowler, N. L.. 2001. Effects of white-tailed deer (Odocoileus virginianus) on plants, plant populations and communities: a review. The American Midland Naturalist 146: 1–26.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: 233–240.CrossRefGoogle Scholar
Salisbury, E. J. 1942. The Reproductive Capacity of Plants. Studies in Quantitative Biology. London: G. Bell and Sons Ltd.Google Scholar
Salisbury, F. B. and Ross, C. W.. 1988. Plant Physiology. 3rd edn. Belmont: Wadsworth Publishers.Google Scholar
Salisbury, S. E. 1970. The pioneer vegetation of exposed muds and its biological features. Royal Society of London, Philosophical Transactions, Series B, 259: 207–255.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: 254–258.CrossRefGoogle Scholar
Salzman, A. G. and Parker, M. A.. 1985. Neighbours ameliorate local salinity stress for a rhizomatous plant in a heterogeneous environment. Oecologia 65: 273–277.CrossRefGoogle Scholar
Sandars, N. K. 1972. The Epic of Gilgamesh. An English version with an introduction by N. K. Sanders. Revised edition. London: Penguin Books.Google Scholar
Sarthou, C. and Villiers, J-F.. 1998. Epilithic plant communities on inselbergs in French Guiana. Journal of Vegetation Science 9: 847–860.CrossRefGoogle Scholar
Sather, J. H. and R. D. Smith. 1984. An Overview of Major Wetland Functions. U. S. Fish and Wildlife Service. FWS/OBS-84/18.
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: Surrey Beatty and Sons Pty Limited.Google Scholar
Savile, D. B. O. 1956. Known dispersal rates and migratory potentials as clues to the origin of the North American biota. The American Midland Naturalist 56: 434–453.CrossRefGoogle Scholar
Savile, D. B. O. 1972. Arctic Adaptations in Plants. Monograph No. 6. Ottawa: Canada Department of Agriculture.Google Scholar
Sayre, N. F. 2003. Recognizing history in range ecology: 100 years of science and management on the Santa Rita Experimental Range. pp. 1–15. In USDA Forest Service Proceedings RMRS-P-30. US Department of Agriculture.
Scagel, R. F., Rouse, G. E., Stein, J. R., Bandoni, R. J., Schofield, W. B., and Taylor, T. M. C.. 1965. An Evolutionary Survey of the Plant Kingdom. Belmont: Wadsworth Publishing Company.Google Scholar
Scagel, R. F., Bandoni, R. J., Rouse, G. E., Schofield, W. B., Stein, J. R., and Taylor, T. M. C.. 1969. Plant Diversity: An Evolutionary Approach. Belmont: Wadsworth Publishing Company.Google Scholar
Schaal, B. A. 1984. Life-history variation, natural selection, and maternal effects in plant populations. pp. 188–206. In Dirzo, R. and Sarukhán, J. (eds.) Perspectives on Plant Population Ecology. Sunderland: Sinauer Associates.Google Scholar
Schell, J. 1982. The Fate of the Earth. New York: Alfred A. Knopf.Google Scholar
Schnitzler, A. 1995. Successional status of trees in gallery forest along the river Rhine. Journal of Vegetation Science 6: 479–486.CrossRefGoogle Scholar
Scholander, P. F., Hammel, H. T., Bradstreet, B. D., and Hemmingsen, E. A.. 1965. Sap pressure in vascular plants. Science 148: 339–346.CrossRefGoogle ScholarPubMed
Schopf, J. W. and Barghoorn, E. S.. 1967. Alga-like fossils from the early Precambrian of South Africa. Science 156: 508–512.CrossRefGoogle Scholar
Schwinning, S. and Sala, O. E.. 2004. Hierarchy of responses to resource pulses in arid and semi-arid ecosystems. Oecologia 141: 211–220.CrossRefGoogle ScholarPubMed
Scott, J. M., Csuti, B., Jacobi, J. D., and Estes, J. E.. 1987. Species richness: a geographic approach to protecting future biological diversity. Bioscience 37: 782–788.CrossRefGoogle Scholar
Scott, M. G. and Larson, D. W.. 1985. The effect of winter field conditions on the distribution of two species of Umbilicaria. I. CO2 exchange in reciprocally transplanted thalli. New Phytologist 101: 89–101.CrossRefGoogle Scholar
Sculthorpe, C. D. 1967. The Biology of Aquatic Vascular Plants. Reprinted in 1985. London: Edward Arnold.Google Scholar
Seischab, F. K. and Orwig, D.. 1991. Catastrophic disturbances in the presettlement forests of western New York. Bulletin of the Torrey Botanical Club 118: 117–122.CrossRefGoogle 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: 677–687.CrossRefGoogle Scholar
Shaffer, G. P., J. G. Gosselink, and S. S. Hoeppner. 2005. The Mississippi River alluvial plain. pp. 272–315. In Fraser, L. H. and Keddy, P. A. (eds.) The World's Largest Wetlands: Ecology and Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Shannon, C. E. and Weaver, W.. 1949. The Mathematical Theory of Communication. Urbana: University of Illinois Press.Google Scholar
Sheail, J. and T. C. E. Wells. 1983. The fenlands of Huntingdonshire, England: a case study in catastrophic change. pp. 375–393. In Gore, A. J. P. (ed.) Ecosystems of the World 4B. Mires: Swamp, Bog, Fen and Moor. Amsterdam: Elsevier Scientific Publishing Company.Google Scholar
Shimwell, D. W. 1971. The Description and Classification of Vegetation. Seattle: University of Washington Press.Google Scholar
Shipley, B. 1993. A null model for competitive hierarchies in competition matrices. Ecology 74: 1693–1699.CrossRefGoogle Scholar
Shipley, B. and Dion, J.. 1992. The allometry of seed production in herbaceous angiosperms. The American Naturalist 139: 467–483.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: 340–345.CrossRefGoogle Scholar
Shipley, B. and Peters, R. H.. 1990. A test of the Tilman model of plant strategies: relative growth rate and biomass partioning. The American Naturalist 136: 139–153.CrossRefGoogle Scholar
Shipley, B., Keddy, P. A., and Lefkovitch, L. P.. 1991. Mechanisms producing plant zonation along a water depth gradient: a comparison with the exposure gradient. Canadian Journal of Botany 69: 1420–1424.CrossRefGoogle Scholar
Shrader-Frechette, K. S. and McCoy, E. D.. 1993. Method in Ecology: Strategies for Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Shugart, H. H., D. C. West, and W. R. Emanuel. 1981. Patterns and dynamics of forests: an application of simulation models. pp. 74–106. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession Concepts and Applications. New York: Springer-Verlag.CrossRefGoogle Scholar
Shure, D. J. 1999. Granite outcrops of the southeastern United States. pp. 99–118. In Anderson, R. C., Fralish, J. S., and Baskin, J. M. (eds.) Savannas, Barrens, and Rock Outcrop Plant Communities of North America. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Silliman, B. R. and Bertness, M. D.. 2002. A trophic cascade regulates salt marsh primary production. Proceedings of the National Academy of Sciences USA 99: 10500–10505.CrossRefGoogle ScholarPubMed
Silver, T. 1990. A New Face on the Countryside. Indians, Colonists and Slaves in the South Atlantic Forests, 1500–1800. Cambridge: Cambridge University Press.Google Scholar
Silvertown, J. 1980. The dynamics of a grassland ecosystem: botanical equilibrium in the Park Grass Experiment. Journal of Applied Ecology 17: 491–504.CrossRefGoogle Scholar
Silvertown, J. 1987. Introduction to Plant Population Ecology. 2nd edn. London: Longman.Google Scholar
Silvertown, J. and Charlesworth, D.. 2001. Introduction to Plant Population Biology. 4th edn. Oxford: Blackwell Science.Google Scholar
Silvertown, J., Dodd, M. E., McConway, K., Potts, J., and Crawley, M.. 1994. Rainfall, biomass variation, and community composition in the Park Grass Experiment. Ecology 75: 2430–2437.CrossRefGoogle Scholar
Silvertown, J., Poulton, P., Johnston, E., Edwards, G., Heard, M. and Biss, P. M.. 2006. The Park Grass experiment 1856–2006: its contribution to ecology. Journal of Ecology 94: 801–814.CrossRefGoogle Scholar
Simon, L., Bousquet, J., Lévesque, R. C., and Lalonde, M.. 1993. Origin and diversification of endomycorrhizal fungi and coincidence with vascular land plants. Nature 363: 67–69.CrossRefGoogle Scholar
Simpson, E. H. 1949. Measurement of diversity. Nature 163: 688.CrossRefGoogle Scholar
Sinclair, A. R. E. 1983. The adaptations of African ungulates and their effects on community function. pp. 401–425. In Bouliere, F. (ed.) Tropical Savannas. Amsterdam: 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: 987–994.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: 579–587.CrossRefGoogle Scholar
Sinclair, A. R. E., Krebs, C. J., Fryxell, J. M., Turkington, R., Boutin, S., Boonstra, R., Seccombe-Hett, P., Lundberg, P., and Oksanen, L.. 2000. Testing hypotheses of trophic level interactions: a boreal forest ecosystem. Oikos 89: 313–328.CrossRefGoogle Scholar
Sinclair, A. R. E. and C. J. Krebs. 2001. Trophic interactions, community organization, and the Kluane ecosystem. pp. 25–48. In Krebs, C. J., Boutin, S., and Boonstra, R. (eds.) Ecosystem Dynamics of the Boreal Forest. The Kluane Project. New York: Oxford University Press.Google Scholar
Sioli, H. 1964. General features of the limnology of Amazonia. Verhandlungen/Internationale Vereinigung fur theoretische und angewandte Limnologie 15: 1053–1058.Google Scholar
Sklar, F. H., R. Costanza, and J. W. Day Jr. 1990. Model conceptualization. pp. 625–658. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 1. The Hague: SPB Academic Publishing.Google Scholar
Sklar, F. H.et al. 2005. The ecological–societal underpinnings of Everglades restoration. Frontiers in Ecology and the Environment 3: 161–169.Google Scholar
Slade, A. J. and Hutchings, M. J.. 1987. The effects of nutrient availability on foraging in the clonal herb Glechoma hederacea. Journal of Ecology 75: 95–112.CrossRefGoogle Scholar
Sletvold, N. 2002. Effects of plant size on reproductive output and offspring performance in the facultative biennial Digitalis purpurea. Journal of Ecology 90: 958–996.CrossRefGoogle Scholar
Small, E. 1972a. Water relations of plants in raised Sphagnum peat bogs. Ecology 53: 726–728.CrossRefGoogle Scholar
Small, E. 1972b. Photosynthetic rates in relation to nitrogen recycling as an adaptation to nutrient deficiency in peat bogs. Canadian Journal of Botany 50: 2227–2233.CrossRefGoogle Scholar
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: 487–495.CrossRefGoogle Scholar
Smith, A. 1776. An enquiry into the nature and causes of the wealth of nations. In Adler, M. J. (ed.) 1990. Great Books of the Western World, Vol. 36. Chicago: Encyclopaedia Britannica.Google Scholar
Smith, C. C. 1970. The coevolution of pine squirrels (Tamiasciurus) and conifers. Ecological Monographs 40: 349–371.CrossRefGoogle Scholar
Smith, C. C. and Follmer, D.. 1972. Food preference of squirrels. Ecology 53: 82–91.CrossRefGoogle Scholar
Smith, D. C. 1980. Mechanisms of nutrient movement between lichen symbionts. pp. 197–227. In Cook, C. B., Pappas, P. W., and Rudolph, E. D. (eds.) Cellular Interactions in Symbiosis and Parasitism. Columbus: Ohio State University Press.Google Scholar
Smith, D. C. and Douglas, A. E.. 1987. The Biology of Symbiosis. London: Edward Arnold.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: 673–684.CrossRefGoogle Scholar
Smith, L. M. and Kadlec, J. A.. 1985a. Fire and herbivory in a Great Salt Lake marsh. Ecology 66: 259–265.CrossRefGoogle Scholar
Smith, L. M. and Kadlec, J. A.. 1985b. Comparisons of prescribed burning and cutting of Utah marsh plants. Great Basin Naturalist 45: 463–466.Google Scholar
Smith, R. L. 1986. Elements of Ecology. New York: Harper and Row.Google Scholar
Smith, V. H. 1982. The nitrogen and phosphorus dependence of algal biomass in lakes: an empirical and theoretical analysis. Limnology and Oceanography 27: 1101–1112.CrossRefGoogle Scholar
Smith, V. H. 1983. Low nitrogen to phosphorus ratios favor dominance by blue-green algae in lake phytoplankton. Science 221: 669–671.CrossRefGoogle ScholarPubMed
Smol, J. P. and Cumming, B. F.. 2000. Tracking long-term changes in climate using algal indicators in lake sediments. Journal of Phycology 36: 986–1011.CrossRefGoogle Scholar
Sneath, P. H. A. and Sokal, R. R.. 1973. Numerical Taxonomy. San Francisco: W. H. Freeman.Google Scholar
Snell, T. W. and Burch, D. G.. 1975. The effects of density on resource partitioning in Chamaesyce hirta (Euphorbiaceae). Ecology 56: 742–746.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: 669–684.CrossRefGoogle Scholar
Sobel, D. 1995. Longitude: The True Story of a Lone Genius Who Solved the Greatest Scientific Problem of His Time. New York: Penguin Books.Google Scholar
Soper, J. H. and Maycock, P. F.. 1963. A community of arctic-alpine plants on the east shore of Lake Superior. Canadian Journal of Botany 41: 183–198.CrossRefGoogle Scholar
Sorrie, B. A. 1994. Coastal plain ponds in New England. Biological Conservation 68: 225–233.CrossRefGoogle Scholar
Soulé, M. and Noss, R.. 1998. Rewilding and biodiversity: complementary goals for continental conservation. Wild Earth 8(3): 18–28.Google Scholar
Sousa, W. P. 1984. The role of disturbance in natural communities. Annual Review of Ecology and Systematics 15: 353–391.CrossRefGoogle Scholar
Southwood, T. R. E. 1977. Habitat, the templet for ecological strategies?Journal of Animal Ecology 46: 337–365.CrossRefGoogle Scholar
Southwood, T. R. E. 1985. Interactions of plants and animals: patterns and processes. Oikos 44: 5–11.CrossRefGoogle Scholar
Southwood, T. R. E. 1988. Tactics, strategies, and templets. Oikos 52: 3–18.CrossRefGoogle Scholar
Specht, A. and Specht, R.. 1993. Species richness and canopy productivity of Australian plant communities. Biodiversity and Conservation 2: 152–167.CrossRefGoogle Scholar
Specht, R. L. and Specht, A.. 1989. Species richness of overstorey strata in Australian plant communities – the influence of overstorey growth rates. Australian Journal of Botany 37: 321–336.CrossRefGoogle 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 plant propagules. Wetlands 17: 55–64.CrossRefGoogle Scholar
Sporne, K. R. 1956. The phylogenetic classification of the angiosperms. Biological Reviews 31: 1–29.CrossRefGoogle Scholar
Sporne, K. R. 1970. The Morphology of Pteridophytes: The Structure of Ferns and Allied Plants. 3rd edn. London: Hutchinson and Co.Google Scholar
Sprengel, C. K. 1793. Discovery of the secret of nature in the structure and fertilization of flowers. Vieweq, Berlin. Translation of title and first chapter by P. Haase. pp. 3–43. In Lloyd, D. G. and Barrett, S. C. H. (eds.) Floral Biology. Studies on Floral Evolution in Animal-Pollinated Plants. London: Chapman and Hall.Google Scholar
Starfield, A. M. and Bleloch, A. L.. 1986. Building Models for Conservation and Wildlife Management. New York: Macmillan.Google 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
Stearn, W. T. 1979. Linnaean classification. pp. 96–101. In Black, D. (ed.) 1979. Carl Linnaeus: Travels. New York: Scribner's Sons.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
Steenbergh, W. F. and Lowe, C. H.. 1969. Critical factors during the first years of life of the saguaro (Cereus giganteus) at Saguaro National Monument, Arizona. Ecology 50: 825–834.CrossRefGoogle Scholar
Steila, D. 1993. Soils. pp. 47–54. In Flora of North America, Vol. 1. Introduction. New York: Oxford University Press.Google Scholar
Stein, B. A., Kutner, L. S., and Adams, J. S. (eds.) 2000. Precious Heritage. The Status of Biodiversity in the United States. Oxford: Oxford University Press.Google Scholar
Steneck, R. S. and Dethier, M. N.. 1994. A functional group approach to the structure of algal-dominated communities. Oikos 69: 476–498.CrossRefGoogle Scholar
Stephenson, N. L. 1990. Climatic control of vegetation distribution: the role of the water balance. The American Naturalist 135: 649–680.CrossRefGoogle Scholar
Stephenson, S. N. and Herendeen, P. S.. 1986. Short-term drought effects on the alvar communities of Drummond Island, Michigan. The Michigan Botanist 25: 16–27.Google Scholar
Stevenson, J. C., L. G. Ward, and M. S. Kearney. 1986. Vertical accretion in marshes with varying rates of sea level rise. pp. 241–259. In Wolfe, D. A. (ed.) Estuarine Variability. San Diego: Academic Press.Google Scholar
Stewart, W. N. and Rothwell, G. W.. 1993. Paleobotany and the Evolution of Plants. 2nd edn. Cambridge: Cambridge University Press.Google Scholar
Steyermark, J. A. 1982. Relationships of some Venezuelan forest refuges with lowland tropical floras. pp. 182–220. In Prance, G. T. (ed.) Biological Diversification in the Tropics. New York: Columbia University Press.Google Scholar
Stoltzenberg, D. 2004. Fritz Haber. Chemist, Nobel Laureate, German, Jew. Philadelphia: Chemical Heritage 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 yr B.P. Quaternary Research 12: 83–118.CrossRefGoogle Scholar
Strong, D. R. Jr., Simberloff, D., Abele, L. G., and Thistle, A. B. (eds). 1984. Ecological Communities. Conceptual Issues and the Evidence. Princeton: Princeton University Press.Google Scholar
Stupka, A. 1964. Trees, Shrubs and Woody Vines of Great Smoky National Park. Knoxville: The University of Tennessee Press.Google Scholar
Suffling, R., Lihou, C., and Morand, Y.. 1988. Control of landscape diversity by catastrophic disturbance: a theory and case study of fire in a Canadian boreal forest. Environmental Management 12: 73–78.CrossRefGoogle Scholar
Sun, G., Ji, Q., Dilcher, D. L., Zheng, S., Nixon, K. C., and Wang, X.. 2002. Archaefructaceae, a new basal angiosperm family. Science 296: 899–904.CrossRefGoogle ScholarPubMed
Sutter, R. D. and Kral, R.. 1994. The ecology, status, and conservation of two non-aluvial wetland communities in the south Atlantic and eastern Gulf coastal plain, USA. Biological Conservation 68: 235–243.CrossRefGoogle Scholar
Szarek, S. R. and I. P. Ting. 1975. Photosynthetic efficiency of CAM plants in relation to C3 and C4 plants. pp. 289–297. In Marcelle, R. (ed.) Environmental and Biological Control of Photosynthesis. The Hague: W. Junk.CrossRefGoogle Scholar
Tabachnick, B. G. and Fidell, L. S.. 2001. Using Multivariate Statistics. 4th edn. Boston: Allyn and Bacon.Google Scholar
Taiz, L. and Zeiger, E.. 1991. Plant Physiology. San Francisco: Benjamin-Cummings.Google Scholar
Takhtajan, A. 1969. Flowering Plants: Origin and Dispersal. Edinburgh: Oliver and Boyd. Translated and revised from a Russian second edition published in Moscow in 1961.Google Scholar
Takhtajan, A. 1986. Floristic Regions of the World. Berkeley: University of California Press. Translated by T. J. Crovello.Google Scholar
Tansley, A. G. 1939. The British Islands and their Vegetation. Cambridge: Cambridge University Press.Google Scholar
Tansley, A. 1987. What is ecology?Biological Journal of the Linnean Society 32: 5–16.CrossRefGoogle Scholar
Tansley, A. G. and Adamson, R. S.. 1925. Studies of the vegetation of the English chalk. Part III. The chalk grasslands of the Hampshire-Sussex border. Journal of Ecology XIII: 177–223.CrossRefGoogle Scholar
Tansley, A. G. and Chipp, T. F. (eds.) 1926. Aims and Methods in the Study of Vegetation. London: The British Empire Vegetation Committee and Crown Agents for Colonies.Google 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 template for life history strategies. Oikos 58: 239–250.CrossRefGoogle Scholar
Taylor, T. N., Kerp, H., and Hass, H.. 2005. Life history biology of early land plants: deciphering the gametophyte phase. Proceedings of the National Academy of Sciences 102: 5892–5897.CrossRefGoogle ScholarPubMed
Taylor, J. G. 1984. Louisiana: A History. New York: W. W. Norton & Company.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, T. N. 1988. The origin of land plants: some answers, more questions. Taxon 37: 805–833.CrossRefGoogle Scholar
Taylor, T. N. 1990. Fungal associations in the terrestrial paleoecosystem. Trends in Ecology and Evolution 5: 21–25.CrossRefGoogle Scholar
Temple, S. A. 1977. Plant–animal mutualism: coevolution with dodo leads to near extinction of plant. Science 197: 886–887.CrossRefGoogle ScholarPubMed
Terborgh, J. 1989. Where Have All the Birds Gone?Princeton: Princeton University Press.Google Scholar
Terborgh, J.et al. 2001. Ecological meltdown in predator-free forest fragments. Science 294: 1923–1926.CrossRefGoogle ScholarPubMed
Thirgood, J. V. 1981. Man and the Mediterranean Forest: A History of Resource Depletion. London: Academic Press.Google Scholar
Thompson, D. J. and Shay, J. M.. 1988. First-year response of a Phragmites marsh community to seasonal burning. Canadian Journal of Botany 67: 1448–1455.CrossRefGoogle Scholar
Thorne, R. F. 1963. Some problems and guiding principles of angiosperm phylogeny. The American Naturalist 97: 287–305.CrossRefGoogle Scholar
Tilghman, N. G. 1989. Impacts of white-tailed deer on forest regeneration in northwestern Pennsylvania. Journal of Wildlife Management 53: 524–532.CrossRefGoogle Scholar
Tilman, D. 1982. Resource Competition and Community Structure. Princeton: Princeton University Press.Google ScholarPubMed
Tilman, D. and S. Pacala. 1993. The maintenance of species richness in plant communities. pp. 13–25. In Ricklefs, R. E. and Schluter, D. (eds.) Species Diversity in Ecological Communities, Chicago: University of Chicago Press.Google Scholar
Tilman, D., Wedin, D., and Knops, J.. 1996. Productivity and sustainability influenced by biodiversity in grassland ecosystems. Nature 379: 718–720.CrossRefGoogle Scholar
Tinker, P. B., M. D. Jones, and D. M. Durall. 1992. A functional comparison of ecto- and endomycorrhizas. pp. 303–310. In Read, D. J., Lewis, D. H., Fitter, A. H., and Alexander, I. J. (eds.) Mycorrhizas in Ecosystems. Wallingford: CAB International.Google Scholar
Tomlinson, P. B. 1986. The Botany of Mangroves. Cambridge: Cambridge University Press.Google Scholar
Tschudy, R. H., Pillmore, C. L., Orth, C. J., Gilmore, J. S., and Knight, J. D.. 1984. Disruption of the terrestrial plant ecosystem at the Cretaceous-Tertiary boundary, Western Interior. Science 225: 1030–1032.CrossRefGoogle ScholarPubMed
Tubbs, C. H., R. M. DeGraff, M. Yamasaki, and W. M. Healy. 1987. Guide to wildlife tree management in New England northern hardwoods. United States Department of Agriculture and Forestry Service General Technical Report NE-118.
Turner, F. 1994. Beyond Geography. The Western Spirit Against the Wilderness. Fifth printing, first edition in 1983. New Brunswick: Rutgers University Press.Google Scholar
Turner, R. E. 1977. Intertidal vegetation and commercial yields of penaeid shrimp. Transactions of the American Fisheries Society 106: 411–416.2.0.CO;2>CrossRefGoogle Scholar
Turner, R. M. 1990. Long-term vegetation change at a fully protected Sonoran Desert site. Ecology 71: 464–477.CrossRefGoogle Scholar
Turner, R. M., Alcorn, S. M., Olin, G., and Booth, J. A.. 1966. The influence of shade, soil, and water on saguaro seedling establishment. Botanical Gazette 127: 95–102.CrossRefGoogle Scholar
Twolan-Strutt, L. and Keddy, P.. 1996. Above- and belowground competition intensity in two contrasting wetland plant communities. Ecology 77: 259–270.CrossRefGoogle Scholar
Udvardy, M. D. F. 1975. A Classification of the Biogeographical Provinces of the World. IUCN Occasional Paper No. 18. Morges: International Union for the Conservation of Nature and Natural Resources.Google Scholar
Uhl, C. and Kauffman, J. B.. 1990. Deforestation, fire susceptibility and potential responses to fire in the eastern Amazon. Ecology 71: 437–449.CrossRefGoogle Scholar
Underwood, A. J. 1978. The detection of non-random patterns of distribution of species along a gradient. Oecologia 36: 317–326.CrossRefGoogle ScholarPubMed
Urban, D. L. and H. H. Shugart. 1992. Individual based models of forest succession. pp. 249–292. In Glenn-Lewin, D. C., Peet, R. K., and Veblen, T. T.. (eds.) Plant Succession. London: Chapman and Hall.Google Scholar
U.S. Army Coastal Engineering Research Centre. 1977. Shore Protection Manual, Vol. 1, 3rd edn. Washington, D.C.: US Government Printing Office.
U.S. Army Corps of Engineers. 2004. The Mississippi River and Tributaries Project. New Orleans District Office Website (www.mvn.usace.army.mil/pao/bro/misstrib.htm) accessed 26 Mar. 2006.
U.S.D.A. 1975. Soil Taxonomy: A Basic System of Soil Classification for Making and Interpreting Soil Surveys. Agricultural Handbook 436, Washington, D.C.: U.S.D.A.
U.S.D.A. 2004. Emerald ash borer; the green menace. Animal and Plant Health Inspection Service. Program Aid No. 1769.
Vallentyne, J. R. 1974. The Algal Bowl. Lakes and Man. Miscellaneous Special Publication 22. Ottawa: Department of the Environment, Fisheries and Marine Service.Google Scholar
Breemen, N. 1995. How Sphagnum bogs down other plants. Trends in Ecology and Evolution 10: 270–275.CrossRefGoogle ScholarPubMed
Vandermeer, J. B., B. A. Hazlett, and B. Rathcke. 1985. Indirect facilitation and mutualism. pp. 326–343. In Boucher, D. H. (ed.) The Ecology of Mutualism. New York: Oxford University Press.Google Scholar
Valk, A. G. 1981. Succession in wetlands: a Gleasonian approach. Ecology 62: 688–696.CrossRefGoogle Scholar
Valk, A. G. and Davis, C. B.. 1976. The seed banks of prairie glacial marshes. Canadian Journal of Botany 54: 1832–1838.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: 322–335.CrossRefGoogle Scholar
Werf, A., Welschen, A., Welschen, R., and Lambers, H.. 1988. Respiratory energy costs for the maintenance of biomass, for growth and for iron uptake in roots of Carex diandra and Carex acutiformis. Physiologia Plantarum 72: 483–491.CrossRefGoogle Scholar
Venable, D. L. and C. E. Pake. 1999. Population ecology of Sonoran Desert annual plants. pp. 115–142. In Robichaux, R. H. (ed.) The Ecology of Sonoran Desert Plants and Plant Communities. Tucson: University of Arizona Press.Google Scholar
Veneklaas, E. J., Fajardo, A., Obregon, S., and Lozano, J.. 2005. Gallery forest types and their environmental correlates in a Colombian savanna landscape. Ecography 28: 236–252.CrossRefGoogle Scholar
Verhoeven, J. T. A. and Liefveld, W. M.. 1997. The ecological significance of organochemical compounds in Sphagnum. Acta Botanica Neerlandica 46: 117–130.CrossRefGoogle Scholar
Verhoeven, J. T. A. and Schmitz, M. B.. 1991. Control of plant growth by nitrogen and phosphorus in mesotrophic fens. Biogeochemistry 12: 135–148.CrossRefGoogle Scholar
Verhoeven, J. T. A., R. H. Kemmers, and W. Koerselman. 1993. Nutrient enrichment of freshwater wetlands. pp. 33–59. In Vos, C. C. and Opdam, P.. Landscape Ecology of a Stressed Environment. 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: 494–497.CrossRefGoogle ScholarPubMed
Vernadsky, V. 1929. La Biosphère. Paris: Felix Alcan.Google Scholar
Vernadsky, V. I. 1998. The Biosphere. New York: Copernicus, Springer-Verlag. Translated from the French and Russian, including a new foreword, introduction and appendices.CrossRefGoogle Scholar
Vesey-FitzGerald, D. F. 1960. Grazing succession among East African game animals. Journal of Mammalogy 41: 161–172.CrossRefGoogle 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: 651–667.CrossRefGoogle Scholar
Vitousek, P. M. 1982. Nutrient cycling and nitrogen use efficiency. The American Naturalist 119: 553–572.CrossRefGoogle Scholar
Vitousek, P., Ehrlich, P. R., Ehrlich, A. H., and Matson, P.. 1986. Human appropriation of the products of photosynthesis. Bioscience 36: 368–373.CrossRefGoogle Scholar
Vitousek, P. M.et al. 1997. Human alteration of the global nitrogen cycle: causes and consequences. Ecological Applications 7: 737–750.Google Scholar
Vitt, D. H. and Chee, W.. 1990. The relationship of vegetation to surface water chemistry and peat chemistry in fens of Alberta, Canada. Vegetatio 89: 87–106.CrossRefGoogle Scholar
Vivian-Smith, G. 1997. Microtopographic heterogeneity and floristic diversity in experimental wetland communities. Journal of Ecology 85: 71–82.CrossRefGoogle Scholar
Vogel, S. 1996. Christian Konrad Sprengel's theory of the flower: the cradle of floral ecology. pp. 44–62. In Lloyd, D. G. and Barrett, S. C. H. (eds.) Floral Biology. Studies on Floral Evolution in Animal-Pollinated Plants. London: Chapman and Hall.Google Scholar
Vogl, R. 1969. One hundred and thirty years of plant succession in a southeastern Wisconsin lowland. Ecology 50: 248–255.CrossRefGoogle Scholar
Humboldt, A. 1845. Cosmos: A Sketch of the Physical Description of the Universe, Vol. 1. Translated by E. C. Otté. Foundations of Natural History. Baltimore: Johns Hopkins University Press. 1997. (Originally produced in five volumes: 1845, 1847, 1850–51, 1858, and 1862.)Google Scholar
Waggoner, P. E. and Stephens, G. R.. 1970. Transition probabilities for a forest. Nature 225: 1160–1161.CrossRefGoogle ScholarPubMed
Walker, D. 1970. Direction and rate in some British post-glacial hydroseres. pp. 117–139. In Walker, D. and West, R. G. (eds.) Studies in the Vegetational History of the British Isles. Cambridge: Cambridge University Press.Google Scholar
Walker, J., C. H. Thompson, I. F. Fergus, and B. R. Tunstall. 1981. Plant succession and soil development in coastal sand dunes of subtropical eastern Australia. pp. 107–131. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession. Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
Wallin, I. E. 1927. Symbioticism and the Origin of Species. Baltimore: Williams and Wilkins.Google Scholar
Wardle, D. A. 1995. Impact of disturbance on detritus food-webs in agro-ecosystems of contrasting tillage and weed management practices. Advances in Ecological Research 26: 105–185.CrossRefGoogle Scholar
Wardle, D. A. 2002. Communities and Ecosystems: Linking the Aboveground and Belowground Components. Princeton: Princeton University Press.Google Scholar
Wardle, D. A., Huston, M. A., Grime, J. P., Berendse, F., Garnier, E., Laurenroth, W. K., Setala, H., and Wilson, S. D.. 2000. Biodiversity and ecosystem function: an issue in ecology. Bulletin of the Ecological Society of America 81: 235–239.Google Scholar
Wardle, D. A., Bardgett, R. D., Klironomos, J. N., Setälä, H., Putten, W. H., and Wall, D. H.. 2004. Ecological linkages between aboveground and belowground biota. Science 304: 1629–1633.CrossRefGoogle ScholarPubMed
Watkinson, A. R. 1985a. Plant responses to crowding. pp. 275–289. In White, J. (ed.) Studies in Plant Demography: A Festschrift for John L. Harper. London: Academic Press.Google Scholar
Watkinson, A. R. 1985b. On the abundance of plants along an environmental gradient. Journal of Ecology 73: 569–578.CrossRefGoogle Scholar
Watkinson, A. R. and Freckleton, R. P.. 1997. Quantifying the impact of arbuscular mycorrhizae on plant competition. Journal of Ecology 85: 541–545.CrossRefGoogle Scholar
Watkinson, A. R. and C. C. Gibson. 1988. Plant parasitism: the population dynamics of parasitic plants and their effects upon plant community structure. pp. 393–411. In Davy, A. J., Hutchings, M. J., and Watkinson, A. R.. Plant Population Biology. Oxford: Blackwell Scientific Publications.Google Scholar
Watt, A. S. 1919. On the causes of failure of natural regeneration in British oakwoods. Journal of Ecology 7: 173–203.CrossRefGoogle Scholar
Watt, A. S. 1923. On the ecology of British beechwoods with special reference to their regeneration. Part I. The causes of failure of natural regeneration of the beech. Journal of Ecology 11: 1–48.CrossRefGoogle Scholar
Weaver, J. E. and Clements, F. E.. 1929. Plant Ecology. New York: McGraw-Hill.Google Scholar
Weaver, J. E. and Clements, F. E.. 1938. Plant Ecology. 2nd edn. New York: McGraw-Hill Book Company.Google Scholar
Weber, W. and Rabinowitz, A.. 1996. A global perspective on large carnivore conservation. Conservation Biology 10: 1046–1054.CrossRefGoogle Scholar
Weetman, G. F. 1983. Forestry practices and stress on Canadian forest land. pp. 260–301. In Simpson-Lewis, W., McKechnie, R., and Neimanis, V. (eds.) Stress on Land in Canada. Ottawa: Lands Directorate, Environment Canada.Google Scholar
Weiher, E. and Keddy, P. A.. 1995a. The assembly of experimental wetland plant communities. Oikos 73: 323–335.CrossRefGoogle Scholar
Weiher, E. and Keddy, P. A.. 1995b. Assembly rules, null models, and trait dispersion: new questions from old patterns. Oikos 74: 159–165.CrossRefGoogle Scholar
Weiher, E. and Keddy, P. A. (eds.). 1999. Ecological Assembly Rules: Perspectives, Advances, Retreats. Cambridge: Cambridge University Press.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: 609–620.CrossRefGoogle Scholar
Wein, R. W. 1983. Fire behaviour and ecological effects in organic terrain. pp. 81–95. In Wein, R. W. and MacLean, D. A. (eds.) The Role of Fire in Northern Circumpolar Ecosystems. New York: John Wiley and Sons Ltd.Google Scholar
Wein, R. W. and Moore, J. M.. 1977. Fire history and rotations in the New Brunswick Acadian forest. Canadian Journal of Forest Research 7: 285–294.CrossRefGoogle Scholar
Weiner, J. 1985. Size hierarchies in experimental populations of annual plants. Ecology 66: 743–752.CrossRefGoogle Scholar
Weiner, J. 1986. How competition for light and nutrients affects size variablility in Ipomea tricolor populations. Ecology 67: 1425–1427.CrossRefGoogle Scholar
Weiner, J. and Thomas, S. C.. 1986. Size variability and competition in plant monocultures. Oikos 47: 221–222.CrossRefGoogle Scholar
Weisner, S. E. B. 1990. Emergent Vegetation in Eutrophic Lakes: Distributional Patterns and Ecophysiological Constraints. Sweden: Grahns Boktryckeri.
Welcomme, R. L. 1976. Some general and theoretical considerations on the fish yield of African rivers. Journal of Fish Biology 8: 351–364.CrossRefGoogle Scholar
Welcomme, R. L. 1979. Fisheries Ecology of Floodplain Rivers. London: Longman.Google Scholar
Weldon, C. W. and Slauson, W. L.. 1986. The intensity of competition versus its importance: an overlooked distinction and some implications. The Quarterly Review of Biology 61: 23–44.CrossRefGoogle Scholar
Weller, D. E. 1990. Will the real self-thinning rule please stand up? A reply to Osawa and Sugita. Ecology 71: 1204–1207.CrossRefGoogle Scholar
Weller, M. W. 1978. Management of freshwater marshes for wildlife. pp. 267–284. In Good, R. E., Whigham, D. F., and Simpson, R. L. (eds.) Freshwater Wetlands: Ecological Processes and Management Potential. New York: Academic Press.Google Scholar
Weller, M. W. 1994. Freshwater Marshes: Ecology and Wildlife Management. 3rd edn. Minneapolis: University of Minnesota.Google Scholar
Wells, H. G. 1956. The Outline of History: Being a Plain History of Life and Mankind. Garden City, NY: Garden City Books. Revised and brought up to the end of the Second World War by Raymond Postgate.Google Scholar
West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) 1981. Forest Succession: Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
Westhoff, V. and E. van der Maarel. 1973. The Braun–Blanquet approach. pp. 617–707. In Whittaker, R. H. (ed.) Ordination and Classification of Communities. The Hague: Junk.CrossRefGoogle Scholar
Westoby, M. 1984. The self-thinning rule. Advances in Ecological Research 14: 167–225.CrossRefGoogle Scholar
Westoby, M. 1998. Leaf–height–seed (LHS) plant ecology strategy scheme. Plant and Soil 199: 213–227.CrossRefGoogle Scholar
Westoby, M., M. Leishman, and J. Lord. 1997. Comparative ecology of seed size and dispersal. pp. 143–162. In Silvertown, J., Franco, M., and Harper, J. L. (eds.) Plant Life Histories: Ecology, Phylogeny and Evolution. Cambridge: Cambridge University Press.Google 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
Whelan, P. M. and Hamann, O.. 1989. Vegetation regrowth on Isla Pinta: a success story. Noticias de Galápagos 48: 11–13.Google Scholar
Whelan, R. J. 1995. The Ecology of Fire. Cambridge: Cambridge University Press.Google Scholar
Whisenant, S. G. 1999. Repairing Damaged Wildlands. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
White, I. D., Mottershead, D. N., and Harrison, S. J.. 1992. Environmental Systems: An Introductory Text. 2nd edn. London: Chapman and Hall.Google Scholar
White, P. S. 1979. Pattern, process and natural disturbance in vegetation. The Botanical Review 45: 229–299.CrossRefGoogle Scholar
White, P. S. 1994. Synthesis: vegetation pattern and process in the Everglades ecosystem. pp. 445–460. In Davis, S. and Ogden, J. (eds.) Everglades: The Ecosystem and its Restoration. Delray Beach: St. Lucie Press.Google Scholar
White, P. S., S. P. Wilds, and G. A. Thunhorst. 1998. Southeast. pp. 255–314. In Mac, M. J., Opler, P. A., Haecker, C. E. Puckett, and Doran, P. D.. (eds.) Status and Trends of the Nation's Biological Resources, 2 Vols. Reston: U.S. Department of the Interior, U.S. Geological Survey.Google Scholar
White, T. C. R. 1993. The Inadequate Environment: Nitrogen and the Abundance of Animals. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Whittaker, R. H. 1952. A study of summer foliage insect communities in the Great Smoky Mountains. Ecological Monographs 22: 1–44.CrossRefGoogle Scholar
Whittaker, R. H. 1954a. The ecology of serpentine soils. I. Introduction. Ecology 35: 258–259.CrossRefGoogle Scholar
Whittaker, R. H. 1954b. The ecology of serpentine soils. IV. The vegetational response to serpentine soils. Ecology 35: 275–288.Google Scholar
Whittaker, R. H. 1956. Vegetation of the Great Smoky Mountains. Ecological Monographs 26: 1–79.CrossRefGoogle Scholar
Whittaker, R. H. 1960. Vegetation of the Siskiyou Mountains, Oregon and California. Ecological Monographs 30: 279–338.CrossRefGoogle Scholar
Whittaker, R. H. 1962. Classification of natural communities. The Botanical Review 28: 1–239.CrossRefGoogle Scholar
Whittaker, R. H. 1965. Dominance and diversity in land plant communities. Science 147: 250–260.CrossRefGoogle ScholarPubMed
Whittaker, R. H. 1967. Gradient analysis of vegetation. Biological Reviews 42: 207–264.CrossRefGoogle ScholarPubMed
Whittaker, R. H. 1972. Evolution and measurement of species diversity. Taxon 21: 213–251.CrossRefGoogle Scholar
Whittaker, R. H. 1973a. Direct gradient analysis: techniques. pp. 9–31. In Whittaker, R. H. (ed.) Ordination and Classification of Communities. Part V. The Hague: W. Junk.CrossRefGoogle Scholar
Whittaker, R. H. (ed.) 1973b. Ordination and Classification of Communities. Part V. The Hague: W. Junk.CrossRefGoogle Scholar
Whittaker, R. H. 1975. Communities and Ecosystems. 2nd edn. London: Macmillan.Google Scholar
Wiens, J. A. 1977. On competition and variable environments. American Scientist 65: 590–597.Google Scholar
Wilcove, D. S., C. H. McLellan, and A. P. Dobson. 1986. Habitat fragmentation in the temperate zone. pp. 237–256. In Soulé, M. E. (ed.) Conservation B; the Science of Scarcity and Diversity. Sunderland: Sinauer Associates.Google Scholar
Wilde, S. A. 1958. Forest Soils: Their Properties and Relation to Silviculture. New York: The Ronald Press Company.Google Scholar
Wilf, P., Cúneo, N. R., Johnson, K. R., Hicks, J. F., Wing, S. L. and Obradovich, J. D.. 2003. High plant diversity in Eocene South America: evidence from Patagonia. Science 300: 122–125.CrossRefGoogle ScholarPubMed
Williams, C. B. 1964. Patterns in the Balance of Nature. London: Academic Press.Google Scholar
Williams, E. J. 1962. The analysis of competition experiments. Australian Journal of Biological Science 15: 509–525.CrossRefGoogle Scholar
Williams, G. C. 1975. Sex and Evolution. Monographs in Population Biology. No. 8. Princeton: Princeton University Press.Google ScholarPubMed
Williams, M. 1989. The lumberman's assault on the southern forest, 1880–1920. pp. 238–288. In Williams, M.. Americans and Their Forests: A Historical Geography. Cambridge: Cambridge University Press.Google Scholar
Williamson, G. B. 1990. Allelopathy, Koch's Postulates and the neck riddle. pp. 143–162. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Willig, M. R., Kaufman, D. M., and Stevens, R. D.. 2003. Latitudinal gradients of biodiversity: pattern, process, scale and synthesis. Annual Review of Ecology, Evolution and Systematics 34: 273–309.CrossRefGoogle 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: 353–374.CrossRefGoogle Scholar
Willson, M. F. 1984. Mating patterns in plants. pp. 261–276. In Dirzo, R. and Sarukháh, J. (eds.) Perspectives on Plant Population Ecology. Sunderland: Sinauer Associates.Google 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: Sinauer Associates.Google Scholar
Wilson, J. B. 1988. Shoot competition and root competition. Journal of Applied Ecology 25: 279–296.CrossRefGoogle 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: 527–538.CrossRefGoogle Scholar
Wilson, S. D. 1993. Competition and resource availability in heath and grassland in the Snowy Mountains of Australia. Journal of Ecology 81: 445–451.CrossRefGoogle Scholar
Wilson, S. D. 1999. Plant interactions during secondary succession. pp. 629–650. In Walker, L. R. (ed.) Ecosystems of Disturbed Ground. Amsterdam: Elsevier.Google Scholar
Wilson, S. D. and Keddy, P. A.. 1986a. Measuring diffuse competition along an environmental gradient: results from a shoreline plant community. The American Naturalist 127: 862–869.CrossRefGoogle Scholar
Wilson, S. D. and Keddy, P. A.. 1986b. Species competitive ability and position along a natural stress/disturbance gradient. Ecology 67: 1236–1242.CrossRefGoogle Scholar
Wimsatt, W. C. 1982. Reductionistic research strategies and their biases in the units of selection controversy. pp. 155–201. In Saarinen, E. (ed.) Conceptual Issues in Ecology. Dordrecht: D. Reidel.CrossRefGoogle Scholar
Wing, S. L. 1997. Global warming and plant species richness: a case study of the Paleocene/Eocene boundary. pp. 163–185. In Reaka-Kudla, M. L., Wilson, D. E., and Wilson, E. O. (eds.) Biodiversity II: Understanding and Protecting Our Biological Resources. Washington, D.C.: Joseph Henry Press.Google Scholar
Wing, S. L. and B. H. Tiffney. 1987. Interactions of angiosperms and herbivorous tetrapods through time. pp. 203–224. In Friis, E., Chaloner, W. G., and Crane, P. R. (eds.) The Origins of Angiosperms and Their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Wiser, S. K., Peet, R. K., and White, P. S.. 1996. High-elevation rock outcrop vegetation of the southern Appalachian Mountains. Journal of Vegetation Science 7: 703–722.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: 1609–1617.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: 229–238.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: 181–188.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: 147–156.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: 253–258.CrossRefGoogle Scholar
Wisheu, I. C., P. A. Keddy, D. R. J. Moore, S. J. McCanny, and C. L. Gaudet. 1991. Effects of eutrophication on wetland vegetation. pp. 112–121. In Kusler, J. and Smardon, R. (eds.) Wetlands of the Great Lakes: Protection and Restoration Policies; Status of the Science. New York: Managers Inc.Google Scholar
Witmer, M. C. and Cheke, A. S.. 1991. The dodo and the tambalacoque tree: an obligate mutualism reconsidered. Oikos 61: 133–137.CrossRefGoogle Scholar
Wium-Anderson, S. 1971. Photosynthetic uptake of free CO2 by the roots of Lobelia dortmanna. Plantarum 25: 245–248.CrossRefGoogle Scholar
Wolbach, W. S., Lewis, R. S., and Anders, E.. 1985. Cretaceous extinctions: evidence for wildfires and search for meteoritic material. Science 230: 167–170.Google Scholar
Wolfe, J. A. 1991. Palaeobotanical evidence for a June “impact winter” at the Cretaceous/Tertiary boundary. Nature 352: 420–423.CrossRefGoogle Scholar
Wolff, W. J. 1993. Netherlands-wetlands. Hydrobiologia 265: 1–14.CrossRefGoogle Scholar
Wolin, C. L. 1985. The population dynamics of mutualistic systems. pp. 248–269. In Boucher, D. H. (ed.) The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Woodbury, A. M. 1947. Distribution of pigmy conifers in Utah and northeastern Arizona. Ecology 28: 113–126.CrossRefGoogle Scholar
Woodley, S., Kay, J., and Francis, G.. (eds.) 1993. Ecological Integrity and the Management of Ecosystems. Delray Beach: St. Lucie Press.Google Scholar
Woods, K. D. and R. H. Whittaker. 1981. Canopy–understory interaction and the internal dynamics of mature hardwood and hemlock-hardwood forests. pp. 305–323. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Applications. New York: Springer-Verlag.CrossRefGoogle Scholar
Woodward, F. I. 1987. Climate and Plant Distribution. Cambridge: Cambridge University Press.Google Scholar
Woodward, F. I. 1992. Predicting plant responses to global environmental change. New Phytologist 122: 239–251.CrossRefGoogle Scholar
Woodward, F. I. and C. K. Kelly. 1997. Plant functional types: towards a definition by environmental constraints. pp. 47–65. In Smith, T. M., Shugart, H. H., and Woodward, F. I. (eds.) Plant Functional Types. Cambridge: Cambridge University Press.Google Scholar
Woodwell, G. M. 1962. Effects of ionizing radiation on terrestrial ecosystems. Science 138: 572–577.CrossRefGoogle ScholarPubMed
Woodwell, G. M. 1963. The ecological effects of radiation. Scientific American 208: 42–47.CrossRefGoogle Scholar
World Commission on Environment and Development. 1987. Our Common Future. Oxford: Oxford University Press.
Wright, D. H. and Reeves, J. H.. 1992. On the meaning and measurement of nestedness of species assemblage. Oecologia 92: 416–428.CrossRefGoogle Scholar
Wright, H. A. and Bailey, A. W.. 1982. Fire Ecology. New York: Wiley.Google Scholar
Wright, J. P., Jones, C. G., and Flecker, A. S.. 2002. An ecosystem engineer, the beaver, increases species richness at the landscape scale. Oecologia 132: 96–101.CrossRefGoogle ScholarPubMed
Wright, R. 2004. A Short History of Progress. Toronto: Anansi Press.Google Scholar
Yoda, K., Kira, T., Ogawa, H., and Hozumi, K.. 1963. Self-thinning in overcrowded pure stands under cultivated and natural conditions. Journal of Biology/Osaka City University 14: 107–129.Google Scholar
Yodzis, P. 1986. Competition, mortality, and community structure. pp. 480–492. In Diamond, J. and Case, T. J. (eds.) Community Ecology. New York: Harper and Row.Google Scholar
Yodzis, P. 1989. Introduction to Theoretical Ecology. New York: Harper and Row.Google Scholar
Young, E. 2006. Easter Island: a monumental collapse?New Scientist 2562: 30–34.Google Scholar
Young, K., Ulloa, C. U., Luteyn, J. L., and Knapp, S.. 2002. Plant evolution and endemism in Andean South America: an introduction. The Botanical Review 68: 4–21.CrossRefGoogle Scholar
Young, T. P. and Augspruger, C. K.. 1991. Ecology and evolution of long-lived semelparous plants. Trends in Ecology and Evolution 6: 285–289.CrossRefGoogle ScholarPubMed
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: 1602–1615.CrossRefGoogle Scholar
Zachos, J., Pagani, M., Sloan, L., Thomas, E., and Billups, K.. 2001. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292: 686–693.CrossRefGoogle ScholarPubMed
Zedler, J. B. and P. A. Beare. 1986. Temporal variability of salt marsh vegetation: the role of low-salinity gaps and environmental stress. pp. 295–306. In Wolfe, D. A. (ed.) Estuarine Variability. San Diego: Academic Press.Google Scholar
Zedler, J. B. and C. P. Onuf. 1984. Biological and physical filtering in arid-region estuaries: seasonality, extreme events, and effects of watershed modification. pp. 415–432. In Kennedy, V. S. (ed.) The Estuary as a Filter. New York: Academic Press.Google Scholar
Zobel, M. 1997. The relative role of species pools in determining plant species richness: an alternative explanation of species coexistence?Trends in Ecology and Evolution 12: 266–269.CrossRefGoogle Scholar
Aaviksoo, K., M. Ilomets, and M. Zobel. 1993. Dynamics of mire communities: a Markovian approach (Estonia). pp. 23–43. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 2. The Hague: SPB Academic Publishing.Google Scholar
Abraham, K. F. and C. J. Keddy. 2005. The Hudson Bay lowland. pp. 118–148. In Fraser, L. H. and Keddy, P. A. (eds.) The World's Largest Wetlands: Ecology and Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Abrahamson, W. G. and Gadgil, M.. 1973. Growth form and reproductive effort in goldenrods (Solidago, Compositae). The American Naturalist 107: 651–661.CrossRefGoogle Scholar
Ackerman, J. D. and Montalvo, A. M.. 1990. Short- and long-term limitations to fruit production in a tropical orchid. Ecology 71: 263–272.CrossRefGoogle Scholar
Adam, P. 1990. Saltmarsh Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Adamus, P. R. 1992. Choices in monitoring wetlands. pp. 571–592. In McKenzie, D. H., Hyatt, D. E., and McDonald, V. J. (eds.) Ecological Indicators. London: Elsevier Applied Science.Google Scholar
Aerts, R. 1996. Nutrient resorption from senescing leaves of perennials: are there general patterns. Journal of Ecology 84: 597–608.CrossRefGoogle Scholar
Agnew, A. D. Q. 1961. The ecology of Juncus effusus L. in North Wales. Journal of Ecology 49: 83–102.CrossRefGoogle Scholar
Allee, W. C. 1951. Cooperation Among Animals with Human Implications. New York: Schuman. (Revised edition of Social Life of Animals. 1938. New York: Norton.)Google Scholar
Allee, W. C., Emerson, A. E., Park, O., Park, T., and Schmidt, K. P.. 1949. Principles of Animal Ecology. Philadelphia: Saunders.Google Scholar
Allen, E. B. 1988. The Reconstruction of Disturbed Arid Ecosystems. Boulder: Westview Press.Google Scholar
Allen, E. B. and M. F. Allen. 1990. The mediation of competition by mycorrhizae in successional and patchy environments. pp. 367–389. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.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
Alvarez, W. 1998. T. rex and the Crater of Doom. New York: Vintage Books.Google Scholar
Alverson, W. S., Waller, D. M., and Solheim, S. J.. 1988. Forests to deer: edge effects in northern Wisconsin. Conservation Biology 2: 348–358.CrossRefGoogle Scholar
Anderson, R. C., Liberta, A. E., Dickman, L. A., and Katz, A. J.. 1983. Spatial variation in vesicular-arbuscular mycorrhizal spore density. Bulletin of the Torrey Botanical Club 110: 519–525.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: 111–117.CrossRefGoogle ScholarPubMed
Anderson, R. C., Fralish, J. S., and Baskin, J. M.. 1999. Savannas, Barrens, and Rock Outcrop Plant Communities of North America. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Antonovics, J. 1984. Genetic variation within populations. pp. 229–241. In Dirzo, R. and Sarukhán, J. (eds.) Perspectives on Plant Population Ecology. Sunderland: Sinauer.Google Scholar
Archer, S. 1989. Have southern Texas savannas been converted to woodlands in recent history?The American Naturalist 134: 545–561.CrossRefGoogle Scholar
Archibold, O. W. 1995. Ecology of World Vegetation. London: Chapman and Hall.CrossRefGoogle Scholar
Arthur, W. 1982. The evolutionary consequences of interspecific competition. Advances in Ecological Research 12: 127–187.CrossRefGoogle Scholar
Arthur, W. 1987. The Niche in Competition and Evolution. Chichester: Wiley.Google Scholar
Ashton, P. S. 1988. Dipterocarp biology as a window to the understanding of tropical forest structure. Annual Review of Ecology and Systematics 19: 347–370.CrossRefGoogle Scholar
Ashton, P. S., Givnish, T. J., and Appanah, S.. 1988. Staggered flowering in the Dipterocarpaceae: new insights into floral induction and the evolution of mast fruiting in the seasonal tropics. The American Naturalist 132: 44–66.CrossRefGoogle Scholar
Atwood, E. L. 1950. Life history studies of the nutria, or coypu, in coastal Louisiana. Journal of Wildlife Management 14: 249–265.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: 941–952.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. 1968. An ordination study of a chalk grassland community. Journal of Ecology 56: 739–757.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: 559–570.CrossRefGoogle Scholar
Austin, M. P. and Austin, B. O.. 1980. Behaviour of experimental plant communities along a nutrient gradient. Journal of Ecology 68: 891–918.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. Australian Journal of Ecology 21: 154–164.CrossRefGoogle Scholar
Austin, M. P., J. G. Pausas, and I. R. Noble. 1997. Modelling environmental and temporal niches of eucalypts. pp. 129–150. In Williams, J. E. and Woinarski, J. C. Z. (eds.) Eucalypt Ecology: Individuals to Ecosystems. Cambridge: Cambridge University Press.Google Scholar
Axelrod, D. I. 1970. Mesozoic paleogeography and early angiosperm history. The Botanical Review 36: 277–319.CrossRefGoogle Scholar
Axelrod, D. I. and P. H. Raven. 1972. Evolutionary biogeography viewed from plate tectonic theory. pp. 218–236. In Behnke, J. A. (ed.) Challenging Biological Problems: Directions Toward Their Solution. Oxford: Oxford University Press.Google Scholar
Bacon, P. R. 1978. Flora and Fauna of the Caribbean. Trinidad: Key Caribbean Publications.Google Scholar
Bailes, K. E. 1990. Science and Russian Culture in an Age of Revolutions. V. I. Vernadsky and his Scientific School, 1863–1945. Bloomington, IN: Indiana University Press.Google Scholar
Baker, H. 1937. Alluvial meadows: a comparative study of grazed and mown meadows. Journal of Ecology 25: 408–420.CrossRefGoogle Scholar
Baker-Brosh, K. and Peet, R. K.. 1997. The ecological significance of lobed and toothed leaves in temperate forest trees. Ecology 78: 1250–1255.Google Scholar
Bakker, R. T. 1978. Dinosaur feeding behaviour and the origin of flowering plants. Nature 274: 661–663.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: 113–120.CrossRefGoogle Scholar
Baldwin, W. K. W. 1958. Plants of the Clay Belt of Northern Ontario and Quebec. National Museum of Canada, Bulletin No. 156.
Ball, P. J. and T. D. Nudds. 1989. Mallard habitat selection: an experiment and implications for management. pp. 659–671. In R. R. Sharitz and J. W. Gibbons (eds.) Freshwater Wetlands and Wildlife. U.S. Department of Energy. Proceedings of a symposium held at Charleston, South Carolina, March 24–27, 1986.
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: 109–117.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
Barnett, V. 1994. Statistics and the long-term experiments: past achievements and future challenges. pp. 165–183. In Leigh, R. A. and Johnston, A. E. (eds.) Long-term Experiments in Agricultural and Ecological Sciences. Proceedings of a conference to celebrate the 150th anniversary of Rothamsted Experimental Station, held at Rothamsted, July 14–17, 1993. Wallingford: CAB International.Google Scholar
Barrett, S. C. H. 2002. The evolution of plant sexual diversity. Nature Reviews Genetics 3: 274–284.CrossRefGoogle ScholarPubMed
Barry, J. M. 1997. Rising Tide. The Great Mississippi Flood of 1927 and How it Changed America. New York: Simon and Schuster.Google Scholar
Barth, F. G. 1985. Insects and Flowers: The Biology of a Partnership. Princeton: Princeton University Press. Translated from 1982 German edition by M. A. Biederman-Thorson.Google Scholar
Barthlott, W., Porembski, S., Fischer, E., and Gemmel, B.. 1998. First protozoa-trapping plant found. Nature 392: 447.CrossRefGoogle Scholar
Baskin, J. M. and Baskin, C. C.. 1985. A floristic study of a cedar glade in Blue Licks Battlefield State Park, Kentucky. Castanea 50: 19–25.Google Scholar
Bauer, C. R., Kellogg, C. H., Bridgham, S. D., and Lamberti, G. A.. 2003. Mycorrhizal colonization across hydrological gradients in restored and reference freshwater wetlands. Wetlands 23: 961–968.CrossRefGoogle Scholar
Baylis, G. T. S. 1980. Mycorrhizas and the spread of beech. New Zealand Journal of Ecology 3: 151–153.Google Scholar
Beard, J. S. 1944. Climax vegetation in tropical America. Ecology 25: 127–158.CrossRefGoogle Scholar
Beard, J. S. 1949. The Natural Vegetation of the Windward and Leeward Islands. Oxford: Clarendon Press.Google Scholar
Beard, J. S. 1973. The physiognomic approach. pp. 355–386. In Whittaker, R. H. (ed.) Ordination and Classification of Communities. Part V. The Hague: W. Junk.CrossRefGoogle Scholar
Beattie, A. J. and Culver, D. C.. 1981. The guild of myrmecochores in the herbaceous flora of West Virginia forests. Ecology 62: 107–115.CrossRefGoogle Scholar
Beeftink, W. G. 1977. The coastal salt marshes of western and northern Europe: an ecological and phytosociological approach. pp. 109–155. In Chapman, V. J. (ed.) Ecosystems of the World 1: Wet Coastal Ecosystems. Amsterdam: Elsevier Scientific Publishing Company.Google Scholar
Bégin, 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. Geographie Physique et Quaternaire 43: 355–366.CrossRefGoogle Scholar
Begon, M. and Mortimer, M.. 1981. Population Ecology: A Unified Study of Animals and Plants. Oxford: Blackwell.Google Scholar
Belcher, J., Keddy, P. A., and Catling, P. M. C.. 1992. Alvar vegetation in Canada: a multivariate description at two scales. Canadian Journal of Botany 70: 1279–1291.CrossRefGoogle Scholar
Belcher, J. W., Keddy, P. A., and Twolan-Strutt, L.. 1995. Root and shoot competition intensity along a soil depth gradient. Journal of Ecology 83: 673–682.CrossRefGoogle Scholar
Bell, A. D. 1984. Dynamic morphology: a contribution to plant population ecology. pp. 48–65. In Dirzo, R. and Sarukhán, J. (eds.) Perspectives on Plant Population Ecology. Sunderland: Sinauer.Google Scholar
Bell, R. A. 1993. Cryptoendolithic algae of hot semiarid lands and deserts. Journal of Phycology 29: 133–139.CrossRefGoogle Scholar
Belsky, A. J. 1992. Effects of grazing, competition, disturbance and fire on species composition and diversity in grassland communities. Journal of Vegetation Science 3: 187–200.CrossRefGoogle Scholar
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
Benecke, P. and R. Mayer. 1971. Aspects of soil water behavior as related to beech and spruce stands: some results of water balance investigations. pp. 153–168. In Ellenburg, H. (ed.) Integrated Experimental Ecology: Methods and Results of Ecosystem Research in the German Solling Project, Vol. 2. Ecological Studies: Analysis and Synthesis. New York: Springer.CrossRefGoogle Scholar
Benson, L. 1950. The Cacti of Arizona, 2nd edn. Tucson: University of Arizona Press.Google Scholar
Benson, L. 1959. Plant Classification. Lexington: D.C. Heath and Company.Google Scholar
Benzing, D. H. 1990. Vascular Epiphytes: General Biology and Related Biota. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Berbee, M. L. and Taylor, J. W.. 1993. Dating the evolutionary radiations of the true fungi. Canadian Journal of Botany 71: 1114–1127.CrossRefGoogle Scholar
Berenbaum, M. R. 1991. Coumarins. pp. 221–249. In Rosenthal, G. A. and Berenbaum, M. R. (eds.) Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego: Academic Press.Google Scholar
Berg, R. Y. 1975. Myrmecochorous plants in Australia and their dispersal by ants. Australian Journal of Botany 23: 475–508.CrossRefGoogle Scholar
Bernard, H. A. and R. J. Leblanc. 1965. Résumé of the quaternary geology of the northwestern Gulf of Mexico province. pp. 137–185. In Wright, H. E. and Frey, D. G. (eds.) The Quaternary of the United States. Princeton: Princeton University Press.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 XIV: 519–545.Google Scholar
Bertness, M. D. and Hacker, S. D.. 1994. Physical stress and positive associations among marsh plants. The American Naturalist 144: 363–372.CrossRefGoogle Scholar
Bertness, M. D. and Yeh, S. M.. 1994. Cooperative and competitive interactions in the recruitment of marsh elders. Ecology 75: 2416–2429.CrossRefGoogle Scholar
Bessey, C. E. 1915. The phylogenetic taxonomy of flowering plants. Annals of the Missouri Botanical Garden 2: 109–164.CrossRefGoogle Scholar
Bierzychudek, P. 1980. The demographic consequences of sexuality and apomixis in Antennaria. pp 293–307. In Kawano, S. (ed.) Biological Approaches and Evolutionary Trends in Plants. London: Academic Press.Google Scholar
Billings, W. D. and Mooney, H. A.. 1968. The ecology of arctic and alpine plants. Biological Reviews 43: 481–529.CrossRefGoogle Scholar
Binford, M. W., Brenner, M., Whitmore, T. J., Higuera-Gundy, A., Deevey, E. S., and Leyden, B.. 1987. Ecosystems, paleoecology and human disturbance in subtropical and tropical America. Quaternary Science Reviews 6: 115–128.CrossRefGoogle Scholar
Björkman, E. 1960. Monotropa hypopitys L. an epiparasite on tree roots. Physiologia Plantarum 13: 308–327.CrossRefGoogle Scholar
Black, D. (ed.) 1979. Carl Linnaeus: Travels. Nature Classics. New York: Charles Scribner's Sons.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: 1095–1107.CrossRefGoogle Scholar
Blizard, D. 1993. The Normandy Landings D-Day: The Invasion of Europe 6 June 1944. London: Reed International Books.Google 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
Bohlen, P. J., Scheu, S., Hale, C. M., McLean, M. A., Migge, S., Groffman, P. M., and Parkinson, D.. 2004. Invasive earthworms as agents of change in north temperate forests. Frontiers in Ecology and the Environment 8: 427–435.CrossRefGoogle Scholar
Bolan, N. S. 1991. A critical review on the role of mychorrhizal fungi in the uptake of phosphorus by plants. Plant and Soil 134: 189–207.CrossRefGoogle Scholar
Bond, W. J. 1997. Functional types for predicting changes in biodiversity: a case study in Cape fynbos. pp. 174–194. In Smith, T. M., Shugart, H. H., and Woodward, F. I. (eds.) Plant Functional Types. Cambridge: Cambridge University Press.Google Scholar
Boot, R. G. A. 1989. The significance of size and morphology of root systems for nutrient acquisition and competition. pp. 299–311. In Lambert, H.et al. (eds.) Causes and Consequences of Variation in Growth Rate and Productivity of Higher Plants. The Hague: SPB Academic Publishing.Google Scholar
Booth, B. and D. W. Larson. 1999. Impact of history, language, and choice of system on the study of assembly rules. pp. 206–229. In Weiher, E. and Keddy, P. (eds.) Ecological Assembly Rules: Perspectives, Advances, Retreats. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Borhidi, A. 1992. The serpentine flora and vegetation of Cuba. pp. 83–95. In Baker, A. J. M., Proctor, J., and Reeves, R. D. (eds.) The Vegetation of Ultramafic (Serpentine) Soils. Andover: Intercept.Google Scholar
Bormann, B. T. and Sidle, R. C.. 1990. Changes in productivity and distribution of nutrients in a chronosequence at Glacier Bay National Park, Alaska. Journal of Ecology 78: 561–578.CrossRefGoogle Scholar
Bormann, F. H. and Likens, G. E.. 1981. Pattern and Process in a Forested Ecosystem. Second corrected printing. 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: 259–270.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: 615–622.CrossRefGoogle ScholarPubMed
Botkin, D. B. 1977. Life and death in a forest: the computer as an aid to understanding. pp. 213–33. In Hall, A. S. and Day, J. W. (eds.) Ecosystem Modelling in Theory and Practice. New York: John Wiley and Sons.Google Scholar
Botkin, D. B. 1990. Discordant Harmonies: A New Ecology for the Twenty-first Century. New York: Oxford University Press.Google Scholar
Botkin, D. B. 1993. Forest Dynamics. Oxford: Oxford University Press.Google Scholar
Boucher, D. H. 1985a. The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Boucher, D. H. 1985b. The idea of mutualism, past and future. pp. 1–28. In Boucher, D. H.. The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Boucher, D. H., James, S., and Keeler, K. H.. 1982. The ecology of mutualism. Annual Review of Ecology and Systematics 13: 315–347.CrossRefGoogle Scholar
Boutin, C. and Keddy, P. A.. 1993. A functional classification of wetland plants. Journal of Vegetation Science 4: 591–600.CrossRefGoogle Scholar
Bowers, M. D. 1991. Iridoid glycosides. pp. 297–325. In Rosenthal, G. A. and Berenbaum, M. R. (eds.) Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego: Academic Press.Google Scholar
Boyd, C. E. 1978. Chemical composition of wetland plants. pp. 155–168. In Good, R. E., Whigham, D. F., and Simpson, R. L. (eds.) Freshwater Wetlands: Ecological Processes and Management Potential. 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: 443–452.Google Scholar
Brackenridge, J. B. and Rosenberg, R. M.. 1970. The Principles of Physics and Chemistry. New York: McGraw-Hill.Google Scholar
Braun, E. L. 1950. The Deciduous Forest of Eastern North America. New York: Hafner.Google Scholar
Brewer, J. S. 1998. Effects of competition and litter on a carnivorous plant, Drosera cappilaris (Droseraceae). American Journal of Botany 85 N 11: 1592–1596.CrossRefGoogle Scholar
Brewer, J. S. 1999. Effects of fire, competition and soil disturbances on regeneration of a carnivorous plant (Drosera capillaris). The American Midland Naturalist 141: 28–42.CrossRefGoogle Scholar
Bridges, E. M., Batjes, N. H., and Nachtergaele, F. O. (eds.) 1998. World Reference Base for Soil Resources: Atlas. Leuven, Belgium: ACCO.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 WRP-DE-4. U.S. Army Corps of Engineers, Washington, D.C.Google Scholar
Brooks, R. R., R. D. Reeves, and A. J. M. Baker. 1992. The serpentine vegetation of Goiás State, Brazil. pp. 67–81. In Baker, A. J. M., Proctor, J., and Reeves, R. D. (eds.) The Vegetation of Ultramafic (Serpentine) Soils. Andover: Intercept.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: 151–158.CrossRefGoogle Scholar
Brown, J. H., D. W. Davidson, J. C. Munger, and R. S. Inouye. 1986. Experimental community ecology: the desert granivore system. pp. 41–61. In Diamond, J. and Case, T. J. (eds.) Community Ecology. New York: Harper and Row.Google Scholar
Browne, J. 1995. Charles Darwin: Voyaging. Princeton: Princeton University Press.Google Scholar
Bryant, D., Nielsen, D., and Tangley, L.. 1997. The Last Frontier Forests: Ecosystems and Economies on the Edge. Washington: World Resources Institute.Google Scholar
Burch, W. Jr. 1999. Daydreams and Nightmares – A Sociological Essay on the American Environment. Madison: Social Ecology Press.Google Scholar
Burdon, J. J. 1982. The effect of fungal pathogens on plant communities. pp. 99–112. In Newman, E. I. (ed.) The Plant Community as a Working Mechanism. Oxford: Blackwell.Google Scholar
Burger, J. C. and Louda, S. V.. 1995. Interaction of diffuse competition and insect herbivory in limiting brittle prickly pear cactus, Opuntia fragilis (Cactaceae). American Journal of Botany 82: 1558–1566.CrossRefGoogle Scholar
Burgess, R. L. and Sharpe, D. M. (eds.) 1981. Forest Island Dynamics in Man-dominated Landscapes. New York: Springer-Verlag.CrossRefGoogle Scholar
Buss, L. W. 1988. The Evolution of Individuality. Princeton: Princeton University Press.CrossRefGoogle Scholar
Cairns, J. (ed.) 1980. The Recovery Process in Damaged Ecosystems. Ann Arbor: Ann Arbor Science.Google Scholar
Cairns, J. (ed.) 1988. Rehabilitating Damaged Ecosystems. Vol. 1 and 2. Boca Raton: CRC Press.Google Scholar
Cairns, J. 1989. Restoring damaged ecosystems: is predisturbance condition a viable option?The Environmental Professional 11: 152–159.Google Scholar
Cairns-Smith, A. G. 1985. Seven Clues to the Origin of Life: A Scientific Detective Story. Canto edition 1990. Cambridge: Cambridge University Press.Google Scholar
Callaway, R. M. and King, L.. 1996. Temperature-driven variation in substrate oxygenation and the balance of competition and facilitation. Ecology 77: 1189–1195.CrossRefGoogle Scholar
Campbell, B. D., Grime, J. P., and Mackey, J. M. L.. 1991. A trade-off between scale and precision in resource foraging. Oecologia 87: 532–538.CrossRefGoogle ScholarPubMed
Campbell, B. D., Grime, J. P., and Mackey, J. M. L.. 1992. Shoot thrust and its role in plant competition. Journal of Ecology 80: 633–641.CrossRefGoogle Scholar
Canfield, R. H. 1948. Perennial grass composition as an indicator of condition of southwestern mixed grass ranges. Ecology 29: 190–204.CrossRefGoogle Scholar
Caputa, J. 1948. Untersuchungen über die Entwicklung einiger Gräser und Kleearten in Reinsaat und Mischung. Landwirtschaftliches Jahrbuch der Schweiz 62: 848–975.Google Scholar
Carleton, T. J. and MacLellan, P.. 1994. Woody vegetation responses to fire versus clear-cutting logging: a comparative survey in the central Canadian boreal forest. Ecoscience 1: 141–152.CrossRefGoogle Scholar
Carpenter, S. R., Chisholm, S. W., Krebs, C. J., Schindler, D. W., and Wright, R. F.. 1995. Ecosystem experiments. Science 269: 324–327.CrossRefGoogle ScholarPubMed
Carroll, G. 1988. Fungal endophytes in stems and leaves: from latent pathogen to mutualistic symboint. Ecology 69: 2–9.CrossRefGoogle Scholar
Carson, R. 1962. Silent Spring. Boston: Houghton Mifflin.Google Scholar
Carson, W. P. and Pickett, S. T. A.. 1990. Role of resources and disturbance in the organization of an old-field plant community. Ecology 71: 226–238.CrossRefGoogle Scholar
Catling, P. M. and Brownell, V. R.. 1995. A review of the alvars of the Great Lakes region: distribution, floristic composition, biogeography and protection. The Canadian Field-Naturalist 109: 143–171.Google Scholar
Catling, P. M. and Brownell, V. R.. 1998. Importance of fire in alvar ecosystems – evidence from the Burnt Lands, Eastern Ontario. The Canadian Field-Naturalist 112: 661–667.Google Scholar
Catling, P. M., Cruise, J. E., McIntosh, K. L., and McKay, S. M.. 1975. Alvar vegetation in southern Ontario. Ontario Field Biologist 29: 1–25.Google Scholar
Cavers, P. B. 1983. Seed demography. Canadian Journal of Botany 61: 3578–3590.CrossRefGoogle Scholar
Chaneton, E. J. and Facelli, J. M.. 1991. Disturbance effects on plant community diversity: spatial scales and dominance hierarchies. Vegetatio 93: 143–156.CrossRefGoogle Scholar
Chapin, F. S. III 1980. The mineral nutrition of wild plants. Annual Review of Ecology and Systematics 11: 233–260.CrossRefGoogle Scholar
Chapin, F. S. III, Vitousek, P. M., and Cleve, K.. 1986. The nature of nutrient limitation in plant communities. The American Naturalist 127: 48–58.CrossRefGoogle Scholar
Chapman, V. J. 1940. The functions of the pneumatophores of Avicennia nitida Jacq. Proceedings of the Linnean Society of London 152: 228–233.CrossRefGoogle Scholar
Charron, D. and Gagnon, D.. 1991. The demography of northern populations of Panax quinquefolium (American ginseng). Journal of Ecology 79: 431–445.CrossRefGoogle Scholar
Cheplick, G. P. 1992. Sibling competition in plants. Journal of Ecology 80: 567–575.CrossRefGoogle Scholar
Christensen, N. L., R. B. Burchell, A. Liggett, and E. L. Simms. 1981. The structure and development of pocosin vegetation. pp. 43–61. In Richardson, C. J. (ed.) Pocosin Wetlands: An Integrated Analysis of Coastal Plain Freshwater Bogs in North Carolina. Stroudsburg, Pennsylvania: Hutchinson Ross Publishing Company.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: 665–691.CrossRefGoogle Scholar
Clarke, 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: 213–234.CrossRefGoogle Scholar
Clay, K. 1990. The impact of parasitic and mutualistic fungi on competitive interactions among plants. pp. 391–412. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Clements, F. E. 1916. Plant Succession: An Analysis of the Development of Vegetation. Pub. 242. Washington, DC: Carnegie Institute.CrossRefGoogle Scholar
Clements, F. E. 1933. Competition in plant societies. In News Service Bulletin. Washington: Carnegie Institution of Washington, April 2, 1933.Google Scholar
Clements, F. E. 1935. Experimental ecology in the public service. Ecology 16: 324–363.CrossRefGoogle Scholar
Clements, F. E. 1936. Nature and structure of climax. Journal of Ecology 24: 254–282.CrossRefGoogle Scholar
Clements, F. E., Weaver, J. E., and Hanson, H. C.. 1929. Plant Competition. Washington, D.C.: Carnegie Institution of Washington.Google Scholar
Cloud, P. 1976. Beginnings of biospheric evolution and their biogeochemical consequences. Paleobiology 2: 351–387.CrossRefGoogle Scholar
Cloudsley-Thompson, J. L. 1996. Biotic Interactions in Arid Lands. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Clymo, R. S. and Duckett, J. G.. 1986. Regeneration of Sphagnum. New Phytologist 102: 589–614.CrossRefGoogle Scholar
Clymo, R. S. and P. M. Hayward. 1982. The ecology of Sphagnum. pp. 229–289. In Smith, A. J. E. (ed.) Bryophyte Ecology. London: Chapman and Hall.CrossRefGoogle Scholar
Cody, M. L. 1993. Do cholla cacti (Opuntia spp., Subgenus Cylindropuntia) use or need nurse plants in the Mojave Desert?Journal of Arid Environments 24: 1–16.CrossRefGoogle Scholar
Coe, M. J., D. L. Dilcher, J. O. Farlow, D. M. Jarzen, and D. A. Russel. 1987. Dinosaurs and land plants. pp. 225–258. In Friis, E., Chaloner, W. G., and Crane, P. R. (eds.) The Origins of Angiosperms and Their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Cole, L. C. 1949. The measurement of interspecific association. Ecology 30: 411–424.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
Coleman, R. G. and C. Jove. 1992. Geological origin of serpentines. pp. 1–17. In Baker, A. J. M., Proctor, J., and Reeves, R. D. (eds.) The Vegetation of Ultramafic (Serpentine) Soils. Andover: Intercept.Google Scholar
Coley, P. D. 1983. Herbivory and defensive characteristics of tree species in a lowland tropical forest. Ecological Monographs 53: 209–233.CrossRefGoogle Scholar
Colinvaux, P. 1978. Why Big Fierce Animals are Rare: An Ecologist's Perspective. Princeton: Princeton University Press.Google Scholar
Colinvaux, P. 1986. Ecology. Toronto: Wiley and Sons.Google Scholar
Colinvaux, P. 1993. Ecology 2. New York: Wiley and Sons.Google Scholar
Colinvaux, P. A., Oliveira, P. E., Moreno, J. E., Miller, M. C., and Bush, M. B.. 1996. A long pollen record from lowland Amazonia: forest and cooling in glacial times. Science 274: 85–88.CrossRefGoogle Scholar
Colinvaux, P. A., Oliveira, P. E., and Bush, M. B.. 2000. Amazonian and Neotropical plant communities on glacial time-scales: the failure of the aridity and refuge hypotheses. Quaternary Science Reviews 19: 141–169.CrossRefGoogle Scholar
Colinvaux, P.A., Irion, G., Räsänen, M. E., Bush, M. B., and Mello, J. A. S. Nunes 2001. A paradigm to be discarded: geological and paleoecological data falsify the Haffer & Prance refuge hypothesis of Amazonian speciation. Amazoniana 16: 609–646.Google Scholar
Collinson, M. E. and J. J. Hooker. 1987. Vegetational and mammalian faunal changes in the Early Tertiary of southern England. pp. 259–304. In Friis, E., Chaloner, W. G., and Crane, P. R. (eds.) The Origins of Angiosperms and Their Biological Consequences. Cambridge: Cambridge 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
Connell, J. H. 1978. Diversity in tropical rain forests and coral reefs. Science 199: 1302–1310.CrossRefGoogle ScholarPubMed
Connell, J. H. 1990. Apparent versus “real” competition in plants. pp. 9–26. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Connell, J. H. and Slatyer, R. O.. 1977. Mechanisms of succession in natural communities and their role in community stability and organization. The American Naturalist 111: 1119–1144.CrossRefGoogle Scholar
Conner, W. H. and M. A. Buford. 1998. Southern deepwater swamps. pp. 261–287. In Messina, M. G. and Conner, W. H. (eds.) Southern Forested Wetlands. Ecology and Management. Boca Raton: Lewis Publishers.Google Scholar
Connolly, J. 1986. On difficulties with replacement-series methodology in mixture experiments. Journal of Applied Ecology 23: 125–137.CrossRefGoogle Scholar
Conservation International. 2006. Biodiversity Hotspots. Tropical Andes. (www.biodiversityhotspots.org/xp/Hotspots/andes/) accessed 24 July 2006.
Corfield, T. F. 1973. Elephant mortality in Tsavo National Park, Kenya. East African Wildlife Journal 11: 339–368.CrossRefGoogle Scholar
Cowling, R. M. 1990. Diversity components in a species-rich area of the Cape Floristic Region. Journal of Vegetation Science 1: 699–710.CrossRefGoogle Scholar
Cowling, R. M. and Samways, M. J.. 1995. Predicting global patterns of endemic plant species richness. Biodiversity Letters 2: 127–131.CrossRefGoogle Scholar
Cowling, R. M., Rundel, P. W., Lamont, B. B., Arroyo, M. K., and Arianoutsou, M.. 1996. Plant diversity in mediterranean-climate regions. Trends in Ecology and Evolution 11: 362–366.CrossRefGoogle ScholarPubMed
Craighead, F. C. Sr. 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
Crandell, D. R. and H. H. Waldron. 1969. Volcanic hazards in the Cascade Range. pp. 5–18. In R. Olson and M. Wallace (eds.) Geologic Hazards and Public Problems. Conference Proceedings. U.S. Government Printing Office.
Crawford, R. M. M. 1982. Physiological response to flooding. pp. 453–477. In Lange, O. L., Nobel, P. S., Osmond, C. B., and Ziegler, H. (eds.) Physiological Plant Ecology II. Encyclopedia of Plant Physiology. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Crawley, M. J. 1983. Herbivory: The Dynamics of Animal/Plant Interactions. Oxford: Blackwell.Google Scholar
Crawley, M. J. and Harral, J. E.. 2001. Scale dependence in plant biodiversity. Science 291: 864–868.CrossRefGoogle ScholarPubMed
Crepet, W. L. and E. M. Friis. 1987. The evolution of insect pollination in angiosperms. pp. 181–201. In Friis, E., Chaloner, W. G., and Crane, P. R. (eds.) The Origins of Angiosperms and Their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Crocker, R. L. and Major, J.. 1955. Soil development in relation to vegetation and surface age at Glacier Bay, Alaska. Journal of Ecology 43: 427–448.CrossRefGoogle Scholar
Cronquist, A. 1991. Asterales. pp. 721–726. In Angiosperms: The Flowering Plants. pp. 596–765, Vol. 13. The New Encyclopaedia Britannica, 15th edn. Chicago: The University of Chicago.Google Scholar
Cronquist, A. 1993. A commentary on the general system of classification of flowering plants. pp. 272–293. In Flora of North America Editorial Committee. Flora of North America, Vol. 1. Introduction. New York: Oxford University Press.Google Scholar
Cyr, H. and Pace, M. L.. 1993. Magnitude and patterns of herbivory in aquatic and terrestrial ecosystems. Nature 361: 148–150.CrossRefGoogle Scholar
Dacey, J. W. H. 1980. Internal winds in water lilies: an adaptation for life in anaerobic sediments. Science 210: 1017–1019.CrossRefGoogle ScholarPubMed
Dacey, J. W. H. 1981. Pressurized ventilation in the yellow water lily. Ecology 62: 1137–1147.CrossRefGoogle Scholar
Dafni, A. 1992. Pollination Ecology: A Practical Approach. Oxford: Oxford University Press.Google Scholar
Dale, M. 1999. Spatial Pattern Analysis in Plant Ecology. Cambridge: Cambridge University Press.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
Darwin, C. 1871. The descent of Man and selection in relation to sex. In Adler, M. J. (ed.) Great Books of the Western World, 2nd edn, Vol. 49. Chicago: Encyclopaedia Britannica.Google Scholar
Darwin, C. 1888. Insectivorous Plants. 2nd edn. London: John Murray. Revised by F. Darwin.CrossRefGoogle Scholar
Darwin, C. R. 1881. The Formation of Vegetable Mould Through the Action of Worms, with Observations on Their Habits. London: Murray.CrossRefGoogle Scholar
Darwin, F. (ed.) 1950. Charles Darwin's Autobiography: With his Notes and Letters Depicting the Growth of the Origin of Species. New York: Henry Schuman.Google Scholar
Daubenmire, R. 1978. Plant Geography: With Special Reference to North America. Physiological Ecology. New York: Academic Press.Google Scholar
Davis, D. W. 2000. Historical perspective on crevasses, levees, and the Mississippi River. In Colten, C. E. (ed.) Transforming New Orleans and its Environs, Centuries of Change. Pittsburgh: University of Pittsburgh Press.Google Scholar
Davies, B. R. and Walker, K. F.. 1986. The Ecology of River Systems. Dordrecht: W. Junk Publishers.CrossRefGoogle Scholar
Davis, M. B. 1976. Pleistocene biogeography of temperate deciduous forests. pp. 13–26. In West, R. C. and Haag, W. G. (eds.) Geoscience and Man, Vol. 13. Ecology of the Pleistocene, a Symposium. Baton Rouge: School of Geoscience, Louisiana State University.Google Scholar
Davis, S. and Ogden, J. (eds.) 1994. Everglades: The Ecosystem and its Restoration. Del Ray Beach: St. Lucie Press.Google Scholar
Dawkins, R. 1976. The Selfish Gene. Oxford: Oxford University Press.Google Scholar
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: 1044–1054.CrossRefGoogle Scholar
Day, W. 1984. Genesis on Planet Earth, 2nd edn. New Haven: Yale University Press.Google Scholar
Dayton, P. K. 1979. Ecology: a science and a religion. pp. 3–18. In Livingston, R. J. (ed.) Ecological Processes in Coastal and Marine Systems. New York: Plenum Press.CrossRefGoogle Scholar
Calesta, D. S. 1994. Effect of white-tailed deer on songbirds within managed forests in Pennsylvania. Journal of Wildlife Management 58: 711–718.CrossRefGoogle Scholar
Deckers, J. A., Nachtergaele, F. O., and Spaargaren, O. C. (eds.) 1998. World Reference Base for Soil Resources: Introduction. Leuven, Belgium: ACCO.Google Scholar
Duve, C. 1991. Blueprint for a Cell: The Nature and Origin of Life. Burlington: Neil Patterson.Google Scholar
Groot, R. S. 1992. Functions of Nature. Groningen: Wolters-Noordhoff.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. 1983. Competition as a control mechanism in subalpine meadows. American Journal of Botany 70: 232–245.CrossRefGoogle Scholar
del Moral, R. and Bliss, L. C.. 1993. Mechanisms of primary succession: insights resulting from the eruption of Mount St. Helens. Advances in Ecological Research 24: 1–66.CrossRefGoogle Scholar
del Moral, R. and S. Y. Grishin. 1999. Volcanic disturbances and ecosystem recovery. pp. 137–160. In Walker, L. R. (ed.) Ecosystems of Disturbed Ground. Ecosystems of the World Series. Amsterdam: Elsevier Science.Google Scholar
del Moral, R. and Wood, D. M.. 1993. Early primary succession on the volcano Mount St. Helens. Journal of Vegetation Science 4: 223–234.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: 107–120.CrossRefGoogle Scholar
Denslow, J. L. 1987. Tropical rain forest gaps and tree species diversity. Annual Review of Ecology and Systematics 18: 431–451.CrossRefGoogle Scholar
Deshmukh, I. 1986. Ecology and Tropical Biology. Palo Alto: Blackwell Scientific.Google Scholar
Desmond, A. and Moore, J.. 1991. Darwin. New York: Warner Books.Google Scholar
Wit, C. T. 1960. On competition. Verslagen van Landbouwkundige Onderzoekingen 66: 1–82.Google Scholar
Diamond, J. M. 1975. Assembly of species communities. pp. 342–444. In Cody, M. L. and Diamond, J. M. (eds.) Ecology and Evolution of Communities. Cambridge: Belknap Press of Harvard University Press.Google Scholar
Diamond, J. M. 1986. Overview: laboratory experiments, field experiments, and natural experiments. pp. 3–22. In Diamond, J. M. and Case, T. J. (eds.) Community Ecology. New York: Harper and Row.Google Scholar
Diamond, J. 1994. Ecological collapses of past civilisations. Proceedings of the American Philosophical Society 138: 363–370.Google Scholar
Diamond, J. 2004. Twilight at Easter. New York Review of Books LI (5) (March 25). pp. 6–10.Google Scholar
Díaz, S., Acosta, A., and Cabido, M.. 1992. Morphological analysis of herbaceous communities under different grazing regimes. Journal of Vegetation Science 3: 689–696.CrossRefGoogle Scholar
Dickerson, R. E. 1969. Molecular Thermodynamics. New York: W. A Benjamin Inc.Google Scholar
Digby, P. G. N. and Kempton, R. A.. 1987. Multivariate Analysis of Ecological Communities. London: Chapman and Hall.Google Scholar
Dilcher, D. L. and Crane, P. R.. 1985. Archaeanthus: an early angiosperm from the Cenomanian of the western interior of North America. Annals of the Missouri Botanical Garden 71: 351–383.CrossRefGoogle Scholar
Dilcher, D. L. and Kovach, W. L.. 1986. Early angiosperm reproduction: Caloda delevoryana gen. et sp. nov., a new fructification from the Dakota Formation (Cenomanian) of Kansas. American Journal of Botany 73: 1230–1237.CrossRefGoogle Scholar
Dinerstein, E. 1991. Seed dispersal by greater one-horned rhinoceros (Rhinoceros unicornis) and the flora of Rhinoceros latrines. Mammalia 55: 355–362.CrossRefGoogle Scholar
Dinerstein, E. 1992. Effects of Rhinoceros unicornis on riverine forest structure in lowland Nepal. Ecology 73: 701–704.CrossRefGoogle Scholar
Dirzo, R., Horvitz, C. C., Quevedo, H., and López, M. A.. 1992. The effects of gap size and age on the understorey herb community of a tropical Mexican rain forest. Journal of Ecology 80: 809–822.CrossRefGoogle Scholar
Dodson, C. H. 1991. Orchidales, pp. 738–746. In Angiosperms: The Flowering Plants. pp. 596–765, Vol. 13. The New Encyclopaedia Britannica, 15th edn. Chicago: The University of Chicago.Google Scholar
Douglas, R. J. W. 1972. Geology and Economic Minerals of Canada. Ottawa: Geological Survey of Canada.Google Scholar
Dowdeswell, J. A. 2006. The Greenland Ice Sheet and global sea-level rise. Science 311: 963–964.CrossRefGoogle ScholarPubMed
Dressler, R. L. 1983. Classification of the Orchidaceae and their probable origin. Telopea 2: 413–424.CrossRefGoogle Scholar
Drew, M. C. 1975. Comparison of the effects of a localized supply of phosphate, nitrate, ammonium and potassium on the growth of the seminal root system, and the shoot, in barley. New Phytologist 75: 479–490.CrossRefGoogle Scholar
Duchesne, L. C. and Larson, D. W.. 1989. Cellulose and the evolution of plant life. Bioscience 39: 238–241.CrossRefGoogle Scholar
Dugan, P. (ed.) 1993. Wetlands in Danger. New York: Oxford University Press.Google 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: 403–416.CrossRefGoogle Scholar
Durant, W. 1944. Caesar and Christ. New York: Simon and Schuster.Google Scholar
du Rietz, G. E. 1931. Life-forms of Terrestrial Flowering Plants. Acta Phytogeographica Suecia. III. Uppsala: Almqvist and Wiksells.Google Scholar
Dynesius, M. and Nilsson, C.. 1994. Fragmentation and flow regulation of river systems in the northern third of the world. Science 266: 753–762.CrossRefGoogle Scholar
Earth Impact Database. 2006. (http://www.unb.ca/passc/ImpactDatabase) accessed 12 Sept. 2006.
Edmonds, J. (ed.) 1997. Oxford Atlas of Exploration. New York: Oxford University Press.Google Scholar
Ehrenfeld, J. G. 1983. The effects of changes in land-use on swamps of the New Jersey Pine Barrens. Biological Conservation 25: 353–375.CrossRefGoogle Scholar
Ehrlich, A. and Ehrlich, P.. 1981. Extinction: The Causes and Consequences of the Disappearance of Species. New York: Random House.Google Scholar
Ehrlich, P. and Raven, P. H. 1964. Butterflies and plants: a study in coevolution. Evolution 18: 586–608.CrossRefGoogle Scholar
Eissenstat, D. M. and Newman, E. I.. 1990. Seedling establishment near large plants: effects of vesicular-arbuscular mycorrhizae on the intensity of plant competition. Functional Ecology 4: 95–99.CrossRefGoogle Scholar
Ellenberg, H. 1985. Veränderungen der Flora Mitteleuropas unter dem Einfluß von Düngung und Immissionen. Schweizerische Zeitschrift für Forstwesen 136: 19–39.Google Scholar
Ellenberg, H. 1988a. Floristic changes due to nitrogen deposition in central Europe. In J. Nilsson and P. Grennfelt (eds.) Critical Loads for Sulphur and Nitrogen. Report from a workshop held at Skokloster, Sweden, March 19–24, 1988.
Ellenberg, H. 1988b. Vegetation Ecology of Central Europe. 4th edn. Cambridge: Cambridge University Press. Translated by G. K. Strutt.Google 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: 717–731.CrossRefGoogle Scholar
Elton, C. 1927. Animal Ecology. London: Sidgwick and Jackson Ltd.Google Scholar
Encyclopaedia Britannica. 1991a. Vol. 16. p. 500. Chicago: Encyclopaedia Britannica Inc.
Encyclopaedia Britannica. 1991b. Vol. 12. p. 41. Chicago: Encyclopaedia Britannica Inc.
Encyclopaedia Britannica. 1991c. Vol. 16. p. 481. Chicago: Encyclopaedia Britannica Inc.
Endress, P. K. 1996. Diversity and Evolutionary Biology of Tropical Flowers. Paperback edition (with corrections). Cambridge: Cambridge University Press.Google Scholar
Englert, S. 1970. Islands at the Center of the World: New Light on Easter Island. New York: Charles Scribner. Translated by W. Mulloy.Google Scholar
Environment Canada. 1976. Marine Environmental Data Service, Ocean and Aquatic Sciences. Monthly and Yearly Mean Water Levels, Vol. 1. Inland. Ottawa: Department of Environment.
Eriksson, O. 1993. The species-pool hypothesis and plant community diversity. Oikos 68: 371–374.CrossRefGoogle Scholar
Ernst, W. 1978. Discrepancy between ecological and physiological optima of plant species: a re-interpretation. Oecologia Plantarum 13: 175–188.Google Scholar
Estill, J. C. and Cruzan, M. B.. 2001. Phytogeography and rare plant species endemic to the southeastern United States. Castanea 66: 3–23.Google 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. The American Midland Naturalist 121: 125–133.CrossRefGoogle Scholar
Farjon, A. 1998. World Checklist and Bibliography of Conifers. Royal Botanical Gardens at Kew, Richmond, UK.
Farrow, E. P. 1917. On the ecology of the vegetation of Breckland. III. General effects of rabbits on the vegetation. Journal of Ecology 5: 1–18.CrossRefGoogle Scholar
Faulkner, S. P. and C. J. Richardson. 1989. Physical and chemical characteristics of freshwater wetland soils. pp. 41–72. In Hammer, D. A. (ed.) Constructed Wetlands for Wastewater Treatment. Municipal, Industrial, and Agricultural. Chelsea: Lewis Publishers.Google Scholar
Fedorov, A. V., Dekens, P. S., McCarthy, M., Ravelo, A. C., deMenocal, P. B., Barreiro, M., Pacanowski, R. C., and Philander, S. G.. 2006. The Pliocene paradox (mechanisms for a permanent El Niño). Science 312: 1485–1489.CrossRefGoogle Scholar
Feinsinger, P. 1976. Organisation of a tropical guild of nectivorous birds. Ecological Monographs 46: 257–291.CrossRefGoogle Scholar
Feinsinger, P. 1993. Coevolution and pollination. pp. 282–310. In Futuyma, D. J. and Slatkin, M. (eds.) Coevolution. Sunderland: Sinauer.Google Scholar
Fernald, M. L. 1921. The Gray Herbarium expedition to Nova Scotia 1920. Rhodora 23: 89–111, 130–171, 184–195, 233–245, 257–278, 284–300.Google Scholar
Fernald, M. L. 1922. Notes on the flora of western Nova Scotia 1921. Rhodora 24: 157–164, 165–180, 201–208.Google Scholar
Fernald, M. L. 1935. Critical plants of the upper Great Lakes region of Ontario and Michigan. Rhodora 37: 197–222, 238–262, 272–301, 324–341.Google Scholar
Fernández-Armesto, F. 1989. The Spanish Armada: The Experience of War in 1588. Oxford: Oxford University Press.Google Scholar
Ferris, T. 1988. Coming of Age in the Milky Way. Anchor Books edition 1989. New York: Doubleday.Google Scholar
Fienberg, S. E. and Hinkley, D. V. (eds.). 1980. R. A. Fisher: An Appreciation. New York: Springer-Verlag.CrossRefGoogle Scholar
Firbank, L. G. and Watkinson, A. R.. 1985. On the analysis of competition within two-species mixtures of plants. Journal of Applied Ecology 22: 503–517.CrossRefGoogle Scholar
Fischer, R., W. De Vries, W. Seidling, P. Kennedy, and M. Lorenz. 2000. Forest Condition in Europe. 2000 Executive Report. United Nations Economic Commission for Europe/European Commission, Geneva and Brussels.
Fisher, R. A. 1925. Statistical Methods for Research Workers. London: Oliver and Boyd.Google Scholar
Fitter, A. H. and Hay, R. K. M.. 1983. Environmental Physiology of Plants. London: Academic Press.Google Scholar
Flannery, T. 2001. The Eternal Frontier: An Ecological History of North America and its Peoples. Melbourne: Text Publishing.Google Scholar
Flannery, T. 2005. The Weather Makers: How Man is Changing the Climate and What it Means for Life on Earth. New York: Atlantic Monthly Press.Google Scholar
Fleischner, T. L. 1994. Ecological costs of livestock grazing in western North America. Conservation Biology 8: 629–644.CrossRefGoogle Scholar
Flint, R. F. 1971. Glacial and Quaternary Geology. New York: John Wiley and Sons.Google Scholar
Fonteyn, P. J. and Mahall, B. E.. 1978. Competition among desert perennials. Nature 275: 544–545.CrossRefGoogle Scholar
Fonteyn, P. J. and Mahall, B. E.. 1981. An experimental analysis of structure in a desert plant community. Journal of Ecology 69: 883–896.CrossRefGoogle Scholar
Forde, B. and Zhang, H.. 1998. Response: nitrate and root branching. Trends in Plant Science 3: 204–205.CrossRefGoogle Scholar
Foreman, D. 2004. Rewilding North America: A Vision for Conservation in the 21st Century. Washington, D.C.: Island Press.Google Scholar
Forman, R. T. T. 1964. Growth under controlled conditions to explain the hierarchical distributions of a moss, Tetraphis pellucida. Ecological Monographs 34: 1–25.CrossRefGoogle Scholar
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: Island Press.Google Scholar
Foster, A. S. and Gifford, E. M. Jr. 1974. Comparative Morphology of Vascular Plants. 2nd edn. San Francisco: W. H. Freeman and Company.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. Jr. 1990. Role of ecosystem development and climate change in bog formation in central Sweden. Ecology 71: 450–463.CrossRefGoogle Scholar
Fowler, N. 1981. Competition and coexistence in a North Carolina grassland. II. The effects of the experimental removal of species. Journal of Ecology 69: 843–845.CrossRefGoogle Scholar
Fox, J. F. 1977. Alternation and coexistence of tree species. The American Naturalist 111: 69–89.CrossRefGoogle Scholar
Fragoso, J. M. V. 1997. Tapir-generated seed shadows: scale-dependent patchiness in the Amazon rain forest. Journal of Ecology 85: 519–529.CrossRefGoogle Scholar
Francis, R. and Read, D. J.. 1984. Direct transfer of carbon between plants connected by vesicular-arbuscular mycorrhizal mycelium. Nature 307: 53–56.CrossRefGoogle Scholar
Franco, A. C. and Nobel, P. S.. 1989. Effect of nurse plants on the microhabit and growth of cacti. Journal of Ecology 77: 870–886.CrossRefGoogle Scholar
Fraser, L. H. and Keddy, P. 1997. The role of experimental microcosms in ecological research. Trends in Ecology and Evolution 12: 478–481.CrossRefGoogle ScholarPubMed
Fraser, L. H. and Keddy, P. A. (eds.). 2005. The World's Largest Wetlands: Ecology and Conservation.Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Freedman, B. 1995. Environmental Ecology. 2nd edn. San Diego: Academic Press.Google Scholar
Freedman, B., Zobens, W., Hutchinson, T. C., and Gizyn, W. I.. 1990. Intense, natural pollution affects arctic tundra vegetation at the Smoking Hills, Canada. Ecology 71: 492–503.CrossRefGoogle Scholar
Freemark, K. E. and Merriam, H. G.. 1986. The importance of area and habitat heterogeneity to bird assemblages in temperate forest fragments. Biological Conservation 36: 115–141.CrossRefGoogle Scholar
French, B. M. 1998. Traces of Catastrophe: A Handbook of Shock-Metamorphic Effects in Terrestrial Meteoric Impact Structures. LPI Contribution No. 954. Houston: Lunar and Planetary Institute.Google Scholar
Fretwell, S. D. 1977. The regulation of plant communities by food chains exploiting them. Perspectives in Biology and Medicine 20: 169–185.CrossRefGoogle Scholar
Frey, R. W. and P. B. Basan. 1978. Coastal salt marshes. pp. 101–169. In Davis, R. A. (ed.) Coastal Sedimentary Environments. New York: Springer-Verlag.CrossRefGoogle Scholar
Frey, T. E. 1973. The Finnish school and forest site types. pp. 403–433. In Whittaker, R. H. (ed.) Ordination and Classification of Communities. Part V. The Hague: W. Junk.CrossRefGoogle Scholar
Friedmann, E. I. 1982. Endolithic microorganisms in the antarctic cold desert. Science 215: 1045–1053.CrossRefGoogle ScholarPubMed
Friis, E. M., Pedersen, K. R., and Crane, P. R.. 2006. Cretaceous angiosperm flowers: Innovation and evolution in plant reproduction. Palaeogeography, Palaeoclimatology, Palaeoecology 23: 251–293.CrossRefGoogle Scholar
Frontier, S. 1985. Diversity and structure in aquatic ecosystems. Oceanography and Marine Biology Annual Review 23: 253–312.Google Scholar
Futuyma, D. J. and M. Slatkin. 1993. The study of coevolution. pp. 459–464. In Futuyma, D. J. and Slatkin, M. (eds.) Coevolution. Sunderland: Sinauer.Google Scholar
Galatowitsch, S. M. and Valk, A. G.. 1994. Restoring Prairie Wetlands: An Ecological Approach. Ames: Iowa State University Press.Google Scholar
Gardner, G. 1977. The reproductive capacity of Fraxinus excelsior on the Derbyshire limestone. Journal of Ecology 65: 107–118.CrossRefGoogle Scholar
Gaston, K. J. 2000. Global patterns in biodiversity. Nature 405: 220–227.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-Biological Sciences 260: 149–154.CrossRefGoogle Scholar
Gauch, H. G. Jr. 1982. Multivariate Analysis in Community Ecology. Cambridge Studies in Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Gauch, H. G. Jr. and Wentworth, T. R.. 1976. Canonical correlation analysis as an ordination technique. Vegetatio 33: 17–22.CrossRefGoogle Scholar
Gauch, H. G. Jr. and Whittaker, R. H.. 1972. Comparison of ordination techniques. Ecology 53: 868–875.CrossRefGoogle Scholar
Gauch, H. G. Jr., Whittaker, R. H., and Wentworth, T. R.. 1977. A comparative study of reciprocal averaging and other techniques. Journal of Ecology 65: 157–174.Google Scholar
Gaudet, C. L. 1993. Competition in shoreline plant communities: a comparative approach. PhD thesis. Ottawa: University of Ottawa.
Gaudet, C. L. and Keddy, P. A.. 1988. A comparative approach to predicting competitive ability from plant traits. Nature 334: 242–243.CrossRefGoogle Scholar
Gause, G. F. and Witt, A. A.. 1935. Behavior of mixed populations and the problem of natural selection. The American Naturalist 69: 596–609.CrossRefGoogle Scholar
Geis, J. W. 1985. Environmental influences on the distribution and composition of wetlands in the Great Lakes basin. pp. 15–31. In Prince, H. H. and D'Itri, F. M. (eds.) Coastal Wetlands. Chelsea: 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
Gibson, A. C. and Nobel, P. S.. 1986. The Cactus Primer. Cambridge: Harvard University Press.CrossRefGoogle Scholar
Gibson, C. W. D. and Hamilton, J.. 1983. Feeding ecology and seasonal movements of giant tortoises on Aldabra atoll. Oecologia 56: 84–92.CrossRefGoogle ScholarPubMed
Gidley, I. and Shears, R.. 1986. The Rainbow Warrior Affair. Toronto: Irwin Publishing.Google Scholar
Given, D. R. and Soper, J.. 1981. The Arctic-Alpine Element of the Vascular Flora at Lake Superior. Publications in Botany No. 10. Ottawa: National Museums of Canada.Google Scholar
Givnish, T. J. 1982. On the adaptive significance of leaf height in forest herbs. The American Naturalist 120: 353–381.CrossRefGoogle Scholar
Givnish, T. J. 1984. Leaf and canopy adaptations in tropical forests. pp. 51–84. In Medina, E., Mooney, H. A., and Vásquez-Yánes, C. (eds.) Physiological Ecology of Plants of the Wet Tropics. The Hague: Dr. Junk.Google Scholar
Givnish, T. J. 1987. Comparative studies of leaf form: assessing the relative roles of selective pressures and phylogenetic constraints. New Phytologist 106 (Suppl.): 131–160.CrossRefGoogle Scholar
Givnish, T. J. 1988. Ecology and evolution of carnivorous plants. pp. 243–290. In Abrahamson, W. B. (ed.) Plant-Animal Interactions. New York: McGraw-Hill.Google Scholar
Givnish, T. J. 1994. Does diversity beget stability?Nature 371: 113–114.CrossRefGoogle Scholar
Glaser, P. H. 1992. Raised bogs in eastern North America – regional controls for species richness and floristic assemblages. Journal of Ecology 80: 535–554.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: 1021–1048.CrossRefGoogle Scholar
Gleason, H. A. 1917. The structure and development of the plant association. Bulletin of the Torrey Botanical Club 44: 463–481.CrossRefGoogle Scholar
Gleason, H. A. 1926. The individualistic concept of the plant association. Bulletin of theTorrey Botanical Club 53: 7–26.CrossRefGoogle Scholar
Glenn-Lewin, D. C., Peet, R. K., and Veblen, T. T. (eds.) 1992. Plant Succession: Theory and Prediction. Population and Community Biology. No. 11. London: Chapman and Hall.Google Scholar
Glooschenko, W. A. 1980. Coastal salt marshes in Canada. pp. 39–47. In C. D. A. Rubec and F. C. Pollet (eds.) Proceedings of the Workshop on Canadian Wetlands. Saskatoon, Saskatchewan. Environment Canada, Lands Directorate, Ecological Land Class. Series No. 12.
Gnanadesikan, R. 1997. Methods for Statistical Data Analysis of Multivariate Observations. 2nd edn. New York: Wiley.CrossRefGoogle Scholar
Goebel, K. 1905. Wilhelm Hofmeister. The Plant World 8: 291–298.Google Scholar
Goldberg, D. E. 1982a. The distribution of evergreen and deciduous trees relative to soil type: an example from the Sierra Madre, Mexico, and a general model. Ecology 63: 942–951.CrossRefGoogle Scholar
Goldberg, D. E. 1982b. Comparison of factors determining growth rates of deciduous vs. broad-leaf evergreen trees. The American Midland Naturalist 108: 133–143.CrossRefGoogle Scholar
Goldberg, D. E. 1990. Components of resource competition in plant communities. pp. 27–49. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.
Goldberg, D. E. and Landa, K.. 1991. Competitive effect and response: hierarchies and correlated traits in the early stages of competition. Journal of Ecology 79: 1013–1030.CrossRefGoogle Scholar
Goldberg, D. E. and Werner, P. A.. 1983. Equivalence of competitors in plant communities: a null hypothesis and a field experimental approach. American Journal of Botany 70: 1098–1104.CrossRefGoogle Scholar
Goldsmith, F. B. 1973a. The vegetation of exposed sea cliffs at South Stack, Anglesey: I. The multivariate approach. Journal of Ecology 61: 787–818.CrossRefGoogle Scholar
Goldsmith, F. B. 1973b. The vegetation of exposed sea cliffs at South Stack, Anglesey: II. Experimental studies. Journal of Ecology 61: 819–829.CrossRefGoogle Scholar
Goldsmith, F. B. 1978. Interaction (competition) studies as a step towards the synthesis of sea-cliff vegetation. Journal of Ecology 66: 921–931.CrossRefGoogle Scholar
Goldsmith, F. B. and C. M. Harrison. 1976. Description and analysis of vegetation. pp. 85–155. In Chapman, S. B. (ed.) Methods in Plant Ecology. Oxford: Blackwell Scientific.Google Scholar
Gopal, B. 1990. Nutrient dynamics of aquatic plant communities. pp. 177–197. In Gopal, B. (ed.) Ecology and Management of Aquatic Vegetation in the Indian Subcontinent. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Gopal, B. and Goel, U.. 1993. Competition allelopathy in aquatic plant communities. The Botanical Review 59: 155–210.CrossRefGoogle Scholar
Gopal, B., J. Kvet, H. Loffler, V. Masing, and B. C. Patten. 1990. Definition and classification. pp. 9–15. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 1. Natural and Human Relationships. The Hague: SPB Academic Publishing.Google Scholar
Gore, A. J. P. 1983. Introduction. In Gore, A. J. P. (ed.) Ecosystems of the World 4A. Mires: Swamp, Bog, Fen and Moor. Amsterdam: Elsevier Scientific Publishing Company.Google Scholar
Gore, A. 2006. An Inconvenient Truth. The Planetary Emergency of Global Warming and What We Can Do About It. New York: Melcher Media/Rodale.Google Scholar
Goremykin, V. V., Hirsch-Ernst, K. I., Wölfl, S., and Hellwig, F. H.. 2003. Analysis of the Amborella trichopoda chloroplast genome sequence suggests that Amborella is not a basal angiosperm. Molecular Biology and Evolution 20: 1499–1505.CrossRefGoogle Scholar
Gorham, E. 1953. Some early ideas concerning the nature, origin and development of peat lands. Journal of Ecology 41: 257–274.CrossRefGoogle Scholar
Gorham, E. 1957. The development of peat lands. The Quarterly Review of Biology 32: 145–166.CrossRefGoogle Scholar
Gorham, E. 1979. Shoot height, weight and standing crop in relation to density of nonspecific plant stands. Nature 279: 148–150.CrossRefGoogle Scholar
Gorham, E. 1990. Biotic impoverishment in northern peatlands. pp. 65–98. In Woodwell, G. M. (ed.) The Earth in Transition. Cambridge: Cambridge University Press.Google Scholar
Gorham, E. 1991. Northern peatlands role in the carbon cycle and probable responses to climatic warming. Ecological Applications 1: 182–195.CrossRefGoogle ScholarPubMed
Gosselink, J. G. and R. E. Turner. 1978. The role of hydrology in freshwater wetland ecosystems. pp. 63–78. In Good, R. E., Whigham, D. F., and Simpson, R. L. (eds.) Freshwater Wetlands: Ecological Processes and Management Potential. New York: Academic Press.Google Scholar
Gosselink, J. G., J. M. Coleman, and R. E. Stewart, Jr. 1998. Coastal Louisiana. pp. 385–436. In Mac, M. J., Opler, P. A., Haecker, C. E. Puckett, and Doran, P. D. (eds.) 1998. Status and Trends of the Nation's Biological Resources, 2 Vols. Reston: U.S. Department of the Interior, U.S. Geological Survey.Google Scholar
Gotelli, N. J. and Graves, G. R.. 1996. Null Models in Ecology. Washington, D.C.: Smithsonian Institution Press.Google 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 Manchester4: 1–19. In D. Walker. 1970. Direction and rate in some British post-glacial hydoseres. pp. 117–139. In Walker, D. and West, R. G. (eds.) Studies in the Vegetational History of the British Isles. Cambridge: Cambridge University Press.Google Scholar
Gough, L., Grace, J. B., and Taylor, K. L.. 1994. The relationship between species richness and community biomass: the importance of environmental variables. Oikos 70: 271–279.CrossRefGoogle Scholar
Gould, S. J. 1977. Ever Since Darwin: Reflections in Natural History. New York: W. W. Norton and Company.Google Scholar
Grace, J. B. 1993. The effects of habitat productivity on competition intensity. Trends in Ecology and Evolution 8: 229–230.CrossRefGoogle ScholarPubMed
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. 2001. The roles of community biomass and species pools in the regulation of plant diversity. Oikos 92: 193–207.CrossRefGoogle Scholar
Grace, J. B. 2006. Structural Equation Modeling and Natural Systems. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Grace, J. B. and Pugesek, B. H.. 1997. A structural equation model of plant species richness and its application to a coastal wetland. The American Naturalist 149: 436–460.CrossRefGoogle Scholar
Grace, J. B. and Tilman, D. (eds.) 1990. Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Grant, M. C. 1993. The trembling giant. Discover 4(10): 82–89.Google Scholar
Green, P. T., O'Dowd, D. J., and Lake, P. S.. 1997. Control of seedling recruitment by land crabs in rain forest on a remote oceanic island. Ecology 78: 2472–2486.CrossRefGoogle Scholar
Greenslade, P. J. M. 1983. Adversity selection and the habitat templet. Nature 242: 344–347.Google Scholar
Greig-Smith, P. 1952. Use of random and contiguous quadrats in the study of the structure of plant communities. Annals of Botany 16: 293–316.CrossRefGoogle Scholar
Greig-Smith, P. 1957. Quantitative Plant Ecology. London: Butterworths.Google Scholar
Grime, J. P. 1973a. Control of species density in herbaceous vegetation. Journal of Environmental Management 1: 151–167.Google Scholar
Grime, J. P. 1973b. Competitive exclusion in herbaceous vegetation. Nature 242: 344–347.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: 1169–1194.CrossRefGoogle Scholar
Grime, J. P. 1979. Plant Strategies and Vegetation Processes. Chichester: John Wiley.Google Scholar
Grime, J. P. 1994. The role of plasticity in exploiting environmental heterogeneity. pp. 1–19. In Caldwell, M. M. and Percy, R. W. (eds.) Exploitation of Environmental Heterogeneity by Plants: Ecophysical Processes Above- and Belowground. San Diego: Academic Press.Google Scholar
Grime, J. P. 1997. The humped-back model: a response to Oksanen. Journal of Ecology 85: 97–98.CrossRefGoogle Scholar
Grime, J. P. 2002. Declining plant diversity: empty niches or functional shifts?Journal of Vegetation Science 13: 457–460.CrossRefGoogle 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. and Jeffrey, D. W.. 1965. Seedling establishment in vertical gradients of sunlight. Journal of Ecology 53: 621–642.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: 1017–1059.CrossRefGoogle Scholar
Grime, J. P., Mackey, J. M. L., Hillier, S. H., and Read, D. J.. 1987. Floristic diversity in a model system using experimental microcosms. Nature 328: 420–422.CrossRefGoogle 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: 129–153.CrossRefGoogle Scholar
Groombridge, B. (ed.). 1992. Global Biodiversity: Status of the Earth's Living Resources. London: Chapman and Hall.CrossRefGoogle Scholar
Grover, A. M. and Baldassarre, G. A.. 1995. Bird species richness within beaver ponds in south-central New York. Wetlands 15: 108–118.CrossRefGoogle Scholar
Grubb, P. J. 1977. The maintenance of species-richness in plant communities: the importance of the regeneration niche. Biological Reviews 52: 107–145.CrossRefGoogle Scholar
Grubb, P. J. 1987. Global trends in species-richness in terrestrial vegetation: a view from the Northern Hemisphere. pp. 99–118. In Gee, J. H. R. and Giller, P. S. (eds.) Organization of Communities Past and Present. Oxford: Blackwell Scientific Publications.Google Scholar
Grumbine, R. E. 1997. Reflections on “What is ecosystem management?”Conservation Biology 11: 41–47.CrossRefGoogle Scholar
Guariguata, M. R. 1990. Landslide disturbance and forest regeneration in the Upper Luquillo mountains of Puerto Rico. Journal of Ecology 78: 814–832.CrossRefGoogle Scholar
Guerlac, H. 1975. Antoine-Laurent Lavoisier, Chemist and Revolutionary. New York: Charles Scribner's Sons.Google Scholar
Gurevitch, J. and Unnasch, R. S.. 1989. Experimental removal of a dominant species at two levels of soil fertility. Canadian Journal of Botany 67: 3470–3477.CrossRefGoogle Scholar
Haber, L. F. 1986. The Poisonous Cloud. Chemical Warfare in the First World War. Oxford: Clarendon Press.Google Scholar
Haffer, J. 1969. Speciation in Amazonian forest birds. Science 165: 131–137.CrossRefGoogle ScholarPubMed
Hairston, N. G., Smith, F. E., and Slobodkin, L. B.. 1960. Community structure, population control, and competition. The American Naturalist XCIV: 421–425.CrossRefGoogle Scholar
Hamann, O. 1979. Regeneration of vegetation on Santa Fe and Pinta Islands, Galápagos, after the eradication of goats. Biological Conservation 15: 215–236.CrossRefGoogle Scholar
Hamann, O. 1993. On vegetation recovery, goats and giant tortoises on Pinta Island, Galápagos, Ecuador. Biodiversity and Conservation 2: 138–151.CrossRefGoogle Scholar
Harper, J. L. 1965. The nature and consequence of interference amongst plants. Genetics Today 2: 465–482.Google Scholar
Harper, J. L. 1967. A Darwinian approach to plant ecology. Journal of Ecology 55: 247–270.CrossRefGoogle Scholar
Harper, J. L. 1977. Population Biology of Plants. London: Academic Press.Google Scholar
Harper, J. L. 1982. After description. pp. 11–25. In Newman, E. I. (ed.) The Plant Community as a Working Mechanism. Oxford: Blackwell.Google Scholar
Harper, J. L. and Ogden, J.. 1970. The reproductive strategy of higher plants. I. The concept of strategy with special reference to Senecio vulgaris L. Journal of Ecology 58: 681–698.CrossRefGoogle Scholar
Harper, J. L. and White, J.. 1974. The demography of plants. Annual Review of Ecology and Systematics 5: 419–463.CrossRefGoogle Scholar
Harper, J. L., Rosen, B. R., and White, J.. 1986. The Growth and Form of Modular Organisms. London: The Royal Society.Google Scholar
Harris, L. D. 1984. The Fragmented Forest: Island Biogeography Theory and the Preservation of Biotic Diversity. Chicago: University of Chicago Press.Google Scholar
Hartman, J. M. 1988. Recolonization of small disturbance patches in a New England salt marsh. American Journal of Botany 75: 1625–1631.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
Hatch, A. B. 1937. The physical basis of mycotrophy in Pinus. The Black Rock Forest Bulletin, No. 6. 17 pp.Google Scholar
Hawksworth, D. L. 1988. Coevolution of fungi with algae and cyanobacteria in lichen symbioses. pp. 125–148. In Pirozynski, K. A. and Hawksworth, D. L. (eds.) Coevolution of Fungi with Plants and Animals. London: Academic Press.Google Scholar
Hawksworth, D. L. 1990. The fungal dimension of biodiversity: magnitude, significance, and conservation. Mycological Research 95: 641–655.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
Heady, H. F. (ed.) 1988. The Vale Rangeland Rehabilitation Program: An Evaluation. USDA Forest Service, Resource Bulletin PNW-RB-157, 151 pp.
Heady, H. F. and J. Bartolome. 1977. The Vale Rangeland Rehabilitation Program: The Desert Repaired in Southeastern Oregon. USDA Forest Service, Resource Bulletin PHW-70, 139 pp.
Heckman, D. S., Geiser, D. M., Eidell, B. R., Stauffer, R. L., Kardos, N. L., and Hedges, S. B.. 2001. Molecular evidence for the early colonization of land by fungi and plants. Science 293: 1129–1133.CrossRefGoogle ScholarPubMed
Heinselman, M. L. 1973. Fire in the virgin forests of the Boundary Waters Canoe Area, Minnesota. Quaternary Research 3: 329–382.CrossRefGoogle Scholar
Heinselman, M. L. 1981. Fire and succession in the conifer forests of northern North America. pp. 374–405. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Applications. New York: Springer-Verlag.CrossRefGoogle Scholar
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: 378–382.CrossRefGoogle Scholar
Higgs, E. S. 1997. What is good ecological restoration?Conservation Biology 11: 338–348.CrossRefGoogle Scholar
Hill, N. M. and Keddy, P. A.. 1992. Predicting numbers of rarities from habitat variables: coastal plain plants of Nova Scotian lakeshores. Ecology 73: 1852–1859.CrossRefGoogle Scholar
Hills, G. A. 1961. The Ecological Basis for Land-Use Planning. Report No. 46. Ontario: Ontario Department of Lands and Forests, Research Branch.
Hoagland, B. W. and Collins, S. L.. 1997. Gradient models, gradient analysis, and hierarchical structure in plant communities. Oikos 78: 23–30.CrossRefGoogle Scholar
Hoffman, T. M., Midgley, G. F., and Cowling, R. M.. 1994. Plant richness is negatively related to energy availability in semi-arid southern Africa. Biodiversity Letters 2: 35–38.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: 1991–1997.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
Holechek, J. L., Vavra, M., and Pieper, R. D.. 1982. Botanical composition determination of herbivore diets: a review. Journal of Range Management 31: 309–315.CrossRefGoogle Scholar
Holling, C. S. 1959. The components of predation as revealed by a study of small-mammal predation of the European pine sawfly. Canadian Entomologist 91: 293–320.CrossRefGoogle Scholar
Holling, C. S. (ed.) 1978a. Adaptive Environmental Assessment and Management. New York: John Wiley and Sons.Google Scholar
Holling, C. S. 1978b. The spruce-budworm/forest-management problem. pp. 143–182. In Holling, C. S. (ed.) Adaptive Environmental Assessment and Management. New York: John Wiley and Sons.Google Scholar
Holt, R. D. and Lawton, J. H.. 1993. Apparent competition and enemy-free space in insect host-parasitoid communities. The American Naturalist 142: 623–645.CrossRefGoogle ScholarPubMed
Holt, R. D. and Lawton, J. H.. 1994. The ecological consequences of shared natural enemies. Annual Review of Ecology and Systematics 25: 495–520.CrossRefGoogle Scholar
Hook, D. D. 1984. Adaptations to flooding with fresh water. pp. 265–294. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Hopkins, D. M. (ed.) 1967. The Bering Land Bridge. Stanford: Stanford University Press.Google Scholar
Horn, H. S. 1971. The Adaptive Geometry of Trees. Princeton: Princeton University Press.Google Scholar
Horn, H. 1976. Succession. pp. 187–204. In May, R. M. (ed.) Theoretical Ecology: Principles and Applications. Philadelphia: W. B. Saunders.
Horn, H. S. 1981. Some causes of variety of patterns of secondary succession. pp. 24–35. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
Horn, H. S. and MacArthur, R. H.. 1972. Competition among fugitive species in a harlequin environment. Ecology 53: 749–752.CrossRefGoogle Scholar
Houck, O. 2006. Can we save New Orleans?Tulane Environmental Law Journal 19: 1–68.Google Scholar
Hubbell, S. P. and R. B. Foster. 1986. Biology, chance and history and the structure of the tropical rain forest tree communities. pp. 314–329. In Diamond, J. and Case, T. J. (eds.) Community Ecology. New York: Harper and Row.Google Scholar
Huber, H., Lukács, S., and Watson, M. A.. 1999. Spatial structure of stoloniferous herbs: an interplay between structural blue-print, ontogeny and phenotypic plasticity. Plant Ecology 141: 107–115.CrossRefGoogle Scholar
Hughes, J. D. 1982. Deforestation, erosion, and forest management in ancient Greece and Rome. Journal of Forest History 26: 60–75.Google Scholar
Humphries, C. J. 1981. Biogeographical methods and the southern beeches. pp. 283–297. In Forey, P. L. (ed.) The Evolving Biosphere. British Museum (Natural History). Cambridge: London and Cambridge University Press.Google Scholar
Hunt, R., Hand, D. W., Hannah, M. A., and Neal, A. M.. 1991. Response to CO2 enrichment in 27 herbaceous species. Functional Ecology 5: 410–421.CrossRefGoogle 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: 724–732.Google Scholar
Huntley, B. 1990. European post-glacial forests: compositional changes in response to climatic change. Journal of Vegetation Science 1: 507–518.CrossRefGoogle 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: 257–271.CrossRefGoogle Scholar
Huston, M. A. 1979. A general hypothesis of species diversity. The American Naturalist 113: 81–101.CrossRefGoogle Scholar
Huston, M. A. 1994. Biological Diversity. The Coexistence of Species on Changing Landscapes. Cambridge: Cambridge University Press.Google Scholar
Huston, M. A. 1997. Hidden treatments in ecological experiments: re-evaluating the ecosystem function of biodiversity. Oecologia 110: 449–460.CrossRefGoogle ScholarPubMed
Hutchinson, G. E. 1959. Homage to Santa Rosalia; or, why are there so many kinds of animals?The American Naturalist 93: 145–159.CrossRefGoogle Scholar
Hutchinson, G. E. 1970. The biosphere. pp. 194–203. In Wilson, E. O. (ed.) 1974. Ecology, Evolution, and Population Biology. Readings from Scientific American. San Francisco: W.H. Freeman and Company.Google Scholar
Hutchinson, G. E. 1975. A Treatise on Limnology, Vol. 3. Limnological Botany. New York: John Wiley and Sons.Google Scholar
Huxley, C. R. 1980. Symbiosis between plants and epiphytes. Biological Reviews 55: 321–340.CrossRefGoogle Scholar
Imbrie, J., et al. 1992. On the structure and origin of major glaciation cycles 2. The 100,000-year cycle. Paleoceanography 8: 699–736.CrossRefGoogle Scholar
International Joint Commission. 1980. Pollution in the Great Lakes Basin from Land Use Activities. Washington, D.C.: International Joint Commission.
Irion, G. M., Müller, J., Mello, J. N., and Junk, W. J.. 1995. Quaternary geology of the Amazon lowland. Geo-Marine Letters 15: 172–178.CrossRefGoogle Scholar
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, J. B. C., Buss, L. W., and Cook, R. E.. 1985. Population Biology and Evolution of Clonal Organisms. New Haven: Yale University Press.Google Scholar
Jackson, M. B. and M. C. Drew. 1984. Effects of flooding on growth and metabolism of herbaceous plants. pp. 47–128. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Jaksic, F. M. and Fuentes, E. R.. 1980. Why are native herbs in the Chilean matorral more abundant beneath bushes: microclimate or grazing?Journal of Ecology 68: 665–669.CrossRefGoogle Scholar
James, W. 1907. Pragmatism. Reprinted pp. xv–xvii, 1–89. In Adler, M. J. (ed.) 1990. Great Books of the Western World. Vol. 55. Chicago: Encyclopaedia Britannica.Google Scholar
Janis, C. 1976. The evolutionary strategy of the Equidae and the origins of rumen and cecal digestion. Evolution 30: 757–774.CrossRefGoogle ScholarPubMed
Janssens, F., Peeters, A., Tallowin, J. R. B., Bakker, J. P., Fillat, F., and Oomes, M. J. M.. 1998. Relationship between soil chemical factors and grassland diversity. Plant and Soil 202: 69–78.CrossRefGoogle Scholar
Janzen, D. H. 1966. Coevolution of mutualism between ants and acacias in Central America. Evolution 20: 249–275.CrossRefGoogle ScholarPubMed
Janzen, D. H. 1967. Interaction of the bull's-horn acacia (Acacia cornigera L.) with an ant inhabitant (Pseudomyrmex ferruginea F. Smith) in eastern Mexico. The University of Kansas Science Bulletin XLVII: 315–558.Google Scholar
Janzen, D. H. 1971. Seed predation by animals. Annual Review of Ecology and Systematics 2: 465–492.CrossRefGoogle Scholar
Janzen, D. H. 1974. Epiphytic myrmecophytes in Sarawak: mutualism through the feeding of plants by ants. Biotropica 6: 237–259.CrossRefGoogle Scholar
Janzen, D. H. 1976. Why bamboos wait so long to flower. Annual Review of Ecology and Systematics 7: 347–391.CrossRefGoogle Scholar
Janzen, D. H. 1983. Dispersal of seeds by vertebrate guts. pp. 232–262. In Futuyma, D. J. and Slatkin, M. (eds.) Coevolution. Sunderland: Sinauer.Google Scholar
Janzen, D. H. 1985. The natural history of mutualisms. pp. 40–99. In Boucher, D. H. (ed.) The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Janzen, D. H. and Martin, P. S.. 1982. Neotropical anachronisms: the fruits the gomphotheres ate. Science 215: 19–27.CrossRefGoogle ScholarPubMed
Janzen, D. H., Miller, G. A., Hackforth-Jones, J., Pond, C. M., Hooper, K., and Janos, D. P.. 1976. Two Costa Rican bat-generated seed shadows of Andira inermis (Leguminosae). Ecology 57: 1068–1075.CrossRefGoogle Scholar
Jarvis, P. G. 1964. Interference by Deschampsia flexuosa (L.) Trin. Oikos 15: 56–78.CrossRefGoogle Scholar
Jefferies, R. L. 1977. The vegetation of salt marshes at some coastal sites in arctic North America. Journal of Ecology 65: 661–672.CrossRefGoogle Scholar
Jeglum, J. K. 1983. Changes in tree species composition in naturally regenerating strip clearcuts in shallow-soil upland black spruce. pp. 180–193. In Wein, R. W., Riewe, R. R., and Methven, I. R. (eds.) Resources and Dynamics of the Boreal Zone. Ottawa: Association of Canadian Universities for Northern Studies.
Jensen, T. S. 1985. Seed–seed predator interactions of European beech, Fagus silvatica and forest rodents, Clethrionomys glareolus and Apodemus flavicollis. Oikos 44: 149–156.CrossRefGoogle Scholar
Jickells, T. D., Dorling, S., Deuser, W. G., Church, T. M., Arimoto, R., and Propsero, J. M.. 1998. Air-borne dust fluxes to a deep water sediment trap in the Sargasso Sea. Global Biogeochemical Cycles 12: 311–320.CrossRefGoogle Scholar
Johansson, M. E. and Keddy, P. A.. 1991. Intensity and asymmetry of competition between plant pairs of different degrees of similarity: an experimental study on two guilds of wetland plants. Oikos 60: 27–34.CrossRefGoogle Scholar
Johnson, P. L. and Billings, W. D.. 1962. The alpine vegetation of the Beartooth Plateau in relation to cryopedogenic processes and patterns. Ecological Monographs 32: 105–135.CrossRefGoogle Scholar
Johnston, A. E. 1994. The Rothamsted classical experiments. pp. 9–35. In R. A. Leigh and A. E. Johnston (eds.) Long-term Experiments in Agricultural and Ecological Sciences. Proceedings of a conference to celebrate the 150th anniversary of Rothamsted Experimental Station, held at Rothamsted, July 14–17, 1993. Wallingford: CAB International.
Johnston, C. A. and Naiman, R. J.. 1990. Aquatic patch creation in relation to beaver population trends. Ecology 71: 1617–1621.CrossRefGoogle Scholar
Jones, C. G., Lawton, J. H., and Shachak, M.. 1994. Organisms as ecosystem engineers. Oikos 69: 373–386.CrossRefGoogle Scholar
Jones, R. K., G. Pierpoint, G. M. Wickware, and J. K. Jeglum. 1983a. A classification and ordination of forest ecosystems in the Great Claybelt of northeastern Ontario. pp. 83–96. In R. W. Wein, R. R. Riewe, and I. R. Methven (eds.) Resources and Dynamics of the Boreal Zone. Proceedings of a Conference held at Thunder Bay, Ontario, August 1982. Ottawa: Association of Canadian Universities for Northern Studies.
Jones, R. K., Pierpoint, G., Wickware, G. M., Jeglum, J. K., Arnup, R. W., and Bowles, J. M.. 1983b. Field Guide to Forest Classification for the Clay Belt, Site Region 3E. Toronto: Queen's Printer for Ontario.Google Scholar
Jones, W. G., Hill, K. D., and Allen, J. M.. 1995. Wollemia nobilis, a new living Australian genus and species in the Araucariaceae. Telopea 6: 173–176.CrossRefGoogle Scholar
Jordan, C. F., Golley, F. B., Hall, J. D., and Hall, J.. 1980. Nutrient scavenging of rainfall by the canopy of an Amazonian rain forest. Biotropica 12: 61–66.CrossRefGoogle Scholar
Jordan, W. R. III, Gilpin, M. E., and Aber, J. D.. 1987. Restoration Ecology: A Synthetic Approach to Ecological Research. Cambridge: Cambridge University Press.Google Scholar
Judd, W. S., Campbell, C. S., Kellogg, E. A., Stevens, P. F., and Donoghue, M. J.. 2002. Plant Systematics: A Phylogenetic Approach. 2nd edn. Sunderland: Sinauer.Google Scholar
Judson, S. 1968. Erosion of the land, or what's happening to our continents?American Scientist 56: 356–374.Google Scholar
Junk, W. J. 1983. Ecology of swamps on the Middle Amazon. pp. 269–294. In Gore, A. J. P.. (ed.) Ecosystems of the World 4B: Mires: Swamp, Bog, Fen, and Moor. Amsterdam: Elsevier Science.Google Scholar
Jutila, H. M. and Grace, J. B.. 2002. Effects of disturbance on germination and seedling establishment in a coastal prairie grassland: a test of the competitive release hypothesis. Journal of Ecology 90: 291–302.CrossRefGoogle Scholar
Kalamees, K. 1982. The composition and seasonal dynamics of fungal cover on peat soils. pp. 12–29. In Masing, V. (ed.) Peatland Ecosystems: Researches into the Plant Cover of Estonian Bogs and Their Productivity. Tallinn: Academy of Sciences of the Estonian S.S.R.
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
Kaplan, D. R. and Cooke, T. J.. 1996. The genius of Wilhelm Hofmeister: The origin of causal-analytical research in plant development. American Journal of Botany 83: 1647–1660.CrossRefGoogle Scholar
Kastner, T. P., and Goñi, M. A.. 2003. Constancy in the vegetation of the Amazon Basin during the late Pleistocene: evidence from the organic matter composition of Amazon deep sea fan sediments. Geology 31: 291–294.2.0.CO;2>CrossRefGoogle Scholar
Kay, S. 1993. Factors affecting severity of deer browsing damage within coppiced woodlands in the south of England. Biological Conservation 63: 524–532.CrossRefGoogle Scholar
Kearney, T. H. and Shantz, H. L.. 1912. The water economy of dry-land crops. pp. 351–362. Yearbook of the United States Department of Agriculture-1911. Washington: Department of Agriculture.Google Scholar
Keddy, P. A. 1980. Population ecology in an environmental mosaic: Cakile edentula on a gravel bar. Canadian Journal of Botany 58: 1095–1100.CrossRefGoogle Scholar
Keddy, P. A. 1981a. Why gametophytes and sporophytes are different: form and function in a terrestrial environment. The American Naturalist 118: 452–454.CrossRefGoogle Scholar
Keddy, P. A. 1981b. Vegetation with Atlantic coastal plain affinities in Axe Lake, near Georgian Bay, Ontario. The Canadian Field Naturalist 95: 241–248.Google Scholar
Keddy, P. A. 1981c. Experimental demography of the sand dune annual, Cakile edentula, growing along an environmental gradient in Nova Scotia. Journal of Ecology 69: 615–630.CrossRefGoogle Scholar
Keddy, P. A. 1982. Population ecology on an environmental gradient: Cakile edentula on a sand dune. Oecologia 52: 348–355.CrossRefGoogle ScholarPubMed
Keddy, P. A. 1983. Shoreline vegetation in Axe Lake, Ontario: effects of exposure on zonation patterns. Ecology 64: 331–344.CrossRefGoogle Scholar
Keddy, P. A. 1987. Beyond reductionism and scholasticism in plant community ecology. Vegetatio 69: 209–211.CrossRefGoogle Scholar
Keddy, P. A. 1989. Competition.London: Chapman and Hall.CrossRefGoogle Scholar
Keddy, P. A. 1990a. The use of functional as opposed to phylogenetic systematics: a first step in predictive community ecology. pp. 387–406. In Kawano, S. (ed.) Biological Approaches and Evolutionary Trends in Plants. London: Academic Press.Google Scholar
Keddy, P. A. 1990b. Competitive hierarchies and centrifugal organization in plant communities. pp. 265–289. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Keddy, P. A. 1991. Biological monitoring and ecological prediction: from nature reserve management to national state of environment indicators. pp. 249–267. In Goldsmith, F. B. (ed.) Biological Monitoring for Conservation. London: Chapman and Hall.Google Scholar
Keddy, P. A. 1992. Assembly and response rules: two goals for predictive community ecology. Journal of Vegetation Science 3: 157–164.CrossRefGoogle Scholar
Keddy, P. 1994. Reflections on the 21st birthday of MacArthur's Geographical Ecology – applications of the Hertzprung-Russel star diagram to ecology. Trends in Ecology and Evolution 9: 231–234.CrossRefGoogle Scholar
Keddy, P. 1998. Review of Null Models in Ecology (N. J. Gotelli and G. R. Graves, 1996, Smithsonian Institution Press, Washington). The Canadian Field-Naturalist112: 752–754.
Keddy, P. A. 2000. Wetland Ecology: Principles and Conservation. Cambridge: Cambridge University Press.Google Scholar
Keddy, P. A. 2001. Competition. 2nd edn. Dordrecht: Kluwer.CrossRefGoogle Scholar
Keddy, P. A. 2004. Plants matter. Review of The Ecology Of Plants (J. Gurevitch, S. Scheiner and G. A. Fox. 2002. Sinauer Associates, Sunderland, Massachusetts). The Quarterly Review of Biology 79: 55–59.CrossRefGoogle Scholar
Keddy, P. A. 2005a. Milestones in ecological thought – a canon for plant ecology. Journal of Vegetation Science 16: 145–150.Google Scholar
Keddy, P. A. 2005b. Putting the plants back into plant ecology: six pragmatic models for understanding, conserving and restoring plant diversity. Annals of Botany 96: 177–189.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: 122–141.CrossRefGoogle Scholar
Keddy, P. A. and Drummond, C. G.. 1996. Ecological properties for the evaluation, management, and restoration of temperate deciduous forest ecosystems. Ecological Applications 6: 748–762.CrossRefGoogle Scholar
Keddy, P. A. and L. H. Fraser. 2005. Introduction: big is beautiful. pp. 1–10. In Fraser, L. H. and Keddy, P. A. (eds.) The World's Largest Wetlands: Ecology and Conservation. Cambridge: Cambridge University Press.CrossRefGoogle 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 plant communities. Oikos 49: 234–241.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., H. T. Lee, and I. C. Wisheu. 1993. Choosing indicators of ecosystem integrity: wetlands as a model system. pp. 61–79. In Woodley, S., Kay, J., and Francis, G. (eds.) Ecological Integrity and the Management of Ecosystems. Ottawa: 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 enivronments?Journal of Ecology 82: 635–643.CrossRefGoogle Scholar
Keddy, P. A., Nielsen, K., Weiher, E., and Lawson, L. R.. 2002. Relative competitive performance of 63 species of terrestrial herbaceous plants. Journal of Vegetation Science 13: 5–16.CrossRefGoogle Scholar
Keeler, K. H. 1985. Cost:benefit models of mutualism. pp. 100–127. In Boucher, D. H. (ed.) The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Keeley, J. E. 1998. CAM photosynthesis in submerged aquatic plants. The Botanical Review 64: 121–175.CrossRefGoogle Scholar
Keeley, J. E. and Rundel, P. W.. 2003. Evolution of CAM and C4 carbon-concentrating mechanisms. International Journal of Plant Science 164 (Supplement): S55–S77.CrossRefGoogle Scholar
Keeley, J. E., D. A. DeMason, R. Gonzalez, and K. R. Markham. 1994. Sediment-based carbon nutrition in tropical alpine Isoetes. pp. 167–194. In Rundel, P. W., Smith, A. P., and Meinzer, F. C. (eds.) Tropical Alpine Environments Plant Form and Function. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Keeling, C. D. and T. P. Whorf. 2005. Atmospheric CO2 records from sites in the SIO air sampling network. In Trends: A Compendium of Data on Global Change. Carbon Dioxide Information Analysis Center, Oak Ridge National Laboratory, U.S. Department of Energy, Oak Ridge, TN.
Keller, G., Adatte, T., Stinnesbeck, W. et al. 2004. Chicxulub impact predates the K-T boundary mass extinction. Proceedings of the National Academy of Sciences (of the United States of America) 101: 3753–3758.CrossRefGoogle ScholarPubMed
Kellman, M. 1985. Forest seedling establishment in Neotropical savannas: transplant experiments with Xylopia frutescens and Calophyllum brasiliense. Journal of Biogeography 12: 373–379.CrossRefGoogle Scholar
Kellman, M. and Delfosse, B.. 1993. Effect of the red land crab (Gecarcinus lateralis) on leaf litter in a tropical dry forest in Vera Cruz, Mexico. Journal of Tropical Ecology 9: 55–65.CrossRefGoogle Scholar
Kellman, M. and Kading, M.. 1992. Facilitation of tree seedling establishment in a sand dune succession. Journal of Vegetation Science 3: 679–688.CrossRefGoogle Scholar
Kellman, M. and Roulet, N.. 1990. Nutrient flux and retention in a tropical sand-dune succession. Journal of Ecology 78: 664–676.CrossRefGoogle Scholar
Kellner, L. 1963. Alexander von Humboldt. London: Oxford University Press.Google Scholar
Kenrick, P. and Crane, P. R.. 1997. The origin and early evolution of plants on land. Nature 389: 33–39.CrossRefGoogle Scholar
Kershaw, K. A. 1962. Quantitative ecological studies from Landmannahellir, Iceland. Journal of Ecology 50: 171–179.CrossRefGoogle Scholar
Kershaw, K. A. 1973. Quantitative and Dynamic Plant Ecology. 2nd edn. London: Edward Arnold.Google Scholar
Kershaw, K. A. and Looney, J. H. H.. 1985. Quantitative and Dynamic Plant Ecology. 3rd edn. Victoria: Edward Arnold.Google Scholar
Kevan, P. G. 1975. Sun-tracking solar furnaces in high arctic flowers: significance for pollination and insects. Science 189: 723–726.CrossRefGoogle ScholarPubMed
Kidston, R. and Lang, W. H.. 1921. Transactions of the Royal Society Edinburgh LII(IV): 855–902.CrossRef
Killingbeck, K. T. 1996. Nutrients in senesced leaves: keys to the search for potential resorption and resorption efficiency. Ecology 77: 1716–1727.CrossRefGoogle Scholar
King, J. 1997. Reaching for the Sun: How Plants Work. New York: Cambridge University Press.Google Scholar
Kinzig, A. P., Pacala, S., and Tilman, G. D. (eds.) 2002. The Functional Consequences of Biodiversity: Empirical Progress and Theoretical Extensions. Princeton: Princeton University Press.Google Scholar
Knoll, A. H. 1992. The early evolution of eukaryotes: a geological perspective. Science 256: 622–627.CrossRefGoogle ScholarPubMed
Koerselman, W. and Meulman, A. F. M.. 1996. The vegetation N:P ratio: a new tool to detect the nature of nutrient limitation. Journal of Applied Ecology 33: 1441–1450.CrossRefGoogle Scholar
Koyama, H. and Kira, T.. 1956. Intraspecific competition among higher plants. VIII. Frequency distributions of individual plant weight as affected by the interaction between plants. Journal of the Institute of Polytechnics, Osaka City University Series D 7: 73–94.Google Scholar
Kozlowski, T. T. (ed.) 1984. Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Kozlowski, T. T. and S. G. Pallardy. 1984. Effect of flooding on water, carbohydrate, and mineral relations. pp. 165–193. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Kramer, P. J. 1983. Water Relations of Plants. Orlando: Academic Press.Google Scholar
Krebs, C. J. 1978. Ecology: The Experimental Analysis of Distribution and Abundance. New York: Harper and Row.Google Scholar
Krebs, C. J. 1989. Ecological Methodology. New York: Harper and Row.Google Scholar
Kruckeberg, A. R. 1954. The ecology of serpentine soils. III. Plant species in relation to serpentine soils. Ecology 35: 267–274.Google Scholar
Küchler, A. W. 1949. A physiognomic classification of vegetation. Annals of the Association of American Geographers 39: 201–210.CrossRefGoogle Scholar
Küchler, A. W. 1966. Analyzing the physiognomy and structure of vegetation. Annals of the Association of American Geographers 56: 112–127.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
Kuijt, J. 1969. The Biology of Parasitic Flowering Plants. Berkeley: University of California Press.Google Scholar
Kyte, F. T. 1998. A meteorite from the Cretaceous/Tertiary boundary. Nature 396: 237–239.CrossRefGoogle Scholar
Lack, D. 1947. Darwin's Finches: An Essay on the General Biological Theory of Evolution. New York: Harper and Row.Google Scholar
Laing, H. E. 1940. Respiration of the rhizomes of Nuphar advenum and other water plants. The American Journal of Botany 27: 574–581.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: 712–724.CrossRefGoogle Scholar
Langer, P. 1974. Stomach evolution in the artiodactyla. Mammalia 38: 295–314.CrossRefGoogle Scholar
Larcher, W. 1995. Physiological Plant Ecology: Ecophysiology and Stress Physiology of Functional Groups. 3rd edn. New York: Springer-Verlag.CrossRefGoogle Scholar
Larcher, W. 2003. Physiological Plant Ecology: Ecophysiology and Stress Physiology of Functional Groups. 4th edn. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Larcher, W. and H. Bauer. 1981. Ecological significance of resistance to low temperature. pp. 403–437. In Lange, O. L., Nobel, P. S., Osmond, C. B., and Ziegler, H. (eds.) Physiological Plant Ecology I: Responses to the Physical Environment. Encyclopedia of Plant Physiology: New Series, Vol. 12A. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Larson, D. W. 1980. Patterns of species distribution in an Umbilicaria-dominated community. Canadian Journal of Botany 58: 1269–1279.CrossRefGoogle Scholar
Larson, D. W. 1982. Environmental stress and Umbilicaria lichens: the effect of subzero temperature pretreatments. Oecologia 55: 268–278.CrossRefGoogle ScholarPubMed
Larson, D. W. 1989. The impact of ten years at − 20 °C on gas exchange in five lichen species. Oecologia 78: 87–92.CrossRefGoogle ScholarPubMed
Larson, D. W. 1996. Brown's Woods: an early gravel pit forest restoration project, Ontario, Canada. Society for Ecological Restoration 4: 11–18.CrossRefGoogle Scholar
Larson, D. W. 2001. The paradox of great longevity in a short-lived tree species. Experimental Gerontology 36: 651–673.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 Savanna River. Wetlands 14: 174–183.CrossRefGoogle Scholar
Latham, R. E. and Ricklefs, R. E.. 1993a. Global patterns of tree species richness in moist forests: energy-diversity theory does not account for variation in species richness. Oikos 67: 325–333.CrossRefGoogle Scholar
Latham, R. E. and R. E. Ricklefs. 1993b. Continental comparisons of temperate-zone tree species diversity. pp. 294–314. In Ricklefs, R. E. and Schluter, D. (eds.) Species Diversity in Ecological Communities: Historical and Geographical Perspectives. Chicago: The University of Chicago Press.Google Scholar
Latham, R. E., Beyea, J., Benner, M., Dunn, C. A., Fajvan, M. A., Freed, R. R., Grund, M., Horsley, S. B., Rhoads, A. F., and Shissler, B. P.. 2005. Managing White-tailed Deer in Forest Habitat from an Ecosystem Perspective: Pennsylvania Case Study. Harrisburg: Audubon Pennsylvania and Pennsylvania Habitat Alliance.Google Scholar
Lavoisier, A. L. 1789. Elements of Chemistry. Translated by R. Kerr and reprinted in xi, xii and pp. 1–60. In Adler, M. J. (ed.) 1990. Great Books of the Western World. 2nd edn., Vol. 42. Chicago: Encyclopaedia Britannica.Google Scholar
Lechowicz, M. J. 1981. The effects of climatic pattern on lichen productivity: Cetraria cucullata (Bell.) Ach. in the arctic tundra of northern Alaska. Oecologia 50: 210–216.CrossRefGoogle ScholarPubMed
Leck, M. A. and Graveline, K. J.. 1979. The seed bank of a freshwater tidal marsh. American Journal of Botany 66: 1006–1015.CrossRefGoogle Scholar
Leck, M. A., Parker, V. T., and Simpson, R. L. (eds.) 1989 Ecology of Soil Seed Banks. San Diego: Academic Press.Google Scholar
Lee, K. E. 1985. Earthworms: Their Ecology and Relationships with Soils and Land Use. Sydney: Academic Press.Google Scholar
Legendre, L. and Legendre, P.. 1983. Numerical Ecology. Amsterdam: Elsevier.Google Scholar
Leopold, A. 1949. A Sand County Almanac. London: Oxford University Press.Google Scholar
Page, C. and Keddy, P. A.. 1988. Reserves of buried seeds in beaver ponds. Wetlands 18: 242–248.CrossRefGoogle Scholar
Page, C. and Keddy, P. A.. 1998. Reserves of buried seeds in beaver ponds. Wetlands 18: 242–248.CrossRefGoogle Scholar
Levin, H. L. 1994. The Earth Through Time. 4th edn., updated. Fort Worth: Saunders College Publishing; Harcourt Brace College Publishers.Google Scholar
Levins, R. 1968. Evolution in Changing Environments. Princeton: Princeton University Press.Google Scholar
Levitt, J. 1977. The nature of stress injury and resistance. pp. 11–21. In Levitt, J. (ed.) Responses of Plants to Environmental Stress. New York: Academic Press.Google Scholar
Levitt, J. 1980. Responses of Plants to Environmental Stresses, Vols. I and II. 2nd edn. New York: Academic Press.Google Scholar
Lewis, D. H. 1987. Evolutionary aspects of mutualistic associations between fungi and photosynthetic organisms. pp. 161–178. In Rayner, A. D. M., Brasier, C. M., and Moore, D. (eds.) Evolutionary Biology of the Fungi. Symposium of the British Mycological Society, held at the University of Bristol, April 1986. Cambridge: Cambridge University Press.Google Scholar
Lewis, R. R. III (ed.) 1982. Creation and Restoration of Coastal Plant Communities. Boca Raton: CRC Press.Google Scholar
Leyser, O. and Fitter, A. 1998. Roots are branching out in patches. Trends in Plant Science 3: 203–204.CrossRefGoogle Scholar
Li, X.-L., George, E., and Marschner, H.. 1991. Phosphorus depletion and pH decrease at the root-soil and hyphae-soil interfaces of VA mycorrhizal white clover fertilized with ammonium. New Phytologist 119: 397–404.CrossRefGoogle Scholar
Lieth, H. 1975. Historical survey of primary productivity research. pp. 7–16. In Leith, H. and Whittaker, R. H. (eds.) Primary Productivity of the Biosphere. New York: Springer-Verlag.CrossRefGoogle Scholar
Likens, G. E., Bormann, F. H., Pierce, R. S., Eaton, J. S., and Johnson, N. M.. 1977. Biogeochemistry of a Forested Ecosystem. New York: Springer-Verlag.CrossRefGoogle Scholar
Little, C. E. 1995. The Dying of the Trees: The Pandemic in America's Forests. New York: Penguin Books.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: 1446–1455.CrossRefGoogle Scholar
Lloyd, D. G. and Barrett, S. C. H. (eds.) 1996. Floral Biology: Studies on Floral Evolution in Animal-Pollinated Plants. London: Chapman and Hall.CrossRefGoogle Scholar
Lodge, D. M. 1991. Herbivory on freshwater macrophytes. Aquatic Botany 41: 195–224.CrossRefGoogle Scholar
Loehle, C. 1998a. Height growth rate tradeoffs determine northern and southern range limits for trees. Journal of Biogeography 25: 735–742.CrossRefGoogle Scholar
Loehle, C. 1988b. Problems with the triangular model for representing plant strategies. Ecology 69: 284–286.CrossRefGoogle Scholar
Loehle, C. 1995. Anomalous responses of plants to CO2 enrichment. Oikos 73: 181–187.CrossRefGoogle Scholar
Lopoukhine, N. 1983. Parks Canada in the boreal forest ecosystem (a pilgrim's progress). pp. 167–179. In Wein, R. W., Riewe, R. R., and Methven, I. R. (eds.) Resources and Dynamics of the Boreal Zone. Proceedings of a Conference held at Thunder Bay, Ontario, August 1982. Ottawa: Association of Canadian Universities for Northern Studies.Google Scholar
Loreau, M. L. 2000. Biodiversity and ecosystem functioning: recent theoretical advances. Oikos 91: 3–17.CrossRefGoogle Scholar
Louda, S. M. and S. Mole. 1991. Glucosinolates: chemistry and ecology. pp. 124–164. In Rosenthal, G. A. and Berenbaum, M. R. (eds.) Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego: Academic Press.Google Scholar
Louda, S. M., K. H. Keller, and R. D. Holt. 1990. Herbivore influence on plant performance and competitive interactions. pp. 413–444. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competiton. San Diego: Academic Press.Google Scholar
Lovejoy, T. E., R. O. Bierregaard, Jr., A. B. Rylands, J. R. Malcolm, C. E. Quintela, L. H. Harper, K. S. Brown, Jr., A. H. Powell, G. V. N. Powell, H. O. R. Schubart, and M. B. Hays. 1986. Edge and other effects of isolation on Amazon forest fragments. pp. 257–285. In Soulé, M. E. (ed.) Conservation Biology; the Science of Scarcity and Diversity. Sunderland: Sinauer Associates.Google Scholar
Loveless, C. M. 1959. A study of the vegetation in the Florida Everglades. Ecology 40: 1–9.CrossRefGoogle Scholar
Lowman, M. D. 1992. Leaf growth dynamics and herbivory in five species of Australian rain-forest canopy trees. Journal of Ecology 80: 433–447.CrossRefGoogle Scholar
Lowman, M. D. and Rinker, H. B. (eds). 2004. Forest Canopies. 2nd edn. Burlington: Elsevier Academic Press.Google Scholar
Ludwig, D., Jones, D. D., and Holling, C. S.. 1978. Qualitative analysis of insect outbreak systems: the spruce budworm and forest. Journal of Animal Ecology 47: 315–332.CrossRefGoogle Scholar
Lugo, A. E. and Snedaker, S. C.. 1974. The ecology of mangroves. Annual Review of Ecology and Systematics 5: 39–64.CrossRefGoogle Scholar
Lutman, J. 1978. The role of slugs in an Agrostis–Festuca grassland. pp. 332–347. In Heal, O. W. and Perkins, D. F. (eds.) Production Ecology of British Moors and Montane Grasslands. Ecological Studies, Vol. 27. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Mabry, C. M. 2004. The number and size of seeds in common versus restricted woodland herbaceous species in central Iowa, USA. Oikos 107: 497–504.CrossRefGoogle Scholar
MacArthur, R. H. 1957. On the relative abundance of bird species. Proceedings of the National Academy of Sciences of the USA 43: 293–295.CrossRefGoogle ScholarPubMed
MacArthur, R. H. 1972. Geographical Ecology. New York: Harper and Row.Google Scholar
MacArthur, R. H. and Wilson, E. O.. 1967. The Theory of Island Biogeography. Monographs in Population Biology, No. 1. Princeton: Princeton University Press.Google Scholar
MacDonald, P. (ed.) 1989. The Solar System. The World of Science, Vol. 7. Oxford: Equinox (Oxford) Ltd.Google Scholar
MacFarland, C. G., Villa, J., and Toro, B.. 1974. The Galápagos giant tortoises (Geochelone elephantopus). Part I: Status of the surviving populations. Biological Conservation 6: 118–133.CrossRefGoogle Scholar
MacGillivray, C. W., Grime, J. P. and the Integrated Screening Programme (ISP) team. 1995. Testing predictions of the resistance and resilience of vegetation subjected to extreme events. Functional Ecology 9: 640–649.CrossRefGoogle Scholar
MacMahon, J. A. 1981. Successional processes: comparisons among biomes with special reference to probable roles of and influence on animals. pp. 277–305. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
Magnuson, J. J., H. A. Regier, W. J. Christie, and W. C. Sonzongi. 1980. To rehabilitate and restore Great Lakes ecosystems. pp. 95–122. In Cairns, J. Jr. (ed.) The Recovery Process in Damaged Ecosystmens. Ann Arbour: Ann Arbour Science Publishers.Google Scholar
Magnusson, M. (ed.) 1990. Chambers Biographical Dictionary. 5th edn. Edinburgh: W & R Chambers.Google Scholar
Mains, G. 1972. The Oxygen Revolution. London: David and Charles.Google Scholar
Major, J. 1988. Endemism: a botanical perspective. pp. 117–146. In Myers, A. A. and Giller, P. S. (eds.) Analytical Biogeography. London: Chapman and Hall.CrossRefGoogle Scholar
Margulis, L. 1970. Origin of Eukaryotic Cells. New Haven: Yale University Press.Google Scholar
Margulis, L. 1993. Symbiosis in Cell Evolution. 2nd edn. New York: W. H. Freeman.Google Scholar
Margulis, L. and Sagan, D.. 1986. Microcosmos: Four Billion Years of Evolution from Our Microbial Ancestors. Reprinted in 1997 in paperback. Berkeley: University of California Press.Google Scholar
Marquis, R. 1991. Evolution of resistance in plants to herbivores. Evolutionary Trends in Plants 5: 23–29.Google Scholar
Marquis, R. and Whelan, C.. 1994. Insectivorous birds increase growth of white oak through consumption of leaf-chewing insects. Ecology 75: 2007–2014.CrossRefGoogle Scholar
Marschner, H. 1995. Mineral Nutrition of Higher Plants. 2nd edn. London: Academic Press.Google Scholar
Martin, J. H.et al. 1994. Testing the iron hypothesis in ecosystems of the equatorial Pacific Ocean. Nature 371: 123–129.CrossRefGoogle Scholar
Martin, P. S. and Klein, R. J.. 1984. Quaternary Extinctions: A Prehistoric Revolution. Tucson: The University of Arizona Press.Google Scholar
Marx, K. 1867. In Engles, F. (ed.) Capital. Translated from 3rd German edition by S. Moore and E. Aveling. Revised from 4th edition by M. Sachey and H. Lamm. pp. 1–411. In Adler, M. J. (ed.) Great Books of the Western World. 2nd edn. 1990. Vol. 50. Chicago: Encyclopaedia Britannica.Google Scholar
Matson, P. A., Lohse, K., and Hall, S.. 2002. The globalization of nitrogen deposition: consequences for terrestrial ecosystems. Ambio 31: 113–119.CrossRefGoogle ScholarPubMed
Matthes-Sears, U., Gerrath, J., and Larson, D.. 1997. Abundance, biomass, and productivity of endolithic and epilithic lower plants on the temperate-zone cliffs of the Niagara Escarpment, Canada. International Journal of Plant Science 158: 451–460.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: 450–455.CrossRefGoogle Scholar
May, E. 1982. Budworm Battles: The Fight to Stop the Aerial Insecticide Spraying of the Forests of Eastern Canada. Halifax: Four East Publications Ltd.Google Scholar
May, R. M. 1973. Stability and Complexity in Model Ecosystems. Princeton: Princeton University Press.Google ScholarPubMed
May, R. M. 1977. Thresholds and breakpoints in ecosystems with a multiplicity of stable states. Nature 269: 471–477.CrossRefGoogle Scholar
May, R. M. 1981. Patterns in multi-species communities. pp. 197–227. In May, R. M. (ed.) Theoretical Ecology. Oxford: Blackwell.Google Scholar
May, R. M. 1988. How many species are there on Earth?Science 241: 1441–1449.CrossRefGoogle ScholarPubMed
Smith, Maynard J. 1978. The Evolution of Sex. Cambridge: Cambridge University Press.Google Scholar
Mayr, E. 1982. The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Cambridge: Belknap Press of Harvard University Press.Google 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: 373–381.CrossRefGoogle Scholar
McCarthy, K. A. 1987. Spatial and temporal distributions of species in two intermittent-tent ponds in Atlantic county, N. J. MSc thesis. New Brunswick: Rutgers University.
McClaran, M. P. 2003. A century of vegetation change on the Santa Rita Experimental Range. pp. 16–33. In USDA Forest Service Proceedings RMRS-P-30. US Department of Agriculture.
McClure, J. W. 1970. Secondary constituents of aquatic angiosperms. pp. 233–265. In Harborne, J. B. (ed.) Phytochemical Phylogeny. New York: Academic Press.Google Scholar
McComb, W. C. and Muller, R. N.. 1983. Snag management in old growth and second-growth Appalachian forests. Journal of Wildlife Management 47: 376–382.CrossRefGoogle Scholar
McGraw, J. B. 2001. Evidence for decline in stature of American ginseng plants from herbarium specimens. Biological Conservation 98: 25–32.CrossRefGoogle Scholar
McGraw, J. B. and Furedi, M. A.. 2005. Deer browsing and population viability of a forest understory plant. Science 307: 920–922.CrossRefGoogle ScholarPubMed
McIntosh, R. P. 1967. The continuum concept of vegetation. The Botanical Review 33: 130–187.CrossRefGoogle Scholar
McIntosh, R. P. 1981. Succession and ecological theory. pp. 10–23. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
McIntosh, R. P. 1985. The Background of Ecology: Concept and Theory. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
McKenzie, D. H., Hyatt, D. E., and McDonald, V. J.. 1992. Ecological Indicators, Vols. 1 and 2. London: Elsevier.Google Scholar
McNaughton, S. J. 1985. Ecology of a grazing ecosystem: the Serengeti. Ecological Monographs 55: 259–294.CrossRefGoogle Scholar
McNaughton, S. J., Ruess, R. W., and Seagle, S. W.. 1988. Large mammal and process dynamics in African ecosystems. Bioscience 38: 794–800.CrossRefGoogle Scholar
McNeill, J. R. and Winiwarter, V.. 2004. Breaking the sod: humankind, history and soil. Science 304: 1627–1629.CrossRefGoogle Scholar
McShea, W. J. and Rappole, J. H.. 2000. Managing the abundance and diversity of breeding bird populations through manipulation of deer populations. Conservation Biology 14: 1161–1170.CrossRefGoogle Scholar
McVaugh, R. 1943. The vegetation of the granitic flat-rocks of the southeastern United States. Ecological Monographs 13: 119–166.CrossRefGoogle Scholar
Meadows, D. H., Meadows, D. L., Randers, J., and Behrens, W. W. III. 1974. The Limits to Growth: A Report for the Club of Rome's Project on the Predicament of Mankind. 2nd edn. New York: The New American Library.Google Scholar
Meave, J. and Kellman, M.. 1994. Maintenance of rain forest diversity in riparian forest of tropical savannas: implications for species conservation during Pleistocene drought. Journal of Biogeography 21: 121–135.CrossRefGoogle Scholar
Merrens, E. J. and Peart, D. R.. 1992. Effects of hurricane damage on individual growth and stand structure in a hardwood forest in New Hampshire, USA. Journal of Ecology 80: 787–795.CrossRefGoogle Scholar
Milchunas, D. G. and Lauenroth, W. K.. 1993. Quantitative effects of grazing on vegetation and soils over a global range of environments. Ecological Monographs 63: 327–366.CrossRefGoogle Scholar
Milchunas, D. G., Laurenroth, W. K., and Burk, I. C.. 1998. Livestock grazing: animal and plant biodiversity of shortgrass steppe and the relationship to ecosystem function. Oikos 83: 65–74.CrossRefGoogle Scholar
Miller, G. R. and A. Watson. 1978. Heather productivity and its relevance to the regulation of red grouse populations. pp. 277–285. In Heal, O. W., Perkins, D. F. and Brown, W. M. (eds.) Production Ecology of British Moors and Montaine Grasslands. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Miller, G. R. and A. Watson. 1983. Heather moorland in northern Britain. pp. 101–117. In Warren, A. and Goldsmith, F. B. (eds.) Conservation in Perspective. Chichester: John Wiley and Sons Ltd.Google Scholar
Miller, K. G., Kominz, M. A., Browning, J. V., Wright, J. D., Mountain, G. S., Katz, M. E., Sugarman, P. J., Cramer, B. S., Christie-Blick, N., and Pekar, S. F.. 2005. The Phanerozoic record of global sea-level changes. Science 310: 1293–1298.CrossRefGoogle Scholar
Miller, R. S. 1967. Pattern and process in competition. Advances in Ecological Research 4: 1–74.CrossRefGoogle Scholar
Miller, S. L. 1953. A production of amino acids under possible primitive earth conditions. Science 117: 528–529.CrossRefGoogle ScholarPubMed
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
Milton, S. J., Dean, W. R. J., Plessis, M. A., and Siegfried, W. R.. 1994. A conceptual model of arid rangeland degradation. Bioscience 44: 70–76.CrossRefGoogle Scholar
Mitchley, J. 1988. Control of relative abundance of perennials in chalk grassland in southern England. II. Vertical canopy structure. Journal of Ecology 76: 341–350.CrossRefGoogle Scholar
Mitchley, J. and Grubb, P. J.. 1986. Control of relative abundance of perennials in chalk grassland in southern England. I. Constancy of rank order and results of pot- and field-experiments on the role of interference. Journal of Ecology 74: 1139–1166.CrossRefGoogle Scholar
Mitsch, W. J. and Gosselink, J. G.. 1986. Wetlands. New York: Van Nostrand Reinhold.Google Scholar
Mitsch, W. J. and Gosselink, J. G.. 2000. Wetlands. 3rd edn. New York: John Wiley & Sons.Google Scholar
Mitsch, W. J. and S. E. Jørgensen. 1990. Modelling and management. pp. 727–744. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 1. The Hague: SPB Academic Publishing.Google Scholar
Moffett, M. W. 1994. The High Frontier: Exploring the Tropical Rainforest Canopy. Cambridge: Harvard University Press.Google Scholar
Moles, A. T., Ackerly, D. D., Webb, C. O., Tweddle, J. C., Dickie, J. B., and Westoby, M.. 2005. A brief history of seed size. Science 307: 576–580.CrossRefGoogle ScholarPubMed
Molisch, H. 1937. Der Einfluss einer Pflanze auf die andere. Allelopathie. Jena: Gustav Fischer.Google Scholar
Moolman, H. J. and Cowling, R. M.. 1994. The impact of elephant and goat grazing on the endemic flora of South African succulent thicket. Biological Conservation 68: 53–61.CrossRefGoogle Scholar
Mooney, H. A. and Dunn, E. L.. 1970. Convergent evolution of mediterranean-climate evergreen sclerophyll shrubs. Evolution 24: 292–303.CrossRefGoogle ScholarPubMed
Mooney, H. A., Drake, B. G., Luxmoore, R. J., Oechel, W. C., and Pitelka, L. F.. 1991. Predicting ecosystem responses to elevated CO2 concentrations. Bioscience 41: 96–104.CrossRefGoogle Scholar
Moore, B. and Bolin, B.. 1987. The oceans, carbon dioxide and global climate change. Oceanus 29: 9–15.Google Scholar
Moore, D. R. J. 1990. Pattern and process in wetlands of varying standing crop: the importance of scale. PhD thesis. Ottawa: University of Ottawa.
Moore, D. R. J. 1998. The ecological component of ecological risk assessment: lessons from a field experiment. Human and Ecological Risk Assessment 4: 1103–1123.CrossRefGoogle Scholar
Moore, D. R. J. and Keddy, P. A.. 1989a. 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 P. A. Keddy. 1989b. Infertile wetlands: conservation priorities and management. pp. 391–397. In Bardecki, M. J. and Patterson, N. (eds.) Wetlands: Inertia or Momentum. Proceedings of a Conference held in Toronto, Ontario, October 21–22, 1988. Federation of Ontario Naturalists.Google 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: 2408–2412.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: 203–217.CrossRefGoogle Scholar
Moore, P. D., Webb, J. A., and Collinson, M. E.. 1991. Pollen Analysis. London: Blackwell Scientific.Google Scholar
Moreno, M. T., Cubero, J. I, Berner, D., Joel, D., Musselman, L. J., and Parker, C. (eds.) 1996. Advances in Parasitic Plant Research. Cordoba: Junta de Andalucia, Dirección General de Investigación Agraria.Google Scholar
Morgan, R. and Whitaker, B. 1986. Rainbow Warrior: The French Attempt to Sink Greenpeace. London: Arrow Books Ltd.Google Scholar
Morowitz, H. J. 1968. Energy Flow in Biology: Biological Organization as a Problem in Thermal Physics. New York: Academic Press.Google Scholar
Morris, E. C. and Myerscough, P. J.. 1991. Self-thinning and competition intensity over a gradient of nutrient availability. Journal of Ecology 79: 903–923.CrossRefGoogle Scholar
Moss, B. 1983. The Norfolk Broadland: experiments in the restoration of a complex wetland. Biological Reviews of the Cambridge Philosophical Society 58: 521–561.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: 115–128.CrossRefGoogle Scholar
Mueller-Dombois, D. 1987. Natural dieback in forests. Bioscience 37: 575–583.CrossRefGoogle Scholar
Mueller-Dombois, D. and Ellenberg, H.. 1974. Aims and Methods of Vegetation Ecology. New York: John Wiley and Sons.Google Scholar
Muller, C. H. 1966. The role of chemical inhibition (allelopathy) in vegetational composition. Bulletin of the Torrey Botanical Club 93: 332–351.CrossRefGoogle Scholar
Muller, C. H. 1969. Allelopathy as a factor in ecological process. Vegetatio 18: 348–357.CrossRefGoogle 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: 232–235.CrossRefGoogle Scholar
Muller, R. A. and MacDonald, G. J.. 1997. Glacial cycles and astronomical forcing. Science 277: 215–218.CrossRefGoogle Scholar
Musselman, L. J. and Mann, W. F. Jr. 1978. Root Parasites of Southern Forests. Southern Forest Experiment Station, Forest Service, U.S. Department of Agriculture.Google Scholar
Myers, C. W. and Donnelly, M. A.. 1997. A tepui herpetofauna on a granitic mountain (Tamacuari) in the borderland between Venezuela and Brazil: report from the Phipps Tapirapecó expedition. American Museum Novitates 3213: 1–71.Google Scholar
Myers, N. 1988. Threatened biotas: “hotspots” in tropical forests. Environmentalist 8: 1–20.CrossRefGoogle Scholar
Myers, N., Mittermeier, R. A., Mittermeier, C. G., Fonseca, G. A. B., and Kent, J.. 2000. Biodiversity hotspots for conservation priorities. Nature 403: 853–858.CrossRefGoogle ScholarPubMed
Myers, R., J. O'Brien, and S. Morrison. 2006. Fire Management Overview of the Caribbean Pine (Pinus caribaea) Savannas of the Mosquitia, Honduras. GFI Technical Report 2006-1b. Arlington: The Nature Conservancy.
Nabokov, P. 1993. Long threads. pp. 301–383. In Ballantine, B. and Ballantine, I. (eds.) The Native Americans: An Illustrated History. Atlanta: Turner Publishing Inc.Google Scholar
Naeem, S. 2002. Ecosystem consequences of biodiversity loss: the evolution of a paradigm. Ecology 83: 1537–1552.CrossRefGoogle Scholar
Naeem, S. L., Thompson, J., Lawler, S. P., Lawton, J. H., and Woodfin, R. M.. 1994. Declining biodiversity can alter the performance of ecosystems. Nature 368: 734–737.CrossRefGoogle Scholar
Naiman, R. J., Johnston, C. A., and Kelley, J. C.. 1988. Alteration of North American streams by beaver. Bioscience 38: 753–762.CrossRefGoogle Scholar
Nakamura, R. P. 1980. Plant kin selection. Evolutionary Theory 5: 113–117.Google 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: 229–251.CrossRefGoogle Scholar
Nantel, P. and Neuman, P.. 1992. Ecology of ectomycorrhizal-basidiomycete communities on a local vegetation gradient. Ecology 73: 99–117.CrossRefGoogle Scholar
Nantel, P., Gagnon, D., and Nault, A.. 1996. Population viability analysis of American ginseng and wild leek harvested in stochastic environments. Conservation Biology 10: 608–621.CrossRefGoogle Scholar
National Parks and Wildlife Service. 1998. Wollemi Pine Recovery Plan. Sydney: NPWS.
Newman, E. I. 1988. Mycorrhizal links between plants: their functioning and ecological significance. Advances in Ecological Research 18: 243–270.CrossRefGoogle Scholar
Newman, E. I. 1993. Applied Ecology. Oxford: Blackwell Scientific Publications.Google Scholar
Newman, S., Grace, J. B., and Koebel, J. W.. 1996. Effects of nutrients and hydroperiod on Typha, Cladium and Eleocharis: implications of Everglades restoration. Ecological Applications 6: 774–783.CrossRefGoogle Scholar
Nicholson, A. and Keddy, P. A.. 1983. The depth profile of a shoreline seed bank in Matchedash Lake, Ontario. Canadian Journal of Botany 61: 3293–3296.CrossRefGoogle Scholar
Nickrent, D. L. 2006. The parasitic plant connection. (http://www.parasiticplants.siu/index.htm) accessed 10 Nov. 2006.
Nicolson, M. and McIntosh, R. P.. 2002. H. A. Gleason and the individualistic hypothesis revisited. Bulletin of the Ecological Society of America (April 2002): 133–142.Google Scholar
Niering, W. A. and Warren, R. S.. 1980. Vegetation patterns and processes in New England salt marshes. Bioscience 30: 301–307.CrossRefGoogle Scholar
Niklas, K. J. 1994. Predicting the height of fossil plant remains: an allometric approach to an old problem. American Journal of Botany 81: 1235–1242.CrossRefGoogle Scholar
Niklas, K. J., Tiffney, B. H., and Knoll, A. H.. 1983. Patterns in vascular land plant diversification. Nature 303: 614–616.CrossRefGoogle Scholar
Niklas, K. J., B. H. Tiffney, and A. H. Knoll. 1985. Patterns in vascular plant diversification: an analysis at the species level. pp. 97–128. In Valentine, J. W. (ed.) Phanerozoic Diversity Pattern: Profiles in Macroevolution. Princeton: Princeton University Press.Google Scholar
Nilsson, L. A. 1992. Orchid pollination biology. Trends in Ecology and Evolution 7: 255–259.CrossRefGoogle Scholar
Nobel, P. S. 1976. Water relations and photosynthesis of a desert CAM plant, Agave deserti. Plant Physiology 58: 576–582.CrossRefGoogle ScholarPubMed
Nobel, P. S. 1977. Water relations and photosynthesis of a barrel cactus, Ferocactus acanthodes, in the Colorado Desert. Oecologia 27: 117–133.CrossRefGoogle ScholarPubMed
Nobel, P. S. 1985. Desert succulents. pp. 181–197. In Chabot, B. F. and Mooney, H. A. (eds.) Physiological Ecology of North American Plant Communities. London: Chapman and Hall.CrossRefGoogle Scholar
Noble, I. R. and Slatyer, R. O.. 1980. The use of vital attributes to predict successional changes in plant communities subject to recurrent disturbances. Vegetatio 43: 5–21.CrossRefGoogle Scholar
Noss, R. 1995. Maintaining Ecological Integrity in Representative Reserve Networks. A World Wildlife Fund Canada/United States Discussion Paper, WWF.
Noss, R. F. and Cooperrider, A. Y.. 1994. Saving Nature's Legacy. Washington, D.C.: Island Press.Google Scholar
Noy-Meir, I. 1973. Desert ecosystems: environment and producers. Annual Review of Ecology and Systematics 4: 25–51.CrossRefGoogle Scholar
Noy-Meir, L. 1975. Stability of grazing systems: an application of predator–prey graphs. Journal of Ecology 63: 459–481.CrossRefGoogle Scholar
Oakes, E. H. 2002. Notable Scientists. A to Z of Chemists. New York: Facts on File.Google Scholar
Ocampo, J. A. 1986. Vesicular-arbuscular mycorrhizal infection of “host” and “non-host” plants: effect on the growth responses of the plants and competition between them. Soil Biology and Biochemistry 18: 607–610.CrossRefGoogle Scholar
Ochoa, J. G., Molina, C., and Giner, S.. 1993. Inventario y estudio comunitario de los mamiferos del Parque National Canaima, con una lista de las especies registradas para la Guayana Venezolana. Ecologia Acta Cientifíca Venezolana 44: 245–262.Google Scholar
Okihana, H. and Ponnamperuma, C.. 1982. A protective function of the coacervates against UV light on the primitive Earth. Nature 299: 347–349.CrossRefGoogle Scholar
Oksanen, L. 1990. Predation, herbivory, and plant strategies along gradients of primary production. pp. 445–474. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Oksanen, L., Fretwell, S. D., Arruda, J., and Niemelä, P.. 1981. Exploitation ecosystems in gradients of primary productivity. The American Naturalist 118: 240–261.CrossRefGoogle Scholar
Oksanen, L., M. Aunapuu, T. Oksanen, M. Schneider, P. Ekerholm, P. A. Lundberg, T. Armulik, V. Aruoja, and L. Bondestad. 1997. Outlines of food webs in a low arctic tundra landscape in relation to three theories on trophic dynamics. pp. 351–373. In Gange, A. C. and Brown, V. K. (eds.) Multitrophic Interactions in Terrestrial Systems. The 36th Symposium of The British Ecological Society. Oxford: Blackwell Science.Google Scholar
Olson, D. M.et al. 2001. Terrestrial ecoregions of the world: a new map of life on Earth. Bioscience 51: 933–938.CrossRefGoogle Scholar
Ondok, J. P. 1990. Modelling ecological processes. pp. 659–89. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 1. The Hague: SPB Academic Publishing.Google Scholar
Oosting, H. J. 1956. The Study of Plant Communities, 2nd edn. San Francisco: W. H. Freeman.Google Scholar
Oparin, A. I. 1938. The Origin of Life. New York: The Macmillan Company. Translated by S. Morgulis.Google Scholar
Orloci, L. 1978. Multivariate Analysis in Vegetation Research. 2nd edn. The Hague: Junk.Google 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: 859–869.Google Scholar
Ostrofsky, M. L. and Zettler, E. R.. 1986. Chemical defenses in aquatic plants. Journal of Ecology 74: 279–287.CrossRefGoogle Scholar
Oxford Atlas of the World. 1997. New York: Oxford University Press.
Parkinson, C. L., Adams, K. L., and Palmer, J. D.. 1999. Multigene analyses identify the three earliest lineages of extant flowering plants. Current Biology 9: 1485–1491.CrossRefGoogle ScholarPubMed
Parks, J. C. and Werth, C. R.. 1993. A study of spatial features of clones in a population of bracken fern, Pteridium aquilinum (Dennstaedtiaceae). American Journal of Botany 80: 537–544.CrossRefGoogle Scholar
Pärtel, M., Zobel, M., Zobel, K., and Maarel, E.. 1996. The species pool and its relationship to species richness: evidence from Estonia plant communities. Oikos 75: 111–117.CrossRefGoogle Scholar
Pearce, C. M. 1993. Coping with forest fragmentation in southwestern Ontario. pp. 100–113. In Poser, S. F., Crins, W. J., and Beechey, T. J. (eds.) Size and Integrity Standards for Natural Heritage Areas in Ontario. Proceedings of a Seminar. Parks and Natural Heritage Policy Branch, Ontario Ministry of Natural Resources, Queen's Printer, Toronto.Google Scholar
Pearsall, W. H. 1920. The aquatic vegetation of the English Lakes. Journal of Ecology 8: 163–201.CrossRefGoogle Scholar
Peat, H. J. and Fitter, A. H.. 1993. The distribution of arbuscular mycorrhizas in the British flora. New Phytologist 125: 845–854.CrossRefGoogle Scholar
Peattie, D. C. 1922. The Atlantic coastal plain element in the flora of the Great Lakes. Rhodora 24: 50–70, 80–88.Google 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: 1536–1545.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. 1978. Forest vegetation of the Colorado Front Range: patterns of species diversity. Vegetatio 37: 65–78.CrossRefGoogle Scholar
Peet, R. K. 1992. Community structure and ecosystem function. pp. 103–151. In Glenn-Lewin, D. C., Peet, R. K., and Veblen, T. T. (eds.) Plant Succession: Theory and Prediction. Population and Community Biology, Vol. 11. London: Chapman and Hall.Google Scholar
Peet, R. K. and D. J Allard. 1993. Longleaf pine vegetation of the southern Atlantic and eastern Gulf coast regions: a preliminary classification. pp. 45–81. In Hermann, S. M. (ed.) Proceedings of the Tall Timbers Fire Ecology Conference. No. 18. The Longleaf Pine Ecosystem: Ecology, Restoration, and Management. Florida: Tall Timbers Research Station.Google Scholar
Pennings, S. C. and Callaway, R. M.. 1996. Impact of a parasitic plant on the structure and dynamics of salt marsh vegetation. Ecology 77: 1410–1419.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: 1968–1979.CrossRefGoogle Scholar
Percival, M. S. 1965. Floral Diversity. Oxford: Pergamon Press.Google Scholar
Perry, D. 1986. Life Above the Jungle Floor. San José: Don Perro Press.Google Scholar
Peters, R. H. 1992. A Critique for Ecology. Cambridge: Cambridge University Press.Google Scholar
Petit, J. R. et al. 1999. Climate and atmospheric history of the past 420,000 years from the Vostok ice core, Antarctica. Nature 399: 429–436.CrossRefGoogle Scholar
Petterson, B. 1965. Gotland and Öland: two limestone islands compared. Acta Phytogeographic Suecica 50: 131–140.Google Scholar
Phillips, D. L. and Shure, D. J.. 1990. Patch-size effects on early succession in southern Appalachian forest. Ecology 71: 204–212.CrossRefGoogle Scholar
Phipps, R. W. 1883. On the Necessity of Preserving and Replanting Forests. Toronto: Blackett and Robinson.CrossRefGoogle Scholar
Pianka, E. R. 1981. Competition and niche theory. pp. 167–196. In May, R. M. (ed.) Theoretical Ecology. Oxford: Blackwell.Google Scholar
Pianka, E. R. 1983. Evolutionary Ecology. 3rd edn. New York: Harper and Row.Google Scholar
Pickett, S. T. A. 1980. Non-equilibrium coexistence of plants. Bulletin of the Torrey Botanical Club 107: 238–248.CrossRefGoogle Scholar
Pickett, S. T. A. and White, P. S.. 1985. The Ecology of Natural Disturbance and Patch Dynamics. Orlando: Academic Press.Google Scholar
Pielou, E. C. 1975. Ecological Diversity. New York: John Wiley and Sons.Google Scholar
Pielou, E. C. 1977. Mathematical Ecology. New York: John Wiley and Sons.Google Scholar
Pielou, E. C. 1979. Biogeography. New York: John Wiley and Sons.Google Scholar
Pielou, E. C. 1984. The Interpretation of Ecological Data: A Primer on Classification and Ordination. New York: Wiley.Google Scholar
Pielou, E. C. and Routledge, R. D.. 1976. Salt marsh vegetation: latitudinal gradients in the zonation patterns. Oecologia 24: 311–321.CrossRefGoogle ScholarPubMed
Pietropaolo, J. and Pietropaolo, P.. 1986. Carnivorous Plants of the World. Portland: Timber Press.Google Scholar
Pimm, S. L. 2001. The World According to Pimm: A Scientist Audits the Earth. New York: McGraw-Hill.Google Scholar
Pirozynski, D. W. and D. W. Malloch. 1988. Seeds, spores and stomachs: coevolution in seed dispersal mutualisms. pp. 228–244. In Pirozynski, K. A. and Hawksworth, D. L. (eds.) Coevolution of Fungi with Plants and Animals. London: Academic Press.Google Scholar
Pirozynski, K. A. and Dalpé, Y.. 1989. Geological history of the Glomaceae with particular reference to mycorrhizal symbiosis. Symbiosis 7: 1–36.Google Scholar
Pitman, N. C. A. and Jørgensen, P. M.. 2002. Estimating the size of the world's threatened flora. Science 298: 989.CrossRefGoogle ScholarPubMed
Platt, W. J. 1999. Southeastern pine savannas. pp. 23–51. In Anderson, R. C., Fralish, J. S., and Baskin, J. M. (eds.) Savannas, Barrens and Rock Outcrop Communities of North America. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Poljakoff-Mayber, A. and Gale, J.. (eds.) 1975. Plants in Saline Environments. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Ponnamperuma, F. N. 1984. Effects of flooding on soils. pp. 9–45. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Poole, R. W. and Rathcke, B. J.. 1979. Regularity, randomness, and aggregation in flowering phenologies. Science 203: 470–471.CrossRefGoogle ScholarPubMed
Porter, C. L. 1967. Taxonomy of Flowering Plants. 2nd edn. San Francisco: W. H. Freeman.Google Scholar
Porter, H. 1993. Interspecific variation in the growth response of plants to an elevated ambient CO2 concentration. Vegetatio 104/105: 77–97.CrossRefGoogle Scholar
Postel, S. L., Daily, G. C., and Ehrlich, P. R.. 1996. Human appropriation of renewable fresh water. Science 271: 785–788.CrossRefGoogle Scholar
Pound, R. 1893. Symbiosis and mutualism. The American Naturalist 27: 509–520.CrossRefGoogle Scholar
Power, M. E. 1992. Top-down and bottom-up forces in food webs: do plants have primacy?Ecology 73: 733–746.CrossRefGoogle Scholar
Prance, G. T. and Johnson, D. M.. 1991. Plant collections from the plateau of Serra do Aracá (Amazonas, Brazil) and their phytogeographic affinities. Kew Bulletin 47: 1–24.CrossRefGoogle Scholar
Press, M. C. and Graves, J. D. (eds.) 1995. Parasitic Plants. London: Chapman and Hall.Google 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: 124–128.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: 410–432.CrossRefGoogle Scholar
Price, M. V. 1983. Ecological consequences of body size: a model for patch choice in desert rodents. Oecologia 59: 384–392.CrossRefGoogle ScholarPubMed
Price, M. V. 1984. Alternative paradigms in community ecology. pp. 354–383. In Price, P. W., Slobodchikoff, C. N., and Gaud, W. S. A. (eds.) A New Ecology: Novel Approaches to Interactive Systems. New York: John Wiley and Sons.Google Scholar
Price, M. V. and Reichman, O. J.. 1987. Distribution of seeds in Sonoran Desert soils: implications for heteromyid rodent foraging. Ecology 68: 1797–1811.CrossRefGoogle ScholarPubMed
Pringle, H. 1996. In Search of Ancient North America. New York: John Wiley and Sons.Google Scholar
Pritchard, P. C. H. 1996. The Galápagos Tortoises: Nomenclatural and Survival Status. Chelonian Research Monographs No. 1. Lunenbug: Chelonian Research Foundation.Google Scholar
Puckett, L. J. 1994. Nonpoint and Point Sources of Nitrogen in Major Watersheds of the United States. U.S. Geological Survey Water-Resources Investigations Report 94–4001.
Puerto, A., Rico, M., Matias, M. D., and García, J. A.. 1990. Variation in structure and diversity in Mediterranean grasslands related to trophic status and grazing intensity. Journal of Vegetation Science 1: 445–452.CrossRefGoogle Scholar
Pusey, A. and Wolf, M.. 1996. Inbreeding avoidance in animals. Trends in Ecology and Evolution 11: 201–206.CrossRefGoogle ScholarPubMed
Putwain, P. D. and Harper, J. L.. 1970. Studies in the dynamics of plant populations. III. The influence of associated species on populations of Rumex acetosa L. and R. acetosella L. in grassland. Journal of Ecology 58: 251–264.CrossRefGoogle Scholar
Putz, F. E. and Canham, C. D.. 1992. Mechanisms of arrested succession in shrublands: root and shoot competition between shrubs and tree seedlings. Forest Ecology and Management 49: 267–275.CrossRefGoogle Scholar
Quinn, C. J. and R. A. Price. 2003. Phylogeny of the southern hemisphere conifers. pp. 129–136. In R. R. Mill (ed.) Proceedings of the 4th International Conifer Conference. ISHS Acta Horticulturae 615.
Radford, A. E., Ahles, H. E. and Bell, C. R.. 1968. Manual of the Vascular Flora of the Carolinas. Chapel Hill: The University of North Carolina Press.Google Scholar
Rapport, D. J. 1989. What constitutes ecosystem health?Perspectives in Biology and Medicine 33: 120–132.CrossRefGoogle Scholar
Rapport, D. J., Thorpe, C., and Hutchinson, T. C.. 1985. Ecosystem behaviour under stress. The American Naturalist 125: 617–640.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. 1907. The life-forms of plants and their bearing on geography. Translated from Danish and republished in 1934. In The Life Forms of Plants and Statistical Plant Geography. pp. 2–104. Oxford: Clarendon Press.Google Scholar
Raunkiaer, C. 1908. The statistics of life-forms as a basis for biological plant geography. Translated from Danish and republished in 1934. In The Life Forms of Plants and Statistical Plant Geography. pp. 111–147. Oxford: Clarendon Press.Google Scholar
Raunkiaer, C. 1918. On the biological normal spectrum. Translated from German and republished in 1934 in The Life Forms of Plants and Statistical Plant Geography. pp. 425–434. Oxford: Clarendon Press.Google Scholar
Raunkiaer, C. 1934. The Life Forms of Plants and Statistical Plant Geography: Being the Collected Papers of Raunkiaer. Translated from the Danish, French and German. Preface by A. G. Tansley. Oxford: Clarendon Press.Google Scholar
Raven, P. H., Evert, R. F., and Eichhorn, S. E.. 1999. Biology of Plants. 6th edn. New York: W. H. Freeman and Company/Worth Publishers.Google Scholar
Raven, P. H., Evert, R. F. and Eichhorn, S. E.. 2005. Biology of Plants. 7th edn. New York: W. H. Freeman and Company Publishers.Google Scholar
Ravera, O. 1989. Lake ecosystem degradation and recovery studied by the enclosure method. pp. 217–243. In Ravera, O. (ed.) Ecological Assessment of Environmental Degradation, Pollution and Recovery. Amsterdam: Elsevier Science Publishers.Google Scholar
Rawes, M. and O. W. Heal. 1978. The blanket bog as part of a Pennine moorland. pp. 224–243. In Heal, O. W. and Perkins, D. F. (eds.) Production Ecology of British Moors and Montane Grasslands. Ecological Studies, Vol. 27. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Read, D. J., Koucheki, H. K., and Hodgson, J.. 1976. Vesicular-arbuscular mycorrhizae in natural vegetation systems. 1. The occurrence of infection. New Phytologist 77: 641–653.CrossRefGoogle Scholar
Reader, R. J. and Best, B. J.. 1989. Variation in competition along an environmental gradient: Hieracium floribundum in an abandoned pasture. Journal of Ecology 77: 673–684.CrossRefGoogle Scholar
Rees, W. E. and M. Wackernagel. 1994. Ecological footprints and appropriated carrying capacity: measuring the natural capital requirements of the human economy. pp. 362–390. In Jansson, A., Hammer, M., Folke, C., and Costanza, R. (eds.) Investing in Natural Capital: The Ecological Economics Approach to Sustainability. Washington, D.C.: Island Press.Google Scholar
Regal, P. J. 1977. Ecology and evolution of flowering plant dominance. Science 196: 622–629.CrossRefGoogle ScholarPubMed
Reid, D. M. and K. J. Bradford. 1984. Effect of flooding on hormone relations. pp. 195–219. In Kozlowski, T. T. (ed.) Flooding and Plant Growth. Orlando: Academic Press.Google Scholar
Reid, W. V., McNeely, J. A., Tunstall, D. B., Bryant, D. A., and Winograd, M.. 1993. Biodiversity Indicators for Policymakers. Washington, D.C.: World Resources Institute.Google Scholar
Richards, P. W. 1952. The Tropical Rain Forest: An Ecological Study. Paperback edition 1979. Cambridge: Cambridge University Press.Google Scholar
Richardson, C. J. (ed.) 1981. Pocosin Wetlands: An Integrated Analysis of Coastal Plain Freshwater Bogs in North Carolina. Stroudsburg: Hutchinson Ross Publishing Company.Google Scholar
Richardson, S. J., Peltzer, D. A., Allen, R. B., and McGlone, M. S.. 2005. Resorption proficiency along a chronosequence: responses among communities and within species. Ecology 80: 20–25.CrossRefGoogle Scholar
Rickerl, D. H., Sancho, F. O., and Ananth, S.. 1994. Vesicular-arbuscular endomycorrhizal colonization of wetland plants. Journal of Environmental Quality 23: 913–916.CrossRefGoogle Scholar
Rigler, F. H. 1982. Recognition of the possible: an advantage of empiricism in ecology. Canadian Journal of Fisheries and Aquatic Sciences 39: 1323–1331.CrossRefGoogle Scholar
Rigler, F. H. and Peters, R. H.. 1995. Science and Limnology. Oldendorf/Lutie: Ecology Institute.Google Scholar
Ritchie, J. C. 1987. Postglacial Vegetation of Canada. New York: Cambridge University Press.Google Scholar
Roberts, B. A. 1992. The serpentinized areas of Newfoundland, Canada: a brief review of their soils and vegetation. pp. 53–66. In Baker, A. J. M., Proctor, J., and Reeves, R. D. (eds.) The Vegetation of Ultramafic (Serpentine) Soils. Andover: Intercept.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: 117–127.CrossRefGoogle Scholar
Robinson, A. R. 1973. Sediment, our greatest pollutant? In Tank, R. W. (ed.) Focus on Environmental Geology. London: University Press.Google Scholar
Robinson, D. 1996. Resource capture by localised root proliferation: why do plants bother?Annals of Botany 77: 179–185.CrossRefGoogle Scholar
Robinson, J. M. 1990. Lignin, land plants, and fungi: biological evolution affecting phanerozoic oxygen balance. Geology 18: 607–610.2.3.CO;2>CrossRefGoogle Scholar
Rodman, J. E. 1974. Systematics and evolution of the genus Cakile (Cruciferae). Contributions from the Gray Herbarium 205: 3–146.Google Scholar
Rodrigues, A. S. L.et al. 2004. Effectiveness of the global protected areas network in representing species diversity. Nature 428: 640–643.CrossRefGoogle ScholarPubMed
Roger, A. J. 1999. Reconstructing early events in eukaryotic evolution. The American Naturalist 154 (Suppl.): S146–S163.CrossRefGoogle ScholarPubMed
Rohde, K. 1997. The larger area of tropics does not explain latitudinal gradients in species diversity. Oikos 79: 169–172.CrossRefGoogle Scholar
Rolston, H. 1994. Foreword. pp. xi–xiii. In L. Westra (ed.) An Environmental Proposal for Ethics: The Principle of Integrity. Lanham: Rowman and Littlefield. In R. Noss (ed.) 1995. Maintaining Ecological Integrity in Representative Reserve Networks. A World Wildlife Fund Canada/World Wildlife Fund/United States Discussion Paper, WWF.
Rose, R. 1924. Man and the Galapagos Islands. pp. 332–417. In Beebe, W. (ed.) Galapagos: World's End. New York: Putnam's Sons.Google Scholar
Rosenthal, G. A. and Berenbaum, M. R.. (eds.) 1991. Herbivores: Their Interactions with Secondary Plant Metabolites. San Diego: Academic Press.Google Scholar
Rosenzweig, M. L. 1995. Species Diversity in Space and Time. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Rosgen, D. L. 1995. River restoration utilizing natural stability concepts. pp. 55–62. In Kusler, J. A., Willard, D. E., and Hull, H. C. Jr. (eds.) Wetlands and Watershed Management: Science Applications and Public Policy. A collection of papers from a national symposium and several workshops at Tampa, Florida, April 23–26. New York: The Association of State Wetland Managers.Google Scholar
Rowe, J. S. and Scotter, G. W.. 1973. Fire in the boreal forest. Quaternary Research 3: 444–464.CrossRefGoogle Scholar
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: 564–569.CrossRefGoogle Scholar
Rundel, P. W., Cowling, R. M., Esler, K. J., Mustart, P. M., Jaarsveld, E., and Bezuidenhout, H.. 1995. Winter growth phenology and leaf orientation in Pachypodium namaquanum (Apocynaceae) in the succulent karoo of the Richtersveld, South Africa. Oecologia 101: 472–477.CrossRefGoogle ScholarPubMed
Russell, F. L., Zippin, D. B., and Fowler, N. L.. 2001. Effects of white-tailed deer (Odocoileus virginianus) on plants, plant populations and communities: a review. The American Midland Naturalist 146: 1–26.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: 233–240.CrossRefGoogle Scholar
Salisbury, E. J. 1942. The Reproductive Capacity of Plants. Studies in Quantitative Biology. London: G. Bell and Sons Ltd.Google Scholar
Salisbury, F. B. and Ross, C. W.. 1988. Plant Physiology. 3rd edn. Belmont: Wadsworth Publishers.Google Scholar
Salisbury, S. E. 1970. The pioneer vegetation of exposed muds and its biological features. Royal Society of London, Philosophical Transactions, Series B, 259: 207–255.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: 254–258.CrossRefGoogle Scholar
Salzman, A. G. and Parker, M. A.. 1985. Neighbours ameliorate local salinity stress for a rhizomatous plant in a heterogeneous environment. Oecologia 65: 273–277.CrossRefGoogle Scholar
Sandars, N. K. 1972. The Epic of Gilgamesh. An English version with an introduction by N. K. Sanders. Revised edition. London: Penguin Books.Google Scholar
Sarthou, C. and Villiers, J-F.. 1998. Epilithic plant communities on inselbergs in French Guiana. Journal of Vegetation Science 9: 847–860.CrossRefGoogle Scholar
Sather, J. H. and R. D. Smith. 1984. An Overview of Major Wetland Functions. U. S. Fish and Wildlife Service. FWS/OBS-84/18.
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: Surrey Beatty and Sons Pty Limited.Google Scholar
Savile, D. B. O. 1956. Known dispersal rates and migratory potentials as clues to the origin of the North American biota. The American Midland Naturalist 56: 434–453.CrossRefGoogle Scholar
Savile, D. B. O. 1972. Arctic Adaptations in Plants. Monograph No. 6. Ottawa: Canada Department of Agriculture.Google Scholar
Sayre, N. F. 2003. Recognizing history in range ecology: 100 years of science and management on the Santa Rita Experimental Range. pp. 1–15. In USDA Forest Service Proceedings RMRS-P-30. US Department of Agriculture.
Scagel, R. F., Rouse, G. E., Stein, J. R., Bandoni, R. J., Schofield, W. B., and Taylor, T. M. C.. 1965. An Evolutionary Survey of the Plant Kingdom. Belmont: Wadsworth Publishing Company.Google Scholar
Scagel, R. F., Bandoni, R. J., Rouse, G. E., Schofield, W. B., Stein, J. R., and Taylor, T. M. C.. 1969. Plant Diversity: An Evolutionary Approach. Belmont: Wadsworth Publishing Company.Google Scholar
Schaal, B. A. 1984. Life-history variation, natural selection, and maternal effects in plant populations. pp. 188–206. In Dirzo, R. and Sarukhán, J. (eds.) Perspectives on Plant Population Ecology. Sunderland: Sinauer Associates.Google Scholar
Schell, J. 1982. The Fate of the Earth. New York: Alfred A. Knopf.Google Scholar
Schnitzler, A. 1995. Successional status of trees in gallery forest along the river Rhine. Journal of Vegetation Science 6: 479–486.CrossRefGoogle Scholar
Scholander, P. F., Hammel, H. T., Bradstreet, B. D., and Hemmingsen, E. A.. 1965. Sap pressure in vascular plants. Science 148: 339–346.CrossRefGoogle ScholarPubMed
Schopf, J. W. and Barghoorn, E. S.. 1967. Alga-like fossils from the early Precambrian of South Africa. Science 156: 508–512.CrossRefGoogle Scholar
Schwinning, S. and Sala, O. E.. 2004. Hierarchy of responses to resource pulses in arid and semi-arid ecosystems. Oecologia 141: 211–220.CrossRefGoogle ScholarPubMed
Scott, J. M., Csuti, B., Jacobi, J. D., and Estes, J. E.. 1987. Species richness: a geographic approach to protecting future biological diversity. Bioscience 37: 782–788.CrossRefGoogle Scholar
Scott, M. G. and Larson, D. W.. 1985. The effect of winter field conditions on the distribution of two species of Umbilicaria. I. CO2 exchange in reciprocally transplanted thalli. New Phytologist 101: 89–101.CrossRefGoogle Scholar
Sculthorpe, C. D. 1967. The Biology of Aquatic Vascular Plants. Reprinted in 1985. London: Edward Arnold.Google Scholar
Seischab, F. K. and Orwig, D.. 1991. Catastrophic disturbances in the presettlement forests of western New York. Bulletin of the Torrey Botanical Club 118: 117–122.CrossRefGoogle 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: 677–687.CrossRefGoogle Scholar
Shaffer, G. P., J. G. Gosselink, and S. S. Hoeppner. 2005. The Mississippi River alluvial plain. pp. 272–315. In Fraser, L. H. and Keddy, P. A. (eds.) The World's Largest Wetlands: Ecology and Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Shannon, C. E. and Weaver, W.. 1949. The Mathematical Theory of Communication. Urbana: University of Illinois Press.Google Scholar
Sheail, J. and T. C. E. Wells. 1983. The fenlands of Huntingdonshire, England: a case study in catastrophic change. pp. 375–393. In Gore, A. J. P. (ed.) Ecosystems of the World 4B. Mires: Swamp, Bog, Fen and Moor. Amsterdam: Elsevier Scientific Publishing Company.Google Scholar
Shimwell, D. W. 1971. The Description and Classification of Vegetation. Seattle: University of Washington Press.Google Scholar
Shipley, B. 1993. A null model for competitive hierarchies in competition matrices. Ecology 74: 1693–1699.CrossRefGoogle Scholar
Shipley, B. and Dion, J.. 1992. The allometry of seed production in herbaceous angiosperms. The American Naturalist 139: 467–483.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: 340–345.CrossRefGoogle Scholar
Shipley, B. and Peters, R. H.. 1990. A test of the Tilman model of plant strategies: relative growth rate and biomass partioning. The American Naturalist 136: 139–153.CrossRefGoogle Scholar
Shipley, B., Keddy, P. A., and Lefkovitch, L. P.. 1991. Mechanisms producing plant zonation along a water depth gradient: a comparison with the exposure gradient. Canadian Journal of Botany 69: 1420–1424.CrossRefGoogle Scholar
Shrader-Frechette, K. S. and McCoy, E. D.. 1993. Method in Ecology: Strategies for Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Shugart, H. H., D. C. West, and W. R. Emanuel. 1981. Patterns and dynamics of forests: an application of simulation models. pp. 74–106. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession Concepts and Applications. New York: Springer-Verlag.CrossRefGoogle Scholar
Shure, D. J. 1999. Granite outcrops of the southeastern United States. pp. 99–118. In Anderson, R. C., Fralish, J. S., and Baskin, J. M. (eds.) Savannas, Barrens, and Rock Outcrop Plant Communities of North America. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Silliman, B. R. and Bertness, M. D.. 2002. A trophic cascade regulates salt marsh primary production. Proceedings of the National Academy of Sciences USA 99: 10500–10505.CrossRefGoogle ScholarPubMed
Silver, T. 1990. A New Face on the Countryside. Indians, Colonists and Slaves in the South Atlantic Forests, 1500–1800. Cambridge: Cambridge University Press.Google Scholar
Silvertown, J. 1980. The dynamics of a grassland ecosystem: botanical equilibrium in the Park Grass Experiment. Journal of Applied Ecology 17: 491–504.CrossRefGoogle Scholar
Silvertown, J. 1987. Introduction to Plant Population Ecology. 2nd edn. London: Longman.Google Scholar
Silvertown, J. and Charlesworth, D.. 2001. Introduction to Plant Population Biology. 4th edn. Oxford: Blackwell Science.Google Scholar
Silvertown, J., Dodd, M. E., McConway, K., Potts, J., and Crawley, M.. 1994. Rainfall, biomass variation, and community composition in the Park Grass Experiment. Ecology 75: 2430–2437.CrossRefGoogle Scholar
Silvertown, J., Poulton, P., Johnston, E., Edwards, G., Heard, M. and Biss, P. M.. 2006. The Park Grass experiment 1856–2006: its contribution to ecology. Journal of Ecology 94: 801–814.CrossRefGoogle Scholar
Simon, L., Bousquet, J., Lévesque, R. C., and Lalonde, M.. 1993. Origin and diversification of endomycorrhizal fungi and coincidence with vascular land plants. Nature 363: 67–69.CrossRefGoogle Scholar
Simpson, E. H. 1949. Measurement of diversity. Nature 163: 688.CrossRefGoogle Scholar
Sinclair, A. R. E. 1983. The adaptations of African ungulates and their effects on community function. pp. 401–425. In Bouliere, F. (ed.) Tropical Savannas. Amsterdam: 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: 987–994.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: 579–587.CrossRefGoogle Scholar
Sinclair, A. R. E., Krebs, C. J., Fryxell, J. M., Turkington, R., Boutin, S., Boonstra, R., Seccombe-Hett, P., Lundberg, P., and Oksanen, L.. 2000. Testing hypotheses of trophic level interactions: a boreal forest ecosystem. Oikos 89: 313–328.CrossRefGoogle Scholar
Sinclair, A. R. E. and C. J. Krebs. 2001. Trophic interactions, community organization, and the Kluane ecosystem. pp. 25–48. In Krebs, C. J., Boutin, S., and Boonstra, R. (eds.) Ecosystem Dynamics of the Boreal Forest. The Kluane Project. New York: Oxford University Press.Google Scholar
Sioli, H. 1964. General features of the limnology of Amazonia. Verhandlungen/Internationale Vereinigung fur theoretische und angewandte Limnologie 15: 1053–1058.Google Scholar
Sklar, F. H., R. Costanza, and J. W. Day Jr. 1990. Model conceptualization. pp. 625–658. In Patten, B. C. (ed.) Wetlands and Shallow Continental Water Bodies, Vol. 1. The Hague: SPB Academic Publishing.Google Scholar
Sklar, F. H.et al. 2005. The ecological–societal underpinnings of Everglades restoration. Frontiers in Ecology and the Environment 3: 161–169.Google Scholar
Slade, A. J. and Hutchings, M. J.. 1987. The effects of nutrient availability on foraging in the clonal herb Glechoma hederacea. Journal of Ecology 75: 95–112.CrossRefGoogle Scholar
Sletvold, N. 2002. Effects of plant size on reproductive output and offspring performance in the facultative biennial Digitalis purpurea. Journal of Ecology 90: 958–996.CrossRefGoogle Scholar
Small, E. 1972a. Water relations of plants in raised Sphagnum peat bogs. Ecology 53: 726–728.CrossRefGoogle Scholar
Small, E. 1972b. Photosynthetic rates in relation to nitrogen recycling as an adaptation to nutrient deficiency in peat bogs. Canadian Journal of Botany 50: 2227–2233.CrossRefGoogle Scholar
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: 487–495.CrossRefGoogle Scholar
Smith, A. 1776. An enquiry into the nature and causes of the wealth of nations. In Adler, M. J. (ed.) 1990. Great Books of the Western World, Vol. 36. Chicago: Encyclopaedia Britannica.Google Scholar
Smith, C. C. 1970. The coevolution of pine squirrels (Tamiasciurus) and conifers. Ecological Monographs 40: 349–371.CrossRefGoogle Scholar
Smith, C. C. and Follmer, D.. 1972. Food preference of squirrels. Ecology 53: 82–91.CrossRefGoogle Scholar
Smith, D. C. 1980. Mechanisms of nutrient movement between lichen symbionts. pp. 197–227. In Cook, C. B., Pappas, P. W., and Rudolph, E. D. (eds.) Cellular Interactions in Symbiosis and Parasitism. Columbus: Ohio State University Press.Google Scholar
Smith, D. C. and Douglas, A. E.. 1987. The Biology of Symbiosis. London: Edward Arnold.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: 673–684.CrossRefGoogle Scholar
Smith, L. M. and Kadlec, J. A.. 1985a. Fire and herbivory in a Great Salt Lake marsh. Ecology 66: 259–265.CrossRefGoogle Scholar
Smith, L. M. and Kadlec, J. A.. 1985b. Comparisons of prescribed burning and cutting of Utah marsh plants. Great Basin Naturalist 45: 463–466.Google Scholar
Smith, R. L. 1986. Elements of Ecology. New York: Harper and Row.Google Scholar
Smith, V. H. 1982. The nitrogen and phosphorus dependence of algal biomass in lakes: an empirical and theoretical analysis. Limnology and Oceanography 27: 1101–1112.CrossRefGoogle Scholar
Smith, V. H. 1983. Low nitrogen to phosphorus ratios favor dominance by blue-green algae in lake phytoplankton. Science 221: 669–671.CrossRefGoogle ScholarPubMed
Smol, J. P. and Cumming, B. F.. 2000. Tracking long-term changes in climate using algal indicators in lake sediments. Journal of Phycology 36: 986–1011.CrossRefGoogle Scholar
Sneath, P. H. A. and Sokal, R. R.. 1973. Numerical Taxonomy. San Francisco: W. H. Freeman.Google Scholar
Snell, T. W. and Burch, D. G.. 1975. The effects of density on resource partitioning in Chamaesyce hirta (Euphorbiaceae). Ecology 56: 742–746.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: 669–684.CrossRefGoogle Scholar
Sobel, D. 1995. Longitude: The True Story of a Lone Genius Who Solved the Greatest Scientific Problem of His Time. New York: Penguin Books.Google Scholar
Soper, J. H. and Maycock, P. F.. 1963. A community of arctic-alpine plants on the east shore of Lake Superior. Canadian Journal of Botany 41: 183–198.CrossRefGoogle Scholar
Sorrie, B. A. 1994. Coastal plain ponds in New England. Biological Conservation 68: 225–233.CrossRefGoogle Scholar
Soulé, M. and Noss, R.. 1998. Rewilding and biodiversity: complementary goals for continental conservation. Wild Earth 8(3): 18–28.Google Scholar
Sousa, W. P. 1984. The role of disturbance in natural communities. Annual Review of Ecology and Systematics 15: 353–391.CrossRefGoogle Scholar
Southwood, T. R. E. 1977. Habitat, the templet for ecological strategies?Journal of Animal Ecology 46: 337–365.CrossRefGoogle Scholar
Southwood, T. R. E. 1985. Interactions of plants and animals: patterns and processes. Oikos 44: 5–11.CrossRefGoogle Scholar
Southwood, T. R. E. 1988. Tactics, strategies, and templets. Oikos 52: 3–18.CrossRefGoogle Scholar
Specht, A. and Specht, R.. 1993. Species richness and canopy productivity of Australian plant communities. Biodiversity and Conservation 2: 152–167.CrossRefGoogle Scholar
Specht, R. L. and Specht, A.. 1989. Species richness of overstorey strata in Australian plant communities – the influence of overstorey growth rates. Australian Journal of Botany 37: 321–336.CrossRefGoogle 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 plant propagules. Wetlands 17: 55–64.CrossRefGoogle Scholar
Sporne, K. R. 1956. The phylogenetic classification of the angiosperms. Biological Reviews 31: 1–29.CrossRefGoogle Scholar
Sporne, K. R. 1970. The Morphology of Pteridophytes: The Structure of Ferns and Allied Plants. 3rd edn. London: Hutchinson and Co.Google Scholar
Sprengel, C. K. 1793. Discovery of the secret of nature in the structure and fertilization of flowers. Vieweq, Berlin. Translation of title and first chapter by P. Haase. pp. 3–43. In Lloyd, D. G. and Barrett, S. C. H. (eds.) Floral Biology. Studies on Floral Evolution in Animal-Pollinated Plants. London: Chapman and Hall.Google Scholar
Starfield, A. M. and Bleloch, A. L.. 1986. Building Models for Conservation and Wildlife Management. New York: Macmillan.Google 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
Stearn, W. T. 1979. Linnaean classification. pp. 96–101. In Black, D. (ed.) 1979. Carl Linnaeus: Travels. New York: Scribner's Sons.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
Steenbergh, W. F. and Lowe, C. H.. 1969. Critical factors during the first years of life of the saguaro (Cereus giganteus) at Saguaro National Monument, Arizona. Ecology 50: 825–834.CrossRefGoogle Scholar
Steila, D. 1993. Soils. pp. 47–54. In Flora of North America, Vol. 1. Introduction. New York: Oxford University Press.Google Scholar
Stein, B. A., Kutner, L. S., and Adams, J. S. (eds.) 2000. Precious Heritage. The Status of Biodiversity in the United States. Oxford: Oxford University Press.Google Scholar
Steneck, R. S. and Dethier, M. N.. 1994. A functional group approach to the structure of algal-dominated communities. Oikos 69: 476–498.CrossRefGoogle Scholar
Stephenson, N. L. 1990. Climatic control of vegetation distribution: the role of the water balance. The American Naturalist 135: 649–680.CrossRefGoogle Scholar
Stephenson, S. N. and Herendeen, P. S.. 1986. Short-term drought effects on the alvar communities of Drummond Island, Michigan. The Michigan Botanist 25: 16–27.Google Scholar
Stevenson, J. C., L. G. Ward, and M. S. Kearney. 1986. Vertical accretion in marshes with varying rates of sea level rise. pp. 241–259. In Wolfe, D. A. (ed.) Estuarine Variability. San Diego: Academic Press.Google Scholar
Stewart, W. N. and Rothwell, G. W.. 1993. Paleobotany and the Evolution of Plants. 2nd edn. Cambridge: Cambridge University Press.Google Scholar
Steyermark, J. A. 1982. Relationships of some Venezuelan forest refuges with lowland tropical floras. pp. 182–220. In Prance, G. T. (ed.) Biological Diversification in the Tropics. New York: Columbia University Press.Google Scholar
Stoltzenberg, D. 2004. Fritz Haber. Chemist, Nobel Laureate, German, Jew. Philadelphia: Chemical Heritage 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 yr B.P. Quaternary Research 12: 83–118.CrossRefGoogle Scholar
Strong, D. R. Jr., Simberloff, D., Abele, L. G., and Thistle, A. B. (eds). 1984. Ecological Communities. Conceptual Issues and the Evidence. Princeton: Princeton University Press.Google Scholar
Stupka, A. 1964. Trees, Shrubs and Woody Vines of Great Smoky National Park. Knoxville: The University of Tennessee Press.Google Scholar
Suffling, R., Lihou, C., and Morand, Y.. 1988. Control of landscape diversity by catastrophic disturbance: a theory and case study of fire in a Canadian boreal forest. Environmental Management 12: 73–78.CrossRefGoogle Scholar
Sun, G., Ji, Q., Dilcher, D. L., Zheng, S., Nixon, K. C., and Wang, X.. 2002. Archaefructaceae, a new basal angiosperm family. Science 296: 899–904.CrossRefGoogle ScholarPubMed
Sutter, R. D. and Kral, R.. 1994. The ecology, status, and conservation of two non-aluvial wetland communities in the south Atlantic and eastern Gulf coastal plain, USA. Biological Conservation 68: 235–243.CrossRefGoogle Scholar
Szarek, S. R. and I. P. Ting. 1975. Photosynthetic efficiency of CAM plants in relation to C3 and C4 plants. pp. 289–297. In Marcelle, R. (ed.) Environmental and Biological Control of Photosynthesis. The Hague: W. Junk.CrossRefGoogle Scholar
Tabachnick, B. G. and Fidell, L. S.. 2001. Using Multivariate Statistics. 4th edn. Boston: Allyn and Bacon.Google Scholar
Taiz, L. and Zeiger, E.. 1991. Plant Physiology. San Francisco: Benjamin-Cummings.Google Scholar
Takhtajan, A. 1969. Flowering Plants: Origin and Dispersal. Edinburgh: Oliver and Boyd. Translated and revised from a Russian second edition published in Moscow in 1961.Google Scholar
Takhtajan, A. 1986. Floristic Regions of the World. Berkeley: University of California Press. Translated by T. J. Crovello.Google Scholar
Tansley, A. G. 1939. The British Islands and their Vegetation. Cambridge: Cambridge University Press.Google Scholar
Tansley, A. 1987. What is ecology?Biological Journal of the Linnean Society 32: 5–16.CrossRefGoogle Scholar
Tansley, A. G. and Adamson, R. S.. 1925. Studies of the vegetation of the English chalk. Part III. The chalk grasslands of the Hampshire-Sussex border. Journal of Ecology XIII: 177–223.CrossRefGoogle Scholar
Tansley, A. G. and Chipp, T. F. (eds.) 1926. Aims and Methods in the Study of Vegetation. London: The British Empire Vegetation Committee and Crown Agents for Colonies.Google 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 template for life history strategies. Oikos 58: 239–250.CrossRefGoogle Scholar
Taylor, T. N., Kerp, H., and Hass, H.. 2005. Life history biology of early land plants: deciphering the gametophyte phase. Proceedings of the National Academy of Sciences 102: 5892–5897.CrossRefGoogle ScholarPubMed
Taylor, J. G. 1984. Louisiana: A History. New York: W. W. Norton & Company.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, T. N. 1988. The origin of land plants: some answers, more questions. Taxon 37: 805–833.CrossRefGoogle Scholar
Taylor, T. N. 1990. Fungal associations in the terrestrial paleoecosystem. Trends in Ecology and Evolution 5: 21–25.CrossRefGoogle Scholar
Temple, S. A. 1977. Plant–animal mutualism: coevolution with dodo leads to near extinction of plant. Science 197: 886–887.CrossRefGoogle ScholarPubMed
Terborgh, J. 1989. Where Have All the Birds Gone?Princeton: Princeton University Press.Google Scholar
Terborgh, J.et al. 2001. Ecological meltdown in predator-free forest fragments. Science 294: 1923–1926.CrossRefGoogle ScholarPubMed
Thirgood, J. V. 1981. Man and the Mediterranean Forest: A History of Resource Depletion. London: Academic Press.Google Scholar
Thompson, D. J. and Shay, J. M.. 1988. First-year response of a Phragmites marsh community to seasonal burning. Canadian Journal of Botany 67: 1448–1455.CrossRefGoogle Scholar
Thorne, R. F. 1963. Some problems and guiding principles of angiosperm phylogeny. The American Naturalist 97: 287–305.CrossRefGoogle Scholar
Tilghman, N. G. 1989. Impacts of white-tailed deer on forest regeneration in northwestern Pennsylvania. Journal of Wildlife Management 53: 524–532.CrossRefGoogle Scholar
Tilman, D. 1982. Resource Competition and Community Structure. Princeton: Princeton University Press.Google ScholarPubMed
Tilman, D. and S. Pacala. 1993. The maintenance of species richness in plant communities. pp. 13–25. In Ricklefs, R. E. and Schluter, D. (eds.) Species Diversity in Ecological Communities, Chicago: University of Chicago Press.Google Scholar
Tilman, D., Wedin, D., and Knops, J.. 1996. Productivity and sustainability influenced by biodiversity in grassland ecosystems. Nature 379: 718–720.CrossRefGoogle Scholar
Tinker, P. B., M. D. Jones, and D. M. Durall. 1992. A functional comparison of ecto- and endomycorrhizas. pp. 303–310. In Read, D. J., Lewis, D. H., Fitter, A. H., and Alexander, I. J. (eds.) Mycorrhizas in Ecosystems. Wallingford: CAB International.Google Scholar
Tomlinson, P. B. 1986. The Botany of Mangroves. Cambridge: Cambridge University Press.Google Scholar
Tschudy, R. H., Pillmore, C. L., Orth, C. J., Gilmore, J. S., and Knight, J. D.. 1984. Disruption of the terrestrial plant ecosystem at the Cretaceous-Tertiary boundary, Western Interior. Science 225: 1030–1032.CrossRefGoogle ScholarPubMed
Tubbs, C. H., R. M. DeGraff, M. Yamasaki, and W. M. Healy. 1987. Guide to wildlife tree management in New England northern hardwoods. United States Department of Agriculture and Forestry Service General Technical Report NE-118.
Turner, F. 1994. Beyond Geography. The Western Spirit Against the Wilderness. Fifth printing, first edition in 1983. New Brunswick: Rutgers University Press.Google Scholar
Turner, R. E. 1977. Intertidal vegetation and commercial yields of penaeid shrimp. Transactions of the American Fisheries Society 106: 411–416.2.0.CO;2>CrossRefGoogle Scholar
Turner, R. M. 1990. Long-term vegetation change at a fully protected Sonoran Desert site. Ecology 71: 464–477.CrossRefGoogle Scholar
Turner, R. M., Alcorn, S. M., Olin, G., and Booth, J. A.. 1966. The influence of shade, soil, and water on saguaro seedling establishment. Botanical Gazette 127: 95–102.CrossRefGoogle Scholar
Twolan-Strutt, L. and Keddy, P.. 1996. Above- and belowground competition intensity in two contrasting wetland plant communities. Ecology 77: 259–270.CrossRefGoogle Scholar
Udvardy, M. D. F. 1975. A Classification of the Biogeographical Provinces of the World. IUCN Occasional Paper No. 18. Morges: International Union for the Conservation of Nature and Natural Resources.Google Scholar
Uhl, C. and Kauffman, J. B.. 1990. Deforestation, fire susceptibility and potential responses to fire in the eastern Amazon. Ecology 71: 437–449.CrossRefGoogle Scholar
Underwood, A. J. 1978. The detection of non-random patterns of distribution of species along a gradient. Oecologia 36: 317–326.CrossRefGoogle ScholarPubMed
Urban, D. L. and H. H. Shugart. 1992. Individual based models of forest succession. pp. 249–292. In Glenn-Lewin, D. C., Peet, R. K., and Veblen, T. T.. (eds.) Plant Succession. London: Chapman and Hall.Google Scholar
U.S. Army Coastal Engineering Research Centre. 1977. Shore Protection Manual, Vol. 1, 3rd edn. Washington, D.C.: US Government Printing Office.
U.S. Army Corps of Engineers. 2004. The Mississippi River and Tributaries Project. New Orleans District Office Website (www.mvn.usace.army.mil/pao/bro/misstrib.htm) accessed 26 Mar. 2006.
U.S.D.A. 1975. Soil Taxonomy: A Basic System of Soil Classification for Making and Interpreting Soil Surveys. Agricultural Handbook 436, Washington, D.C.: U.S.D.A.
U.S.D.A. 2004. Emerald ash borer; the green menace. Animal and Plant Health Inspection Service. Program Aid No. 1769.
Vallentyne, J. R. 1974. The Algal Bowl. Lakes and Man. Miscellaneous Special Publication 22. Ottawa: Department of the Environment, Fisheries and Marine Service.Google Scholar
Breemen, N. 1995. How Sphagnum bogs down other plants. Trends in Ecology and Evolution 10: 270–275.CrossRefGoogle ScholarPubMed
Vandermeer, J. B., B. A. Hazlett, and B. Rathcke. 1985. Indirect facilitation and mutualism. pp. 326–343. In Boucher, D. H. (ed.) The Ecology of Mutualism. New York: Oxford University Press.Google Scholar
Valk, A. G. 1981. Succession in wetlands: a Gleasonian approach. Ecology 62: 688–696.CrossRefGoogle Scholar
Valk, A. G. and Davis, C. B.. 1976. The seed banks of prairie glacial marshes. Canadian Journal of Botany 54: 1832–1838.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: 322–335.CrossRefGoogle Scholar
Werf, A., Welschen, A., Welschen, R., and Lambers, H.. 1988. Respiratory energy costs for the maintenance of biomass, for growth and for iron uptake in roots of Carex diandra and Carex acutiformis. Physiologia Plantarum 72: 483–491.CrossRefGoogle Scholar
Venable, D. L. and C. E. Pake. 1999. Population ecology of Sonoran Desert annual plants. pp. 115–142. In Robichaux, R. H. (ed.) The Ecology of Sonoran Desert Plants and Plant Communities. Tucson: University of Arizona Press.Google Scholar
Veneklaas, E. J., Fajardo, A., Obregon, S., and Lozano, J.. 2005. Gallery forest types and their environmental correlates in a Colombian savanna landscape. Ecography 28: 236–252.CrossRefGoogle Scholar
Verhoeven, J. T. A. and Liefveld, W. M.. 1997. The ecological significance of organochemical compounds in Sphagnum. Acta Botanica Neerlandica 46: 117–130.CrossRefGoogle Scholar
Verhoeven, J. T. A. and Schmitz, M. B.. 1991. Control of plant growth by nitrogen and phosphorus in mesotrophic fens. Biogeochemistry 12: 135–148.CrossRefGoogle Scholar
Verhoeven, J. T. A., R. H. Kemmers, and W. Koerselman. 1993. Nutrient enrichment of freshwater wetlands. pp. 33–59. In Vos, C. C. and Opdam, P.. Landscape Ecology of a Stressed Environment. 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: 494–497.CrossRefGoogle ScholarPubMed
Vernadsky, V. 1929. La Biosphère. Paris: Felix Alcan.Google Scholar
Vernadsky, V. I. 1998. The Biosphere. New York: Copernicus, Springer-Verlag. Translated from the French and Russian, including a new foreword, introduction and appendices.CrossRefGoogle Scholar
Vesey-FitzGerald, D. F. 1960. Grazing succession among East African game animals. Journal of Mammalogy 41: 161–172.CrossRefGoogle 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: 651–667.CrossRefGoogle Scholar
Vitousek, P. M. 1982. Nutrient cycling and nitrogen use efficiency. The American Naturalist 119: 553–572.CrossRefGoogle Scholar
Vitousek, P., Ehrlich, P. R., Ehrlich, A. H., and Matson, P.. 1986. Human appropriation of the products of photosynthesis. Bioscience 36: 368–373.CrossRefGoogle Scholar
Vitousek, P. M.et al. 1997. Human alteration of the global nitrogen cycle: causes and consequences. Ecological Applications 7: 737–750.Google Scholar
Vitt, D. H. and Chee, W.. 1990. The relationship of vegetation to surface water chemistry and peat chemistry in fens of Alberta, Canada. Vegetatio 89: 87–106.CrossRefGoogle Scholar
Vivian-Smith, G. 1997. Microtopographic heterogeneity and floristic diversity in experimental wetland communities. Journal of Ecology 85: 71–82.CrossRefGoogle Scholar
Vogel, S. 1996. Christian Konrad Sprengel's theory of the flower: the cradle of floral ecology. pp. 44–62. In Lloyd, D. G. and Barrett, S. C. H. (eds.) Floral Biology. Studies on Floral Evolution in Animal-Pollinated Plants. London: Chapman and Hall.Google Scholar
Vogl, R. 1969. One hundred and thirty years of plant succession in a southeastern Wisconsin lowland. Ecology 50: 248–255.CrossRefGoogle Scholar
Humboldt, A. 1845. Cosmos: A Sketch of the Physical Description of the Universe, Vol. 1. Translated by E. C. Otté. Foundations of Natural History. Baltimore: Johns Hopkins University Press. 1997. (Originally produced in five volumes: 1845, 1847, 1850–51, 1858, and 1862.)Google Scholar
Waggoner, P. E. and Stephens, G. R.. 1970. Transition probabilities for a forest. Nature 225: 1160–1161.CrossRefGoogle ScholarPubMed
Walker, D. 1970. Direction and rate in some British post-glacial hydroseres. pp. 117–139. In Walker, D. and West, R. G. (eds.) Studies in the Vegetational History of the British Isles. Cambridge: Cambridge University Press.Google Scholar
Walker, J., C. H. Thompson, I. F. Fergus, and B. R. Tunstall. 1981. Plant succession and soil development in coastal sand dunes of subtropical eastern Australia. pp. 107–131. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession. Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
Wallin, I. E. 1927. Symbioticism and the Origin of Species. Baltimore: Williams and Wilkins.Google Scholar
Wardle, D. A. 1995. Impact of disturbance on detritus food-webs in agro-ecosystems of contrasting tillage and weed management practices. Advances in Ecological Research 26: 105–185.CrossRefGoogle Scholar
Wardle, D. A. 2002. Communities and Ecosystems: Linking the Aboveground and Belowground Components. Princeton: Princeton University Press.Google Scholar
Wardle, D. A., Huston, M. A., Grime, J. P., Berendse, F., Garnier, E., Laurenroth, W. K., Setala, H., and Wilson, S. D.. 2000. Biodiversity and ecosystem function: an issue in ecology. Bulletin of the Ecological Society of America 81: 235–239.Google Scholar
Wardle, D. A., Bardgett, R. D., Klironomos, J. N., Setälä, H., Putten, W. H., and Wall, D. H.. 2004. Ecological linkages between aboveground and belowground biota. Science 304: 1629–1633.CrossRefGoogle ScholarPubMed
Watkinson, A. R. 1985a. Plant responses to crowding. pp. 275–289. In White, J. (ed.) Studies in Plant Demography: A Festschrift for John L. Harper. London: Academic Press.Google Scholar
Watkinson, A. R. 1985b. On the abundance of plants along an environmental gradient. Journal of Ecology 73: 569–578.CrossRefGoogle Scholar
Watkinson, A. R. and Freckleton, R. P.. 1997. Quantifying the impact of arbuscular mycorrhizae on plant competition. Journal of Ecology 85: 541–545.CrossRefGoogle Scholar
Watkinson, A. R. and C. C. Gibson. 1988. Plant parasitism: the population dynamics of parasitic plants and their effects upon plant community structure. pp. 393–411. In Davy, A. J., Hutchings, M. J., and Watkinson, A. R.. Plant Population Biology. Oxford: Blackwell Scientific Publications.Google Scholar
Watt, A. S. 1919. On the causes of failure of natural regeneration in British oakwoods. Journal of Ecology 7: 173–203.CrossRefGoogle Scholar
Watt, A. S. 1923. On the ecology of British beechwoods with special reference to their regeneration. Part I. The causes of failure of natural regeneration of the beech. Journal of Ecology 11: 1–48.CrossRefGoogle Scholar
Weaver, J. E. and Clements, F. E.. 1929. Plant Ecology. New York: McGraw-Hill.Google Scholar
Weaver, J. E. and Clements, F. E.. 1938. Plant Ecology. 2nd edn. New York: McGraw-Hill Book Company.Google Scholar
Weber, W. and Rabinowitz, A.. 1996. A global perspective on large carnivore conservation. Conservation Biology 10: 1046–1054.CrossRefGoogle Scholar
Weetman, G. F. 1983. Forestry practices and stress on Canadian forest land. pp. 260–301. In Simpson-Lewis, W., McKechnie, R., and Neimanis, V. (eds.) Stress on Land in Canada. Ottawa: Lands Directorate, Environment Canada.Google Scholar
Weiher, E. and Keddy, P. A.. 1995a. The assembly of experimental wetland plant communities. Oikos 73: 323–335.CrossRefGoogle Scholar
Weiher, E. and Keddy, P. A.. 1995b. Assembly rules, null models, and trait dispersion: new questions from old patterns. Oikos 74: 159–165.CrossRefGoogle Scholar
Weiher, E. and Keddy, P. A. (eds.). 1999. Ecological Assembly Rules: Perspectives, Advances, Retreats. Cambridge: Cambridge University Press.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: 609–620.CrossRefGoogle Scholar
Wein, R. W. 1983. Fire behaviour and ecological effects in organic terrain. pp. 81–95. In Wein, R. W. and MacLean, D. A. (eds.) The Role of Fire in Northern Circumpolar Ecosystems. New York: John Wiley and Sons Ltd.Google Scholar
Wein, R. W. and Moore, J. M.. 1977. Fire history and rotations in the New Brunswick Acadian forest. Canadian Journal of Forest Research 7: 285–294.CrossRefGoogle Scholar
Weiner, J. 1985. Size hierarchies in experimental populations of annual plants. Ecology 66: 743–752.CrossRefGoogle Scholar
Weiner, J. 1986. How competition for light and nutrients affects size variablility in Ipomea tricolor populations. Ecology 67: 1425–1427.CrossRefGoogle Scholar
Weiner, J. and Thomas, S. C.. 1986. Size variability and competition in plant monocultures. Oikos 47: 221–222.CrossRefGoogle Scholar
Weisner, S. E. B. 1990. Emergent Vegetation in Eutrophic Lakes: Distributional Patterns and Ecophysiological Constraints. Sweden: Grahns Boktryckeri.
Welcomme, R. L. 1976. Some general and theoretical considerations on the fish yield of African rivers. Journal of Fish Biology 8: 351–364.CrossRefGoogle Scholar
Welcomme, R. L. 1979. Fisheries Ecology of Floodplain Rivers. London: Longman.Google Scholar
Weldon, C. W. and Slauson, W. L.. 1986. The intensity of competition versus its importance: an overlooked distinction and some implications. The Quarterly Review of Biology 61: 23–44.CrossRefGoogle Scholar
Weller, D. E. 1990. Will the real self-thinning rule please stand up? A reply to Osawa and Sugita. Ecology 71: 1204–1207.CrossRefGoogle Scholar
Weller, M. W. 1978. Management of freshwater marshes for wildlife. pp. 267–284. In Good, R. E., Whigham, D. F., and Simpson, R. L. (eds.) Freshwater Wetlands: Ecological Processes and Management Potential. New York: Academic Press.Google Scholar
Weller, M. W. 1994. Freshwater Marshes: Ecology and Wildlife Management. 3rd edn. Minneapolis: University of Minnesota.Google Scholar
Wells, H. G. 1956. The Outline of History: Being a Plain History of Life and Mankind. Garden City, NY: Garden City Books. Revised and brought up to the end of the Second World War by Raymond Postgate.Google Scholar
West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) 1981. Forest Succession: Concepts and Application. New York: Springer-Verlag.CrossRefGoogle Scholar
Westhoff, V. and E. van der Maarel. 1973. The Braun–Blanquet approach. pp. 617–707. In Whittaker, R. H. (ed.) Ordination and Classification of Communities. The Hague: Junk.CrossRefGoogle Scholar
Westoby, M. 1984. The self-thinning rule. Advances in Ecological Research 14: 167–225.CrossRefGoogle Scholar
Westoby, M. 1998. Leaf–height–seed (LHS) plant ecology strategy scheme. Plant and Soil 199: 213–227.CrossRefGoogle Scholar
Westoby, M., M. Leishman, and J. Lord. 1997. Comparative ecology of seed size and dispersal. pp. 143–162. In Silvertown, J., Franco, M., and Harper, J. L. (eds.) Plant Life Histories: Ecology, Phylogeny and Evolution. Cambridge: Cambridge University Press.Google 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
Whelan, P. M. and Hamann, O.. 1989. Vegetation regrowth on Isla Pinta: a success story. Noticias de Galápagos 48: 11–13.Google Scholar
Whelan, R. J. 1995. The Ecology of Fire. Cambridge: Cambridge University Press.Google Scholar
Whisenant, S. G. 1999. Repairing Damaged Wildlands. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
White, I. D., Mottershead, D. N., and Harrison, S. J.. 1992. Environmental Systems: An Introductory Text. 2nd edn. London: Chapman and Hall.Google Scholar
White, P. S. 1979. Pattern, process and natural disturbance in vegetation. The Botanical Review 45: 229–299.CrossRefGoogle Scholar
White, P. S. 1994. Synthesis: vegetation pattern and process in the Everglades ecosystem. pp. 445–460. In Davis, S. and Ogden, J. (eds.) Everglades: The Ecosystem and its Restoration. Delray Beach: St. Lucie Press.Google Scholar
White, P. S., S. P. Wilds, and G. A. Thunhorst. 1998. Southeast. pp. 255–314. In Mac, M. J., Opler, P. A., Haecker, C. E. Puckett, and Doran, P. D.. (eds.) Status and Trends of the Nation's Biological Resources, 2 Vols. Reston: U.S. Department of the Interior, U.S. Geological Survey.Google Scholar
White, T. C. R. 1993. The Inadequate Environment: Nitrogen and the Abundance of Animals. Berlin: Springer-Verlag.CrossRefGoogle Scholar
Whittaker, R. H. 1952. A study of summer foliage insect communities in the Great Smoky Mountains. Ecological Monographs 22: 1–44.CrossRefGoogle Scholar
Whittaker, R. H. 1954a. The ecology of serpentine soils. I. Introduction. Ecology 35: 258–259.CrossRefGoogle Scholar
Whittaker, R. H. 1954b. The ecology of serpentine soils. IV. The vegetational response to serpentine soils. Ecology 35: 275–288.Google Scholar
Whittaker, R. H. 1956. Vegetation of the Great Smoky Mountains. Ecological Monographs 26: 1–79.CrossRefGoogle Scholar
Whittaker, R. H. 1960. Vegetation of the Siskiyou Mountains, Oregon and California. Ecological Monographs 30: 279–338.CrossRefGoogle Scholar
Whittaker, R. H. 1962. Classification of natural communities. The Botanical Review 28: 1–239.CrossRefGoogle Scholar
Whittaker, R. H. 1965. Dominance and diversity in land plant communities. Science 147: 250–260.CrossRefGoogle ScholarPubMed
Whittaker, R. H. 1967. Gradient analysis of vegetation. Biological Reviews 42: 207–264.CrossRefGoogle ScholarPubMed
Whittaker, R. H. 1972. Evolution and measurement of species diversity. Taxon 21: 213–251.CrossRefGoogle Scholar
Whittaker, R. H. 1973a. Direct gradient analysis: techniques. pp. 9–31. In Whittaker, R. H. (ed.) Ordination and Classification of Communities. Part V. The Hague: W. Junk.CrossRefGoogle Scholar
Whittaker, R. H. (ed.) 1973b. Ordination and Classification of Communities. Part V. The Hague: W. Junk.CrossRefGoogle Scholar
Whittaker, R. H. 1975. Communities and Ecosystems. 2nd edn. London: Macmillan.Google Scholar
Wiens, J. A. 1977. On competition and variable environments. American Scientist 65: 590–597.Google Scholar
Wilcove, D. S., C. H. McLellan, and A. P. Dobson. 1986. Habitat fragmentation in the temperate zone. pp. 237–256. In Soulé, M. E. (ed.) Conservation B; the Science of Scarcity and Diversity. Sunderland: Sinauer Associates.Google Scholar
Wilde, S. A. 1958. Forest Soils: Their Properties and Relation to Silviculture. New York: The Ronald Press Company.Google Scholar
Wilf, P., Cúneo, N. R., Johnson, K. R., Hicks, J. F., Wing, S. L. and Obradovich, J. D.. 2003. High plant diversity in Eocene South America: evidence from Patagonia. Science 300: 122–125.CrossRefGoogle ScholarPubMed
Williams, C. B. 1964. Patterns in the Balance of Nature. London: Academic Press.Google Scholar
Williams, E. J. 1962. The analysis of competition experiments. Australian Journal of Biological Science 15: 509–525.CrossRefGoogle Scholar
Williams, G. C. 1975. Sex and Evolution. Monographs in Population Biology. No. 8. Princeton: Princeton University Press.Google ScholarPubMed
Williams, M. 1989. The lumberman's assault on the southern forest, 1880–1920. pp. 238–288. In Williams, M.. Americans and Their Forests: A Historical Geography. Cambridge: Cambridge University Press.Google Scholar
Williamson, G. B. 1990. Allelopathy, Koch's Postulates and the neck riddle. pp. 143–162. In Grace, J. B. and Tilman, D. (eds.) Perspectives on Plant Competition. San Diego: Academic Press.Google Scholar
Willig, M. R., Kaufman, D. M., and Stevens, R. D.. 2003. Latitudinal gradients of biodiversity: pattern, process, scale and synthesis. Annual Review of Ecology, Evolution and Systematics 34: 273–309.CrossRefGoogle 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: 353–374.CrossRefGoogle Scholar
Willson, M. F. 1984. Mating patterns in plants. pp. 261–276. In Dirzo, R. and Sarukháh, J. (eds.) Perspectives on Plant Population Ecology. Sunderland: Sinauer Associates.Google 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: Sinauer Associates.Google Scholar
Wilson, J. B. 1988. Shoot competition and root competition. Journal of Applied Ecology 25: 279–296.CrossRefGoogle 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: 527–538.CrossRefGoogle Scholar
Wilson, S. D. 1993. Competition and resource availability in heath and grassland in the Snowy Mountains of Australia. Journal of Ecology 81: 445–451.CrossRefGoogle Scholar
Wilson, S. D. 1999. Plant interactions during secondary succession. pp. 629–650. In Walker, L. R. (ed.) Ecosystems of Disturbed Ground. Amsterdam: Elsevier.Google Scholar
Wilson, S. D. and Keddy, P. A.. 1986a. Measuring diffuse competition along an environmental gradient: results from a shoreline plant community. The American Naturalist 127: 862–869.CrossRefGoogle Scholar
Wilson, S. D. and Keddy, P. A.. 1986b. Species competitive ability and position along a natural stress/disturbance gradient. Ecology 67: 1236–1242.CrossRefGoogle Scholar
Wimsatt, W. C. 1982. Reductionistic research strategies and their biases in the units of selection controversy. pp. 155–201. In Saarinen, E. (ed.) Conceptual Issues in Ecology. Dordrecht: D. Reidel.CrossRefGoogle Scholar
Wing, S. L. 1997. Global warming and plant species richness: a case study of the Paleocene/Eocene boundary. pp. 163–185. In Reaka-Kudla, M. L., Wilson, D. E., and Wilson, E. O. (eds.) Biodiversity II: Understanding and Protecting Our Biological Resources. Washington, D.C.: Joseph Henry Press.Google Scholar
Wing, S. L. and B. H. Tiffney. 1987. Interactions of angiosperms and herbivorous tetrapods through time. pp. 203–224. In Friis, E., Chaloner, W. G., and Crane, P. R. (eds.) The Origins of Angiosperms and Their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Wiser, S. K., Peet, R. K., and White, P. S.. 1996. High-elevation rock outcrop vegetation of the southern Appalachian Mountains. Journal of Vegetation Science 7: 703–722.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: 1609–1617.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: 229–238.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: 181–188.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: 147–156.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: 253–258.CrossRefGoogle Scholar
Wisheu, I. C., P. A. Keddy, D. R. J. Moore, S. J. McCanny, and C. L. Gaudet. 1991. Effects of eutrophication on wetland vegetation. pp. 112–121. In Kusler, J. and Smardon, R. (eds.) Wetlands of the Great Lakes: Protection and Restoration Policies; Status of the Science. New York: Managers Inc.Google Scholar
Witmer, M. C. and Cheke, A. S.. 1991. The dodo and the tambalacoque tree: an obligate mutualism reconsidered. Oikos 61: 133–137.CrossRefGoogle Scholar
Wium-Anderson, S. 1971. Photosynthetic uptake of free CO2 by the roots of Lobelia dortmanna. Plantarum 25: 245–248.CrossRefGoogle Scholar
Wolbach, W. S., Lewis, R. S., and Anders, E.. 1985. Cretaceous extinctions: evidence for wildfires and search for meteoritic material. Science 230: 167–170.Google Scholar
Wolfe, J. A. 1991. Palaeobotanical evidence for a June “impact winter” at the Cretaceous/Tertiary boundary. Nature 352: 420–423.CrossRefGoogle Scholar
Wolff, W. J. 1993. Netherlands-wetlands. Hydrobiologia 265: 1–14.CrossRefGoogle Scholar
Wolin, C. L. 1985. The population dynamics of mutualistic systems. pp. 248–269. In Boucher, D. H. (ed.) The Biology of Mutualism: Ecology and Evolution. New York: Oxford University Press.Google Scholar
Woodbury, A. M. 1947. Distribution of pigmy conifers in Utah and northeastern Arizona. Ecology 28: 113–126.CrossRefGoogle Scholar
Woodley, S., Kay, J., and Francis, G.. (eds.) 1993. Ecological Integrity and the Management of Ecosystems. Delray Beach: St. Lucie Press.Google Scholar
Woods, K. D. and R. H. Whittaker. 1981. Canopy–understory interaction and the internal dynamics of mature hardwood and hemlock-hardwood forests. pp. 305–323. In West, D. C., Shugart, H. H., and Botkin, D. B. (eds.) Forest Succession: Concepts and Applications. New York: Springer-Verlag.CrossRefGoogle Scholar
Woodward, F. I. 1987. Climate and Plant Distribution. Cambridge: Cambridge University Press.Google Scholar
Woodward, F. I. 1992. Predicting plant responses to global environmental change. New Phytologist 122: 239–251.CrossRefGoogle Scholar
Woodward, F. I. and C. K. Kelly. 1997. Plant functional types: towards a definition by environmental constraints. pp. 47–65. In Smith, T. M., Shugart, H. H., and Woodward, F. I. (eds.) Plant Functional Types. Cambridge: Cambridge University Press.Google Scholar
Woodwell, G. M. 1962. Effects of ionizing radiation on terrestrial ecosystems. Science 138: 572–577.CrossRefGoogle ScholarPubMed
Woodwell, G. M. 1963. The ecological effects of radiation. Scientific American 208: 42–47.CrossRefGoogle Scholar
World Commission on Environment and Development. 1987. Our Common Future. Oxford: Oxford University Press.
Wright, D. H. and Reeves, J. H.. 1992. On the meaning and measurement of nestedness of species assemblage. Oecologia 92: 416–428.CrossRefGoogle Scholar
Wright, H. A. and Bailey, A. W.. 1982. Fire Ecology. New York: Wiley.Google Scholar
Wright, J. P., Jones, C. G., and Flecker, A. S.. 2002. An ecosystem engineer, the beaver, increases species richness at the landscape scale. Oecologia 132: 96–101.CrossRefGoogle ScholarPubMed
Wright, R. 2004. A Short History of Progress. Toronto: Anansi Press.Google Scholar
Yoda, K., Kira, T., Ogawa, H., and Hozumi, K.. 1963. Self-thinning in overcrowded pure stands under cultivated and natural conditions. Journal of Biology/Osaka City University 14: 107–129.Google Scholar
Yodzis, P. 1986. Competition, mortality, and community structure. pp. 480–492. In Diamond, J. and Case, T. J. (eds.) Community Ecology. New York: Harper and Row.Google Scholar
Yodzis, P. 1989. Introduction to Theoretical Ecology. New York: Harper and Row.Google Scholar
Young, E. 2006. Easter Island: a monumental collapse?New Scientist 2562: 30–34.Google Scholar
Young, K., Ulloa, C. U., Luteyn, J. L., and Knapp, S.. 2002. Plant evolution and endemism in Andean South America: an introduction. The Botanical Review 68: 4–21.CrossRefGoogle Scholar
Young, T. P. and Augspruger, C. K.. 1991. Ecology and evolution of long-lived semelparous plants. Trends in Ecology and Evolution 6: 285–289.CrossRefGoogle ScholarPubMed
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: 1602–1615.CrossRefGoogle Scholar
Zachos, J., Pagani, M., Sloan, L., Thomas, E., and Billups, K.. 2001. Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292: 686–693.CrossRefGoogle ScholarPubMed
Zedler, J. B. and P. A. Beare. 1986. Temporal variability of salt marsh vegetation: the role of low-salinity gaps and environmental stress. pp. 295–306. In Wolfe, D. A. (ed.) Estuarine Variability. San Diego: Academic Press.Google Scholar
Zedler, J. B. and C. P. Onuf. 1984. Biological and physical filtering in arid-region estuaries: seasonality, extreme events, and effects of watershed modification. pp. 415–432. In Kennedy, V. S. (ed.) The Estuary as a Filter. New York: Academic Press.Google Scholar
Zobel, M. 1997. The relative role of species pools in determining plant species richness: an alternative explanation of species coexistence?Trends in Ecology and Evolution 12: 266–269.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 Keddy, Southeastern Louisiana University
  • Book: Plants and Vegetation
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511812989.015
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 Keddy, Southeastern Louisiana University
  • Book: Plants and Vegetation
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511812989.015
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 Keddy, Southeastern Louisiana University
  • Book: Plants and Vegetation
  • Online publication: 05 June 2012
  • Chapter DOI: https://doi.org/10.1017/CBO9780511812989.015
Available formats
×