Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-45l2p Total loading time: 0 Render date: 2024-04-27T23:05:42.785Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  29 March 2019

J. Bastow Wilson
Affiliation:
University of Otago, New Zealand
Andrew D. Q. Agnew
Affiliation:
Aberystwyth University
Stephen H. Roxburgh
Affiliation:
Commonwealth Scientific and Industrial Research Organisation, Canberra
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
Publisher: Cambridge University Press
Print publication year: 2019

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

Aarssen, L.W. (1983) Ecological combining ability and competitive combining ability in plants: towards a general evolutionary theory of coexistence in systems of competition. American Naturalist, 122, 707731.CrossRefGoogle Scholar
Aarssen, L.W. (1985) Interpretation of the evolutionary consequences of competition in plants: an experimental approach. Oikos, 45, 99109.CrossRefGoogle Scholar
Aarssen, L.W. (1988) ‘Pecking order’ of four plant species from pastures of different ages. Oikos, 51, 312.CrossRefGoogle Scholar
Aarssen, L.W. (1989) Competitive ability and species coexistence: a ‘plant’s-eye’ view. Oikos, 56, 386401.CrossRefGoogle Scholar
Aarssen, L.W. & Turkington, R. (1985) Vegetation dynamics and neighbour associations in pasture-community evolution. Journal of Ecology, 73, 585603.Google Scholar
Abades, S.R., Gaxiola, A. & Marquet, P.A. (2014) Fire, percolation thresholds and the savanna forest transition: a neutral model approach. Journal of Ecology, 102, 13861393.CrossRefGoogle Scholar
Abeli, T., Gentili, R., Mondoni, A., Orsenigo, S. & Rossi, G. (2014) Effects of marginality on plant population performance. Journal of Biogeography, 41, 239249.Google Scholar
Abrams, M.D. & Scott, M.L. (1989) Disturbance-mediated accelerated succession in two Michigan forest types. Forest Science, 35, 4249.Google Scholar
Abrams, P.A. (1990) Ecological vs evolutionary consequences of competition. Oikos, 57, 147151.CrossRefGoogle Scholar
Abul-Fatih, H.A. & Bazzaz, F.A. (1979) The biology of Ambrosia trifida L. I. Influence of species removal on the organization of the plant community. New Phytologist, 83, 813816.CrossRefGoogle Scholar
Adams, J.M., Fang, W., Callaway, R.M., Cipollini, D. & Newell, E. (2009) A cross-continental test of the Enemy Release Hypothesis: leaf herbivory on Acer platanoides (L.) is three times lower in North America than in its native Europe. Biological Invasions, 11, 10051016.Google Scholar
Adema, E.B. & Grootjans, A.P. (2003) Possible positive-feedback mechanisms: plants change abiotic soil parameters in wet calcareous dune slacks. Plant Ecology, 167, 141149.CrossRefGoogle Scholar
Adema, E.B., Grootjans, A.P., Petersen, J. & Grijpstra, J. (2002) Alternative stable states in a wet calcareous dune slack in the Netherlands. Journal of Vegetation Science, 13, 107114.Google Scholar
Adema, E.B., Van de Koppel, J., Meijer, H.A.J. & Grootjans, A.P. (2005) Enhanced nitrogen loss may explain alternative stable states in dune slack succession. Oikos, 109, 374386.Google Scholar
Adler, P.B., Lambers, J.H.R., Kyriakidis, P.C., Guan, Q. & Levine, J.M. (2006) Climate variability has a stabilizing effect on the coexistence of prairie grasses. Proceedings of the National Academy of Sciences of the U.S.A., 103, 1279312798.CrossRefGoogle Scholar
Adler, P.B., Lambers, J.H.R. & Levine, J.M. (2009) Weak effect of climate variability on coexistence in a sagebrush steppe community. Ecology, 90, 33033312.CrossRefGoogle Scholar
Agnew, A.D.Q., Rapson, G., Sykes, M.T. & Wilson, J.B. (1993b) The functional ecology of Empodisma minus (Hook.f.) Johnson & Cutler in New Zealand ombrotrophic mires. New Phytologist, 124, 703710.Google Scholar
Agnew, A.D.Q., Wilson, J.B. & Sykes, M.T. (1993a) A vegetation switch as the cause of a forest/mire ecotone in New Zealand. Journal of Vegetation Science, 4, 273278.Google Scholar
Agrawal, A.A. (2004) Resistance and susceptibility of milkweed: competition, root herbivory, and plant genetic variation. Ecology, 85, 21182133.CrossRefGoogle Scholar
Agrawal, A.A., Kotanen, P.M., Mitchell, C.E., Power, A.G., Godsoe, W. & Klironomos, J. (2005) Enemy release? An experiment with congeneric plant pairs and diverse above- and belowground enemies. Ecology, 86, 29792989.Google Scholar
Agrawal, A.A., Lau, J.A. & Hambäck, P.A. (2006) Community heterogeneity and the evolution of interactions between plants and insect herbivores. Quarterly Review of Biology, 81, 349376.CrossRefGoogle ScholarPubMed
Aide, T.M. (1987) Limbfalls: a major cause of sapling mortality for tropical forest plants. Biotropica, 19, 284285.Google Scholar
Aizen, M.A. & Vazquez, D.P. (2006) Flowering phenologies of hummingbird plants from the temperate forest of southern South America: is there evidence of competitive displacement? Ecography, 29, 357366.Google Scholar
Akashi, N., Kohyama, T. & Matsui, K. (2003) Lateral and vertical crown associations in mixed forests. Ecological Research, 18, 455461.CrossRefGoogle Scholar
Alexander, H.D. & Dunton, K.H. (2002) Freshwater inundation effects on emergent vegetation of a hypersaline salt marsh. Estuaries, 25, 14261435.CrossRefGoogle Scholar
Allan, E., van Ruijven, J. & Crawley, M.J. (2010) Foliar fungal pathogens and grassland biodiversity. Ecology, 91, 25722582.CrossRefGoogle ScholarPubMed
Allan, E., Weisser, W., Weigelt, A., Roscher, C., Fischer, M. & Hillebrand, H. (2011) More diverse plant communities have higher functioning over time due to turnover in complementary dominant species. Proceedings of the National Academy of Sciences of the U.S.A., 108, 1703417039.Google Scholar
Allen, B.P., Pauley, E.F. & Sharitz, R.R. (1997) Hurricane impacts on liana populations in an old growth southeastern bottomland forest. Journal of the Torrey Botanical Society, 124, 3442.CrossRefGoogle Scholar
Allen, E.A. (1991) Temporal and spatial organization of desert plant communities. In Semiarid Lands and Deserts: Soil Resource and Reclamation (ed. Skujins, J.), pp. 193208. Dekker, New York.Google Scholar
Allen, E.B. (1988) Some trajectories of succession in Wyoming sagebrush grassland: implications for restoration. In The Reconstruction of Disturbed Arid Lands: An Ecological Approach (ed. Allen, E.B.), pp. 89112. American Association for the Advancement of Science, Washington.Google Scholar
Allen, E.B. & Forman, R.T.T. (1976) Plant species removals and old-field community structure and stability. Ecology, 57, 12331243.Google Scholar
Allen, R.B. & Peet, R.K. (1990) Gradient analysis of forests of the Sangre de Cristo Range, Colorado. Canadian Journal of Botany, 68, 193201.Google Scholar
Allen, R.B., Wilson, J.B. & Mason, C.R. (1995) Vegetation change following exclusion of grazing animals in depleted grassland, Central Otago, New Zealand. Journal of Vegetation Science, 6, 615626.CrossRefGoogle Scholar
Alriksson, A. & Eriksson, H.M. (1998) Variations in mineral nutrient and C distribution in the soil and vegetation compartments of five temperate tree species in NE Sweden. Forest Ecology and Management, 108, 261273.Google Scholar
Andersen, D.C. (1987) Below-ground herbivory in natural communities: a review emphasising fossorial animals. Quarterly Review of Biology, 62, 261286.Google Scholar
Andersen, U.V. (1995) Resistance of Danish coastal vegetation types to human trampling. Biological Conservation, 71, 223230.Google Scholar
Anderson, R.L., Foster, D.R. & Motzkin, G. (2003) Integrating lateral expansion into models of peatland development in temperate New England. Journal of Ecology, 91, 6876.Google Scholar
Angers, D.A. & Caron, J. (1998) Plant-induced changes in soil structure: processes and feedbacks. Biogeochemistry, 42, 5572.Google Scholar
Angert, A.L., Huxman, T.E., Chesson, P. & Venable, L. (2009) Functional tradeoffs determine species coexistence via the storage effect. Proceedings of the National Academy of Sciences of the U.S.A., 106, 1164111645.Google Scholar
Appanah, S. & Putz, F.E. (1984) Climber abundance in virgin dipterocarp forest and the effect of pre-felling climber cutting on logging damage. Malaysian Forester, 47, 335342.Google Scholar
Arceo-Gómez, G. & Ashman, T.L. (2014) Heterospecific pollen receipt affects self pollen more than outcross pollen: implications for mixed-mating plants. Ecology, 95, 29462952.Google Scholar
Archer, S., Scifres, C., Bassham, C.R. & Maggio, R. (1988) Autogenic succession in a subtropical savanna: conversion of grassland to thorn woodland. Ecological Monographs, 58, 111127.Google Scholar
Arianoutsou, M. (1989) Timing of litter production in a maquis ecosystem of North-East Greece. Acta Oecologica Oecologia Plantarum, 10, 371378.Google Scholar
Arizaga, S. & Ezcurra, E. (2002) Propagation mechanisms in Agave macroacantha (Agavaceae), a tropical arid-land succulent rosette. American Journal of Botany, 89, 632641.Google Scholar
Armbruster, W.S. (1986) Reproductive interactions between sympatric Dalechampia species: are natural assemblages “random” or organized? Ecology, 67, 522533.Google Scholar
Armbruster, W.S. (1995) The origins and detection of plant community structure: reproductive versus vegetative processes. Folia Geobotanica et Phytotaxonomica, 30, 483497.CrossRefGoogle Scholar
Armbruster, W.S., Edwards, M.E. & Debevec, E.M. (1994) Floral character displacement generates assemblage structure of Western Australian triggerplants (Stylidium). Ecology, 75, 315329.Google Scholar
Armentrout, S.M. & Pieper, R.D. (1988) Plant distribution surrounding Rocky Mountain pinyon pine and one-seed juniper in south-central New Mexico. Journal of Range Management, 41, 139143.Google Scholar
Ashton, P.S., Givnish, T.J. & Appanah, S. (1988) Staggered flowering in the Dipterocarpaceae: new insights into floral induction and the evolution of mast fruiting in the aseasonal tropics. American Naturalist, 132, 4466.CrossRefGoogle Scholar
Atsatt, P.R. & O’Dowd, D.J. (1976) Plant defense guilds. Science, 193, 2429.CrossRefGoogle ScholarPubMed
Attiwill, P.M. (1994) The disturbance of forest ecosystems: the ecological basis for conservative management. Forest Ecology and Management, 63, 247300.CrossRefGoogle Scholar
Attiwill, P.M. & Wilson, B. (2003) Ecology: An Australian Perspective. Oxford University Press, Melbourne.Google Scholar
Auerbach, M. & Shmida, A. (1993) Vegetation change along an altitudinal gradient on Mt Hermon, Israel – no evidence for discrete communities. Journal of Ecology, 81, 2533.Google Scholar
Augustine, D.J., Frelich, L.E. & Jordan, P.A. (1998) Evidence for two alternate stable states in an ungulate grazing system. Ecological Applications, 8, 12601269.Google Scholar
Augusto, L., Ranger, J., Binkley, D. & Rothe, A. (2002) Impact of several common tree species of European temperate forests on soil fertility. Annals of Forest Science, 59, 233253.Google Scholar
Austin, M.P. (1982) The use of a relative physiological performance value in the prediction of performance in multispecies mixtures from monoculture performance. Journal of Ecology, 70, 559570.Google Scholar
Austin, M.P. (1985) Continuum concept, ordination methods, and niche theory. Annual Review of Ecology and Systematics, 16, 3961.CrossRefGoogle Scholar
Austin, M.P. (2012) Vegetation and environment: discontinuities and continuities. In Vegetation Ecology (eds van der Maarel, E. & Franklin, J.), pp. 71106. Wiley-Blackwell, Malden, MA.Google Scholar
Austin, M.P. & Gaywood, M.J. (1994) Current problems of environmental gradients and species response curves in relation to continuum theory. Journal of Vegetation Science, 5, 473482.Google Scholar
Austin, M.P., Nicholls, A.O., Doherty, M.D. & Meyers, J.A. (1994) Determining species response functions to an environmental gradient by means of a beta-function. Journal of Vegetation Science, 5, 215228.Google Scholar
Austin, M.P. & Smith, T.M. (1989) A new model for the continuum concept. Vegetatio, 83, 3547.Google Scholar
Awasthi, O.P., Sharma, E. & Palni, L.M.S. (1995) Stemflow: a source of nutrients in some naturally growing epiphytic orchids of the Sikkim Himalaya. Annals of Botany, 75, 511.Google Scholar
Babikova, Z., Gilbert, L., Bruce, T.J.A., Birkett, M., Caulfield, J.C., Woodcock, C., Pickett, J.A. & Johnson, D. (2013) Underground signals carried through common mycelial networks warn neighbouring plants of aphid attack. Ecology Letters, 16, 835843.Google Scholar
Badano, E.I., Bustamante, R.O., Villarroel, E., Marquet, P.A. & Cavieres, L.A. (2015) Facilitation by nurse plants regulates community invasibility in harsh environments. Journal of Vegetation Science, 26, 756767.Google Scholar
Bais, H.P., Weir, T.L., Perry, L.G., Gilroy, S. & Vivanco, J.M. (2006) The role of root exudates in rhizosphere interactions with plants and other organisms. Annual Review of Plant Biology, 57, 233266.Google Scholar
Baker, P.J., Bunyavejchewin, S., Oliver, C.D. & Ashton, P.S. (2005) Disturbance history and historical stand dynamics of a seasonal tropical forest in western Thailand. Ecological Monographs, 75, 317343.Google Scholar
Bakker, E.S., Olff, H., Vandenberghe, C., de Maeyer, K., Smit, R., Gleichman, J.M. & Vera, F.W.M. (2004) Ecological anachronisms in the recruitment of temperate light-demanding tree species in wooded pastures. Journal of Applied Ecology, 41, 571582.Google Scholar
Ball, M.C., Egerton, J., Lutze, J.L., Gutschick, V.P. & Cunningham, R.B. (2002) Mechanisms of competition: thermal inhibition of tree seedling growth by grass. Oecologia, 133, 120130.Google Scholar
Ballaré, C.L., Scopel, A.L., Roush, M.L. & Radosevich, S.R. (1995) How plants find light in patchy canopies. A comparison between wild-type and phytochrome-B-deficient mutant plants of cucumber. Functional Ecology, 9, 859868.Google Scholar
Balslev, H., Valencia, R., Paz y Mino, G., Christensen, H. & Nielsen, I. (1998) Species count of vascular plants in one hectare of humid lowland forest in Amazonian Ecuador In Forest Biodiversity in North, Central and South America, and the Caribbean (eds Dallmeier, F. & Comiskey, J.A.), pp. 585594. UNESCO, Paris.Google Scholar
Balyan, R.S., Malik, R.K., Panwar, R.S. & Singh, S. (1991) Competitive ability of winter wheat cultivars with wild oat (Avena ludoviciana). Weed Science, 39, 154158.CrossRefGoogle Scholar
Baraza, E., Zamora, R. & Hodar, J.A. (2006) Conditional outcomes in plant-herbivore interactions: neighbours matter. Oikos, 113, 148156.Google Scholar
Barbour, M.G., Burk, J.H., Pitts, W.D., Gilliam, F.S. & Schwartz, M.W. (1999) Terrestrial Plant Ecology, 3rd edn. Benjamin/Cummings, Menlo Park CA, USA.Google Scholar
Bardgett, R.D., Smith, R.S., Shiel, R.S., Peacock, S., Simkin, J.M., Quirk, H. & Hobbs, P.J. (2006) Parasitic plants indirectly regulate below-ground properties in grassland ecosystems. Nature, 439, 969972.Google Scholar
Barkman, J.J. (1979) The investigation of vegetation texture and structure. In The Study of Vegetation (ed. Werger, M.J.A.), pp. 123160. Junk, The Hague.Google Scholar
Barkman, J.J., Masselink, A.K. & de Vries, B.W.L. (1977) Uber das microclimat in wacholder-fluren. In Vegetation und Klima (ed. Dieschke, H.), pp. 123160. Junk, The Hague.Google Scholar
Barnes, B.V. (1966) The clonal growth habit of American aspens. Ecology, 47, 439447.CrossRefGoogle Scholar
Barot, S. & Gignoux, J. (2004) Mechanisms promoting plant coexistence: can all the proposed processes be reconciled? Oikos, 106, 185192.Google Scholar
Barreiro, R., Guiamét, J.J., Beltrano, J. & Montaldi, E.R. (1992) Regulation of the photosynthetic capacity of primary bean leaves by the red-far-red ratio and photosynthetic photon flux density of incident light. Physiologia Plantarum, 85, 97101.Google Scholar
Barrington, D.S. & Paris, C.A. (2007) Refugia and migration in the quaternary history of the New England flora. Rhodora, 109, 369386.Google Scholar
Barros, C., Thuiller, W., Georges, D., Boulangeat, I. & Münkemüller, T. (2016) N-dimensional hypervolumes to study stability of complex ecosystems. Ecology Letters, 19, 729742.CrossRefGoogle ScholarPubMed
Bartha, S., Czárán, T. & Oborny, B. (1995) Spatial constraints masking community assembly rules: a simulation study. Folia Geobotanica et Phytotaxonomica, 30, 471482.Google Scholar
Bartha, S., Meiners, S.J., Pickett, S.T.A. & Cadenasso, M.L. (2003) Plant colonization windows in a mesic old field succession. Applied Vegetation Science, 6, 205212.Google Scholar
Barto, E.K., Powell, J.R. & Cipollini, D. (2010) How novel are the chemical weapons of garlic mustard in North American forest understories? Biological Invasions, 12, 34653471.Google Scholar
Batanouny, K.H. (1981) Ecology and Flora of Qatar. Alden, Oxford.Google Scholar
Batlla, D., Kruk, C. & Benech-Arnold, R.L. (2000) Very early detection of canopy presence by seeds through perception of subtle modifications in red:far red signals. Functional Ecology, 14, 195202.Google Scholar
Baustian, J.J., Mendelssohn, I.A. & Hester, M.W. (2012) Vegetation’s importance in regulating surface elevation in a coastal salt marsh facing elevated rates of sea level rise. Global Change Biology, 18, 33773382.Google Scholar
Bazzaz, F.A. (1983) Characteristics of populations in natural and man-modified ecosystems. In Disturbance and Ecosystems: Components of Response (eds Mooney, H.A. & Godron, M.), pp. 259275. Springer, Berlin.Google Scholar
Bazzaz, F.A. (1987) Experimental studies on the evolution of niche in successional plant populations. In Colonization, Succession and Stability (ed. Gray, A.J.), pp. 245272. Blackwell, Oxford.Google Scholar
Bazzaz, F.A. & McConnaughay, K.D.M. (1992) Plant-plant interactions in elevated CO2 environments. Australian Journal of Botany, 40, 547563.CrossRefGoogle Scholar
Beans, C.M. (2014) The case for character displacement in plants. Ecology and Evolution, 4, 862875.Google Scholar
Beaton, L.L., VanZandt, P.A., Esselman, E.J. & Knight, T.M. (2011) Comparison of the herbivore defense and competitive ability of ancestral and modern genotypes of an invasive plant, Lespedeza cuneata. Oikos, 120, 14131419.Google Scholar
Beckage, B., Kloeppel, B.D., Yeakley, J., Taylor, S.F. & Coleman, D.C. (2008) Differential effects of understory and overstory gaps on tree regeneration. Journal of the Torrey Botanical Society, 135, 111.CrossRefGoogle Scholar
Becks, L., Ellner, S.P., Jones, L.E. & Hairston, N.G. (2010) Reduction of adaptive genetic diversity radically alters eco-evolutionary community dynamics. Ecology Letters, 13, 989997.Google Scholar
Begon, M., Harper, J.L. & Townsend, J.R. (1996) Ecology: Individuals, Populations and Communities, 3rd edn. Blackwell, Oxford.Google Scholar
Belleau, A., Leduc, A., Lecomte, N. & Bergeron, Y. (2011) Forest succession rate and pathways on different surface deposit types in the boreal forest of northwestern Quebec. Ecoscience, 18, 329340.Google Scholar
Belsky, A.J. (1994) Influences of trees on savanna productivity: tests of shade, nutrients, and tree-grass competition. Ecology, 75, 922932.Google Scholar
Belsky, A.J., Mwonga, S.M., Amundson, R.G., Duxbury, J.M. & Ali, A.R. (1993) Comparative effects of isolated trees on their undercanopy environments in high-rainfall and low-rainfall savannas. Journal of Applied Ecology, 30, 143155.Google Scholar
Belyea, L.R. & Clymo, R.S. (1998) Do hollows control the rate of peat bog growth? In Patterned Mires and Mire Pools (eds Standen, V., Tallis, J.H. & Meade, R.), pp. 5565. Mires Research Group, BES, Durham, UK.Google Scholar
Bengtsson, J., Fagerstrom, T. & Rydin, H. (1994) Competition and coexistence in plant communities. Trends in Ecology and Evolution, 9, 246250.Google Scholar
Bennett, J.A. & Cahill, J.F. (2013) Conservatism of responses to environmental change is rare under natural conditions in a native grassland. Perspectives in Plant Ecology Evolution and Systematics, 15, 328337.Google Scholar
Bennett, J.A. & Pärtel, M. (2017) Predicting species establishment using absent species and functional neighborhoods. Ecology and Evolution, 7, 22232237.Google Scholar
Bennington, J.B. & Bambach, R.K. (1996) Statistical testing for paleocommunity recurrence: are similar fossil assemblages ever the same? Palaeoecology, 127, 107133.Google Scholar
Benzing, D.H. (2004) Vascular epiphytes. In Forest Canopies (eds Lowman, M. & Rinker, H.B.), pp. 175211. Elsevier Academic, Amsterdam, NL.CrossRefGoogle Scholar
Berendse, F. (1994) Litter decomposability – a neglected component of plant fitness. Journal of Ecology, 82, 187190.Google Scholar
Berendse, F. (1999) Implications of increased litter production for plant biodiversity. Trends in Ecology and Evolution, 14, 45.CrossRefGoogle ScholarPubMed
Berendse, F. & Aerts, R. (1984) Competition between Erica tetralix L. and Molinia caerulea (L.) Moench as affected by the availability of nutrients. Acta Oecologica Oecologia Plantarum, 5, 314.Google Scholar
Berendse, F., Oudhof, H. & Bol, J. (1987) A comparative study on nutrient cycling in wet heathland ecosystems. I. Litter production and nutrient losses from the plant. Oecologia, 74, 174184.Google Scholar
Berendse, F., Schmitz, M. & de Visser, W. (1994) Experimental manipulation of succession in heathland ecosystems. Oecologia, 100, 3844.Google Scholar
Berg, S.S. & Dunkerley, D.L. (2004) Patterned mulga near Alice Springs, Central Australia, and the potential threat of firewood collection in this vegetation. Journal of Arid Environments, 59, 313350.Google Scholar
Berg-Binder, M.C. & Suarez, A.V. (2012) Testing the directed dispersal hypothesis: are native ant mounds (Formica sp.) favorable microhabitats for an invasive plant? Oecologia, 169, 763772.Google Scholar
Bergelson, J. (1990) Life after death: site pre-emption by the remains of Poa annua. Ecology, 71, 21572165.Google Scholar
Bergengren, J.C., Thompson, S.L., Pollard, D. & DeConto, R.M. (2001) Modeling global climate-vegetation interactions in a doubled CO2 world. Climatic Change, 50, 3175.Google Scholar
Berkley, H.A., Kendall, B.E., Mitarai, S. & Siegel, D.A. (2010) Turbulent dispersal promotes species co-existence. Ecology Letters, 13, 360371.Google Scholar
Bernard-Verdier, M., Navas, M.L., Vellend, M., Violle, C., Fayolle, A. & Garnier, E. (2012) Community assembly along a soil depth gradient: contrasting patterns of plant trait convergence and divergence in a Mediterranean rangeland. Journal of Ecology, 100, 14221433.Google Scholar
Berntson, G.M. & Wayne, P.M. (2000) Characterizing the size dependence of resource acquisition within crowded plant populations. Ecology, 81, 10721085.Google Scholar
Bertness, M.D. & Callaway, R. (1994) Positive interactions in communities. Trends in Ecology and Evolution, 9, 191193.Google Scholar
Bertness, M.D. & Hacker, S.D. (1994) Physical stress and positive associations among marsh plants. American Naturalist, 144, 363372.Google Scholar
Bertness, M.D. & Yeh, S.M. (1994) Cooperative and competitive interactions in the recruitment of marsh elders. Ecology, 75, 24162429.Google Scholar
Bever, J.D. (2003) Soil community feedback and the coexistence of competitors: conceptual frameworks and empirical tests. New Phytologist, 157, 465473.Google Scholar
Bezemer, T.M., Fountain, M.T., Barea, J.M., Christensen, S., Dekker, S.C., Duyts, H., van Hal, R., Harvey, J.A., Hedlund, K., Maraun, M., Mikola, J., Mladenov, A.G., Robin, C., de Ruiter, P.C., Scheu, S., Setälä, H., Šmilauer, P. & van der Putten, W.H. (2010) Divergent composition but similar function of soil food webs of individual plants: plant species and community effects. Ecology, 91, 30273036.Google Scholar
Bhatt, M.V., Khandelwal, A. & Dudley, S.A. (2011) Kin recognition, not competitive interactions, predicts root allocation in young Cakile edentula seedling pairs. New Phytologist, 189, 11351142.Google Scholar
Biddington, N.L. & Dearman, A.S. (1985) The effects of mechanically-induced stress on the growth of cauliflower, lettuce and celery seedlings. Annals of Botany, 55, 109119.Google Scholar
Biedrzycki, M.L. & Bais, H.P. (2010) Kin recognition in plants: a mysterious behaviour unsolved. Journal of Experimental Botany, 61, 41234128.Google Scholar
Binkley, D. & Valentine, D. (1991) 50-year biogeochemical effects of green ash, white-pine, and Norway spruce in a replicated experiment. Forest Ecology and Management, 40, 1325.CrossRefGoogle Scholar
Bio, A.M.F. (2000) Does vegetation suit our models? Data and model assumptions and the assessment of species distribution in space. PhD thesis, Utrecht University.Google Scholar
Birkett, M.A., Campbell, C.A., Chamberlain, K., Guerrieri, E., Hick, A.J., Martin, J.L., Matthes, M., Napier, J.A., Pettersson, J., Pickett, J.A., Poppy, G.M., Pow, E.M., Pye, B.J., Smart, L.E., Wadhams, G.H., Wadhams, L.J. & Woodcock, C.M. (2000) New roles for cis-jasmone as an insect semiochemical and in plant defense. Proceedings of the National Academy of Sciences of the U.S.A., 97, 93299334.Google Scholar
Birks, H.J.B. (1993) Quaternary palaeoecology and vegetation science: current contributions and possible future developments. Review of Palaeobotany and Palynology, 79, 153177.Google Scholar
Birouste, M., Kazakou, E., Blanchard, A. & Roumet, C. (2012) Plant traits and decomposition: are the relationships for roots comparable to those for leaves? Annals of Botany, 109, 463472.Google Scholar
Bittebiere, A.K., Renaud, N., Clement, B. & Mony, C. (2012) Morphological response to competition for light in the clonal Trifolium repens (Fabaceae). American Journal of Botany, 99, 646654.CrossRefGoogle ScholarPubMed
Black, J.N. (1958) Competition between plants of different initial seed sizes in swards of subterranean clover (Trifolium subterraneum L.) with particular reference to leaf area and the light microclimate. Australian Journal of Agricultural Research, 9, 299318.Google Scholar
Blair, A.C. & Wolfe, L.M. (2004) The evolution of an invasive plant: an experimental study with Silene latifolia. Ecology, 85, 30353042.Google Scholar
Blair, B. (2001) Effect of soil nutrient heterogeneity on the symmetry of belowground competition. Plant Ecology, 156, 199203.Google Scholar
Blindow, I., Andersson, G., Hargeby, A. & Johansson, S. (1993) Long-term pattern of alternative stable states in two shallow eutrophic lakes. Freshwater Biology, 30, 159167.Google Scholar
Blindow, I., Hargeby, A. & Andersson, G. (2002) Seasonal changes of mechanisms maintaining clear water in a shallow lake with abundant Chara vegetation. Aquatic Botany, 72, 315334.Google Scholar
Blondel, J. (2003) Guilds or functional groups: does it matter? Oikos, 100, 223231.Google Scholar
Bobiwash, K., Schultz, S.T. & Schoen, D.J. (2013) Somatic deleterious mutation rate in a woody plant: estimation from phenotypic data. Heredity, 111, 338344.Google Scholar
Bode, M., Bode, L. & Armsworth, P.R. (2011) Different dispersal abilities allow reef fish to co-exist. Proceedings of the National Academy of Sciences of the U.S.A., 108, 1631716321.Google Scholar
Boege, K. (2004) Induced responses in three tropical dry forest plant species – direct and indirect effects on herbivory. Oikos, 107, 541548.Google Scholar
Bogaard, A., Hodgson, J.G., Wilson, P.J. & Band, S.R. (1998) An index of weed size for assessing the soil productivity of ancient crop fields. Vegetation History and Archaeobotany, 7, 1722.Google Scholar
Bolker, B.M. & Pacala, S.W. (1999) Spatial moment equations for plant competition: understanding spatial strategies and the advantages of short dispersal. American Naturalist, 153, 575602.CrossRefGoogle ScholarPubMed
Bonan, G.B., Pollard, D. & Thompson, S.L. (1992) Effects of boreal forest vegetation on global climate. Nature, 359, 716718.Google Scholar
Bond, W.J. (1993) Keystone species. In Ecosystem Functioning and Biodiversity (eds Schulze, E.-D. & Mooney, H.A.), pp. 237253. Springer, Berlin.Google Scholar
Booth, B.D. & Larson, D.W. (1999) Impact of language, history and choice of system on the study of assembly rules. In Ecological Assembly Rules: Perspectives, Advances, Retreats (eds Weiher, E. & Keddy, P.A.), pp. 206232. Cambridge University Press, Cambridge, UK.Google Scholar
Booth, M.G. (2004) Mycorrhizal networks mediate overstorey-understorey competition in a temperate forest. Ecology Letters, 7, 538546.Google Scholar
Borrelli, J.J. (2015) Selection against instability: stable subgraphs are most frequent in empirical food webs. Oikos, 124, 15831588.Google Scholar
Bossdorf, O., Auge, H., Lafuma, L., Rogers, W.E., Siemann, E. & Prati, D. (2005) Phenotypic and genetic differentiation between native and introduced plant populations. Oecologia, 144, 111.Google Scholar
Bossuyt, B., Honnay, O. & Hermy, M. (2005) Evidence for community assembly constraints during succession in dune slack plant communities. Plant Ecology, 178, 201209.Google Scholar
Bosy, J.L. & Reader, R.J. (1995) Mechanisms underlying the suppression of forb seedling emergence by grass (Poa pratensis) litter. Functional Ecology, 9, 635639.Google Scholar
Boyce, C.K., Lee, J.E., Feild, T.S., Brodribb, T.J. & Zwieniecki, M.A. (2010) Angiosperms helped put the rain in the rainforests: the impact of plant physiological evolution on tropical biodiversity. Annals of the Missouri Botanical Garden, 97, 527540.Google Scholar
Boyd, R. & Jaffré, T. (2001) Phytoenrichment of soil Ni content by Sebertia acuminata in New Caledonia and the concept of elemental allelopathy. South African Journal of Science, 97, 535538.Google Scholar
Braam, J. & Davis, R. (1990) Rain-, wind-, and touch-induced expression of calmodulin and calmodulin-related genes in Arabidopsis. Cell, 60, 357364.Google Scholar
Bracken, M.E.S. & Low, N.H.N. (2012) Realistic losses of rare species disproportionately impact higher trophic levels. Ecology Letters, 15, 461467.Google Scholar
Bradshaw, A.D. (1965) Evolutionary significance of phenotypic plasticity in plants. Advances in Genetics, 13, 115155.Google Scholar
Braun-Blanquet, J. (1932) Plant Sociology: The Study of Plant Communities. McGraw-Hill, New York.Google Scholar
Bray, R.H. (1954) A nutrient mobility concept of soil-plant relationships. Soil Science, 78, 922.Google Scholar
Brenchley, W.E. (1916) The effect of the concentration of the nutrient solution on the growth of barley and wheat in water cultures. Annals of Botany, 30, 7790.Google Scholar
Brittingham, S. & Walker, L.R. (2000) Facilitation of Yucca brevifolia recruitment by Mojave Desert shrubs. Western North American Naturalist, 60, 374383.Google Scholar
Brokaw, N. & Busing, R.T. (2000) Niche versus chance and tree diversity in forest gaps. Trends in Ecology and Evolution, 15, 183188.Google Scholar
Brooker, R.W., Scott, D., Palmer, S.C.F. & Swaine, E. (2006) Transient facilitative effects of heather on Scots pine along a grazing disturbance gradient in Scottish moorland. Journal of Ecology, 94, 637645.Google Scholar
Brooks, M.L., D’Antonio, C.M., Richardson, D.M., Grace, J.B., Keeley, J.E., DiTomaso, J.M., Hobbs, R.J., Pellant, M. & Pyke, D. (2004) Effects of invasive alien plants on fire regimes. BioScience, 54, 677688.Google Scholar
Brown, C., Law, R., Illian, J.B. & Burslem, D.F.R.P. (2011) Linking ecological processes with spatial and non-spatial patterns in plant communities. Journal of Ecology, 99, 14021414.Google Scholar
Brown, P.M. & Sieg, C.H. (1999) Historical variability in fire at the ponderosa pine – Northern Great Plains prairie ecotone, southeastern Black Hills, South Dakota. Ecoscience, 6, 539547.Google Scholar
Brown, V.K. (1993) Herbivory: a structuring force in plant communities. In Individuals, Populations and Patterns in Ecology (eds Leather, S.R., Walters, K., Mills, N. & Watt, A.), pp. 299308. Intercept, Andover.Google Scholar
Brown, V.K. & Gange, A.C. (1989a) Herbivory by soil-dwelling insects depresses plant species richness. Functional Ecology, 3, 667671.Google Scholar
Brown, V.K. & Gange, A.C. (1989b) Differential effects of above- and below-ground insect herbivory during early plant succession. Oikos, 54, 6776.Google Scholar
Brownstein, G., Steel, J.B., Porter, S., Gray, A., Wilson, C., Wilson, P.G. & Wilson, J.B. (2012) Chance in plant communities: a new approach to its measurement using the nugget from spatial autocorrelation. Journal of Ecology, 100, 987996.Google Scholar
Buchmann, N., Kao, W.-Y. & Ehleringer, J.R. (1996) Carbon dioxide concentrations within forest canopies – variation with time, stand structure, and vegetation type. Global Change Biology, 2, 421432.Google Scholar
Buck-Sorlin, G.H. & Bell, A.D. (1998) A quantification of shoot shedding in the pedunculate oak (Quercus robur L.). Botanical Journal of the Linnean Society, 127, 371391.Google Scholar
Buck-Sorlin, G.H. & Bell, A.D. (2000) Crown architecture in Quercus petraea and Q. robur: the fate of buds and shoots in relation to age, position and environmental perturbation. Forestry, 73, 331349.Google Scholar
Bullock, J.M., Joe Franklin, J., Stevenson, M.J., Silvertown, J., Coulson, S.J., Gregory, S.J. & Richard Tofts, R. (2001) A plant trait analysis of responses to grazing in a long-term experiment. Journal of Applied Ecology, 38, 253267.Google Scholar
Burdon, J.J., Wennström, A., Ericson, L., Müller, W.J. & Morton, R. (1992) Density-dependent mortality in Pinus sylvestris caused by the snow blight pathogen Phacidium infestans. Oecologia, 90, 7479.Google Scholar
Burke, D.J. (2012) Shared mycorrhizal networks of forest herbs: does the presence of conspecific and heterospecific adult plants affect seedling growth and nutrient acquisition? Botany, 90, 10481057.Google Scholar
Burke, M.J.W. & Grime, J.P. (1996) An experimental study of plant community invasibility. Ecology, 77, 776790.Google Scholar
Burns, K.C. (2005) Is there limiting similarity in the phenology of fleshy fruits? Journal of Vegetation Science, 16, 617624.Google Scholar
Busing, R.T. (1996) Estimation of tree replacement patterns in an Appalachian Picea-Abies forest. Journal of Vegetation Science, 7, 685694.Google Scholar
Buss, L.W. & Jackson, J.B.C. (1979) Competitive networks: nontransitive competitive relationships in cryptic coral reef environments. American Naturalist, 113, 223234.Google Scholar
Bycroft, C.M., Nicolaou, N., Smith, B. & Wilson, J.B. (1993) Community structure (niche limitation and guild proportionality) in relation to the effect of spatial scale, in a Nothofagus forest sampled with a circular transect. New Zealand Journal of Ecology, 17, 95101.Google Scholar
Caccianiga, M., Luzzaro, A., Pierce, S., Ceriani, R.M. & Cerabolini, B. (2006) The functional basis of a primary succession resolved by C-S-R classification. Oikos, 112, 1020.Google Scholar
Cadisch, G. & Giller, K.E. (1997) Driven by Nature: Plant Litter Quality and Decomposition. CAB International, Cambridge.Google Scholar
Cadotte, M.W., Dinnage, R. & Tilman, D. (2012) Phylogenetic diversity promotes ecosystem stability. Ecology, 93, 223233.Google Scholar
Cahill, J.F. (1999) Fertilization effects on interactions between above- and belowground competition in an old field. Ecology, 80, 466480.Google Scholar
Cairney, J.W.G. & Ashford, A.E. (2002) Biology of mycorrhizal associations of epacrids (Ericaceae). New Phytologist, 154, 305326.Google Scholar
Calcagno, V., Sun, C., Schmitz, O.J. & Loreau, M. (2011) Keystone predation and plant species coexistence: the role of carnivore hunting mode. American Naturalist, 177, 113.Google Scholar
Caldwell, M.M., Dawson, T.E. & Richards, J.H. (1998) Hydraulic lift: consequences of water efflux from the roots of plants. Oecologia, 113, 151161.Google Scholar
Callaway, R.M. (1995) Positive interactions among plants. Botanical Review, 61, 306349.Google Scholar
Callaway, R.M. (1998) Are positive interactions species-specific? Oikos, 82, 202207.Google Scholar
Callaway, R.M. & Aschehoug, E.T. (2000) Invasive plants versus their new and old neighbors: a mechanism for exotic invasion. Science, 290, 521523.Google Scholar
Callaway, R.M., Brooker, R.W., Choler, P., Kikvidze, Z., Lortie, C.J., Michalet, R., Paolini, L., Pugnaire, F.I., Newingham, B., Aschehoug, E.T., Armas, C., Kikodze, D. & Cook, B.J. (2002) Positive interactions among alpine plants increase with stress. Nature, 417, 844848.Google Scholar
Callaway, R.M., Ridenour, W.M., Laboski, T., Weir, T. & Vivanco, J.M. (2005) Natural selection for resistance to the allelopathic effects of invasive plants. Journal of Ecology, 93, 576583.Google Scholar
Callaway, R.M., Thelen, G.C., Rodriguez, A. & Holben, W.E. (2004) Soil biota and exotic plant invasion. Nature, 427, 731733.Google Scholar
Campbell, B.D. & Grime, J.P. (1992) An experimental test of plant strategy theory. Ecology, 73, 1529.Google Scholar
Campbell, B.D., Grime, J.P. & Mackey, J.M.L. (1992) Shoot thrust and its role in plant competition. Journal of Ecology, 80, 633641.Google Scholar
Campbell, S.P., Witham, J.W. & Hunter, M.L. (2010) Stochasticity as an alternative to deterministic explanations for patterns of habitat use by birds. Ecological Monographs, 80, 287302.Google Scholar
Cape, J.N., Brown, A.H.F., Robertson, S.M.C., Howson, G. & Paterson, I.S. (1991) Interspecies comparisons of throughfall and stemflow at three sites in northern Britain. Forest Ecology and Management, 46, 165177.Google Scholar
Cappellato, R. & Peters, N.E. (1995) Dry deposition and canopy leaching rates in deciduous and coniferous forests of the Georgia Piedmont: an assessment of a regression model. Journal of Hydrology, 169, 131150.Google Scholar
Cappuccino, N. & Carpenter, D. (2005) Invasive exotic plants suffer less herbivory than non-invasive exotic plants. Biology Letters, 1, 435438.Google Scholar
Cardinale, B.J., Gross, K., Fritschie, K., Flombaum, P., Fox, J.W., Rixen, C., van Ruijven, J., Reich, P.B., Scherer-Lorenzen, M. & Wilsey, B.J. (2013) Biodiversity simultaneously enhances the production and stability of community biomass, but the effects are independent. Ecology, 94, 16971707.Google Scholar
Cardinale, B.J., Wright, J.P., Cadotte, M.W., Carroll, I.T., Hector, A., Srivastava, D.S., Loreau, M. & Weis, J.J. (2007) Impacts of plant diversity on biomass production increase through time because of species complementarity. Proceedings of the National Academy of Sciences of the U.S.A., 104, 1812318128.Google Scholar
Carey, P.D. & Watkinson, A.R. (1993) The dispersal and fates of seeds of the winter annual grass Vulpia ciliata. Journal of Ecology, 81, 759767.Google Scholar
Carey, P.D., Watkinson, A.R. & Gerard, F.F.O. (1995) The determinants of the distribution and abundance of the winter annual grass Vulpia ciliata ssp. ambigua. Journal of Ecology, 83, 177187.Google Scholar
Carlucci, M.B., Debastiani, V.J., Pillar, V.D. & Duarte, L.D.S. (2015) Between- and within-species trait variability and the assembly of sapling communities in forest patches. Journal of Vegetation Science, 26, 2131.Google Scholar
Caron, M.N., Kneeshaw, D.D., Degrandpre, L., Kauhanen, H. & Kuuluvainen, T. (2009) Canopy gap characteristics and disturbance dynamics in old- growth Picea abies stands in northern Fennoscandia: is the forest in quasi-equilibrium? Annales Botanici Fennici, 46, 251262.Google Scholar
Carroll, I.T., Cardinale, B.J. & Nisbet, R.M. (2011) Niche and fitness differences relate the maintenance of diversity to ecosystem function. Ecology, 92, 11571165.Google Scholar
Carson, W.P. & Root, R.B. (2000) Herbivory and plant species coexistence: community regulation by an outbreaking phytophagous insect. Ecological Monographs, 70, 7399.Google Scholar
Carter, R.N. & Prince, S.D. (1985) The geographical distribution of prickly lettuce (Lactuca serriola). I. A general survey of its habitats and performance in Britain. Journal of Ecology, 73, 2738.Google Scholar
Casas, G., Scrosati, R. & Piriz, M.L. (2004) The invasive kelp Undaria pinnatifida (Phaeophyceae, Laminariales) reduces native seaweed diversity in Nuevo Gulf (Patagonia, Argentina). Biological Invasions, 6, 411416.Google Scholar
Case, T.J. (1991) Invasion resistance, species build-up and community collapse in metapopulation models with interspecific competition. Biological Journal of the Linnean Society, 42, 239266.Google Scholar
Cash, F.B.J., Conn, A., Coutts, S., Stephen, M., Mason, N.W.H. & Wilson, J.B. (2012) Assembly rules operate only in equilibrium communities: Is it true? Austral Ecology, 37, 903914.Google Scholar
Castellanos, A.E., Martinez, M.J., Llano, J.M., Halvorson, W.L., Espiricueta, M. & Espejel, I. (2005) Successional trends in Sonoran Desert abandoned agricultural fields in northern Mexico. Journal of Arid Environments, 60, 437455.Google Scholar
Caswell, H. (1976) Community structure: a neutral model analysis. Ecological Monographs, 46, 327354.Google Scholar
Cavender-Bares, J., Ackerly, D.D., Baum, D.A. & Bazzaz, F.A. (2004) Phylogenetic overdispersion in Floridian oak communities. American Naturalist, 163, 823843.Google Scholar
Cavieres, L.A., Badano, E.I., Sierra-Almeida, A. & Molina-Montenegro, M.A. (2007) Microclimatic modifications of cushion plants and their consequences for seedling survival of native and non-native herbaceous species in the high Andes of central Chile. Arctic Antarctic and Alpine Research, 39, 229236.Google Scholar
Cavieres, L.A., Quiroz, C.L. & Molina-Montenegro, M.A. (2008) Facilitation of the non-native Taraxacum officinale by native nurse cushion species in the high Andes of central Chile: are there differences between nurses? Functional Ecology, 22, 148156.Google Scholar
Certini, G. (2005) Effects of fire on properties of forest soils: a review. Oecologia, 143, 110.Google Scholar
Challinor, D. (1968) Alteration of surface soil characteristics by four tree species. Ecology, 49, 286290.Google Scholar
Chambers, J.Q., Higuchi, N., Schimel, J.P., Ferreira, L.V. & Melack, J.M. (2000) Decomposition and carbon cycling of dead trees in tropical forests of the central Amazon. Oecologia, 122, 380388.Google Scholar
Chandrashekara, U.M. & Ramakrishnan, P.S. (1994) Vegetation and gap dynamics of a tropical wet evergreen forest in the Western Ghats of Kerala, India. Journal of Tropical Ecology, 10, 337354.Google Scholar
Chanway, C.P., Holl, F.B. & Turkington, R. (1989) Effect of Rhizobium leguminosarum biovar trifolii on specificity between Trifolium repens and Lolium perenne. Journal of Ecology, 77, 11501160.Google Scholar
Chapin, F.S., Walker, L.R., Fastie, C.L. & Sharman, L.C. (1994) Mechanisms of primary succession following deglaciation at Glacier Bay, Alaska. Ecological Monographs, 64, 149175.Google Scholar
Chase, J.M. (2003) Community assembly: when should history matter? Oecologia, 136, 489498.Google Scholar
Chase, J.M. (2010) Stochastic community assembly causes higher biodiversity in more productive environments. Science, 328, 13881391.Google Scholar
Chase, J.M. (2014) Spatial scale resolves the niche versus neutral theory debate. Journal of Vegetation Science, 25, 319322.Google Scholar
Chase, J.M. & Leibold, M.A. (2003) Ecological Niches: Linking Classical and Contemporary Approaches. Chicago University Press, Chicago, IL.Google Scholar
Chase, J.M. & Myers, J.A. (2011) Disentangling the importance of ecological niches from stochastic processes across scales. Philosophical Transactions of the Royal Society of London B, 366, 23512363.Google Scholar
Chave, J. (2004) Neutral theory and community ecology. Ecology Letters, 7, 241253.Google Scholar
Chen, J., Franklin, J.F. & Spies, T.A. (1993) Contrasting microclimates among clearcut, edge, and interior of old-growth Douglas-fir forest. Agricultural and Forest Meteorology, 63, 219237.Google Scholar
Chen, Y., Wright, S.J., Muller-Landau, H.C., Hubbell, S.P., Wang, Y. & Yu, S. (2016) Positive effects of neighborhood complementarity on tree growth in a Neotropical forest. Ecology, 97, 776785.Google Scholar
Cheng, J.J., Mi, X.C., Nadrowski, K., Ren, H.B., Zhang, J.T. & Ma, K.P. (2012) Separating the effect of mechanisms shaping species-abundance distributions at multiple scales in a subtropical forest. Oikos, 121, 236244.Google Scholar
Chesson, P. & Huntly, N. (1997) The roles of harsh and fluctuating conditions in the dynamics of ecological communities. American Naturalist, 150, 519553.Google Scholar
Chesson, P.L. (1985) Coexistence of competitors in spatially and temporally varying environments: a look at the combined effects of different sorts of variability. Theoretical Population Biology, 28, 263287.Google Scholar
Chesson, P.L. (1990) Geometry, heterogeneity and competition in variable environments. Philosophical Transactions of the Royal Society of London B, 330, 165173.Google Scholar
Chesson, P.L. (2000a) Mechanisms of maintenance of species diversity. Annual Review of Ecology and Systematics, 31, 343366.Google Scholar
Chesson, P.L. (2000b) General theory of competitive coexistence in spatially varying environments. Theoretical Population Biology, 58, 211237.Google Scholar
Chesson, P.L. (2008) Quantifying and testing species coexistence mechanisms. In Unity in Diversity: Reflections on Ecology after the Legacy of Ramon Margalef (eds Valladares, F., Camacho, A., Elosegi, A., Gracia, C., Estrada, M., Senar, J.C. & Gili, J.-P.), pp. 119164. Fundation Banco Bilbao-Vizcaya Argentaria, Bilbao.Google Scholar
Chesson, P.L., Donahue, M., Melbourne, B. & Sears, A. (2005) Scale transition theory for understanding mechanisms in metacommunities. In Metacommunities: Spatial Dynamics and Ecological Communities (eds Holyoak, M., Leibold, M.A. & Holt, R.D.), pp. 279306. The University of Chicago Press, Chicago, IL.Google Scholar
Chiarucci, A., Mistral, M., Bonini, I., Anderson, B.J. & Wilson, J.B. (2002) Canopy occupancy: how much of the space in plant communities is filled? Folia Geobotanica, 37, 333338.Google Scholar
Chisholm, R.A. & Pacala, S.W. (2010) Niche and neutral models predict asymptotically equivalent species abundance distributions in high-diversity ecological communities. Proceedings of the National Academy of Sciences of the U.S.A., 107, 1582115825.Google Scholar
Choi, J.H., Fushimi, K., Abe, N., Tanaka, H., Maeda, S., Morita, A., Hara, M., Motohashi, R., Matsunaga, J., Eguchi, Y., Ishigaki, N., Hashizume, D., Koshino, H. & Kawagishi, H. (2010) Disclosure of the “fairy” of fairy-ring-forming fungus Lepista sordida. ChemBioChem, 11, 13731377.Google Scholar
Christensen, O. (1975) Wood litter fall in relation to abscission, environmental factors, and the decomposition cycle in a Danish oak forest. Oikos, 26, 187195.Google Scholar
Cipollini, D., Rigsby, C.M. & Barto, E.K. (2012) Microbes as targets and mediators of allelopathy in plants. Journal of Chemical Ecology, 38, 714727.Google Scholar
Clark, D.B. & Clark, D.A. (1991) The impact of physical damage on canopy tee regeneration in tropical rain forest. Journal of Ecology, 79, 447457.Google Scholar
Clark, J.S., Dietze, M., Chakraborty, S., Agarwal, P.K., Ibanez, I., LaDeau, S. & Wolosin, M. (2007) Resolving the biodiversity paradox. Ecology Letters, 10, 647659.Google Scholar
Clarke, P.J. (2002) Experiments on tree and shrub establishment in temperate grassy woodlands: Seedling survival. Austral Ecology, 27, 606615.Google Scholar
Clarkson, B.R., Schipper, L.A. & Silvester, W.B. (2009) Nutritional niche separation in coexisting bog species demonstrated by 15N-enriched simulated rainfall. Austral Ecology, 34, 377385.Google Scholar
Clements, C.F., Warren, P.H., Collen, B., Blackburn, T., Worsfold, N. & Petchey, O. (2013) Interactions between assembly order and temperature can alter both short- and long-term community composition. Ecology and Evolution, 3, 52015208.Google Scholar
Clements, F.E. (1904) Studies on the vegetation of the state. III. The development and structure of vegetation. Reports of the Botanical Survey of Nebraska, 7, 1175.Google Scholar
Clements, F.E. (1905) Research Methods in Ecology. University Publishing, Lincoln, Nebraska.Google Scholar
Clements, F.E. (1907) Plant Physiology and Ecology. Henry Holt, New York.Google Scholar
Clements, F.E. (1916) Plant Succession: An Analysis of the Development of Vegetation. Carnegie Institute of Washington, Washington.Google Scholar
Clements, F.E. (1920) Plant Indicators: The Relation of Plant Communities to Process and Practice. Carnegie Institution of Washington, Washington.Google Scholar
Clements, F.E. (1929) The phylogeny of climaxes. Yearbook Carnegie Institute of Washington, 28, 202203.Google Scholar
Clements, F.E. (1934) The relict method in dynamic ecology. Journal of Ecology, 22, 3968.Google Scholar
Clements, F.E. (1935) Experimental ecology in the public service. Ecology, 16, 342363.Google Scholar
Clements, F.E. (1936) Nature and structure of the climax. Journal of Ecology, 24, 252284.Google Scholar
Clements, F.E. & Goldsmith, G.W. (1924) The Phytometer Method in Ecology. Carnegie Institution of Washington, Washington.Google Scholar
Clements, F.E. & Shelford, V.E. (1939) Bio-ecology. Wiley, New York.Google Scholar
Clements, F.E., Weaver, J.E. & Hanson, H.C. (1929) Plant Competition: An Analysis of Community Functions. Carnegie Institute of Washington, Washington.Google Scholar
Clymo, R.S. & Hayward, P.M. (1982) The ecology of Sphagnum. In Bryophyte Ecology (ed. Smith, A.J.E.), pp. 229289. Chapman and Hall, London.Google Scholar
Cockayne, L. (1926) Monograph on The New Zealand Beech Forests. I. The Ecology of the Forests and Taxonomy of the Beeches. Government Printer, Wellington.Google Scholar
Cody, M.L. (1966) A general theory of clutch size. Evolution, 20, 174184.Google Scholar
Cody, M.L. (1974) Competition and the Structure of Bird Communities. Princeton University Press, Princeton, NJ.Google Scholar
Cody, M.L. (1986) Structural niches in plant communities. In Community Ecology (eds Diamond, J.M. & Case, T.J.), pp. 381405. Harper and Row, New York.Google Scholar
Cody, M.L. (1989) Discussion: Structure and assembly of communities. In Perspectives in Ecological Theory (eds Roughgarden, J., May, R.M. & Levin, S.A.), pp. 227241. Princeton University Press, Princeton, NJ.Google Scholar
Cody, M.L. & Prigge, B.A. (2003) Spatial and temporal variations in the timing of leaf replacement in a Quercus cornelius-mulleri population. Journal of Vegetation Science, 14, 789798.Google Scholar
Cohen, J.E. (1968) Alternate derivations of a species-abundance relation. American Naturalist, 102, 165172.Google Scholar
Cole, D.N. (1995) Experimental trampling of vegetation. II. Predictors of resistance and resilience. Journal of Applied Ecology, 32, 215224.Google Scholar
Cole, D.N. & Spildie, D.R. (1998) Hiker, horse and llama trampling effects on native vegetation in Montana, USA. Journal of Environmental Management, 53, 6171.Google Scholar
Cole, D.N. & Trull, S.J. (1992) Quantifying vegetation response to recreational disturbance in the North Cascades, Washington. Northwest Science, 66, 229236.Google Scholar
Collins, S.L., Bradford, J.A. & Sims, P.L. (1987) Succession and fluctuation in Artemisia dominated grassland. Vegetatio, 73, 8999.Google Scholar
Collins, S.L., Suding, K.N., Cleland, E.E., Batty, M., Pennings, S.C., Gross, K.L., Grace, J.B., Gough, L., Fargione, J.E. & Clark, C.M. (2008) Rank clocks and plant community dynamics. Ecology, 89, 35343541.Google Scholar
Colwell, R.K. & Rangel, T.F. (2009) Hutchinson’s duality: the once and future niche. Proceedings of the National Academy of Sciences of the U.S.A., 106, 1965119658.Google Scholar
Connell, J.H. (1978) Diversity in tropical rain forests and coral reefs. Science, 199, 13021310.Google Scholar
Connell, J.H. & Slatyer, R.O. (1977) Mechanisms of succession in natural communities and their role in community stability and organisation. American Naturalist, 111, 11191144.Google Scholar
Connell, J.H. & Sousa, W.P. (1983) On the evidence needed to judge ecological stability or persistence. American Naturalist, 121, 789824.Google Scholar
Connolly, J. (1997) Substitutive experiments and the evidence for competitive hierarchies in plant communities. Oikos, 80, 179182.Google Scholar
Connor, E.F. & McCoy, E.D. (1979) The statistics and biology of the species-area relationship. American Naturalist, 113, 791833.Google Scholar
Connor, E.F., McCoy, E.D. & Cosby, B.J. (1983) Model discrimination and expected slope values in species-area studies. American Naturalist, 122, 789796.Google Scholar
Coomes, D.A., Allen, R.B., Bentley, W.A., Burrows, L.E., Canham, C.D., Fagan, L., Forsyth, D.M., Gaxiola-Alcantar, A., Parfitt, R.L., Ruscoe, W.A., Wardle, D.A., Wilson, D.J. & Wright, E.F. (2005) The hare, the tortoise and the crocodile: the ecology of angiosperm dominance, conifer persistence and fern filtering. Journal of Ecology, 93, 918935.Google Scholar
Coomes, D.A. & Grubb, P.J. (2000) Impacts of root competition in forests and woodlands: a theoretical framework and review of experiments. Ecological Monographs, 70, 171207.Google Scholar
Corbin, J.D. & D’Antonio, C.M. (2004) Competition between native perennial and exotic annual grasses: implications for an historical invasion. Ecology, 85, 12731283.Google Scholar
Cornelissen, J.H.C. & Thompson, K. (1997) Functional leaf attributes predict litter decomposition rate in herbaceous plants. New Phytologist, 135, 109114.Google Scholar
Cornell, H.V. & Lawton, J.H. (1992) Species interactions, local and regional processes, and limits to the richness of ecological communities: a theoretical perspective. Journal of Animal Ecology, 61, 112.Google Scholar
Cornwell, W.K. & Ackerly, D.D. (2009) Community assembly and shifts in plant trait distributions across an environmental gradient in coastal California. Ecological Monographs, 79, 109126.Google Scholar
Correia, M., Montesinos, D., French, K. & Rodríguez-Echeverría, S. (2016) Evidence for enemy release and increased seed production and size for two invasive Australian acacias. Journal of Ecology, 104, 13911399.Google Scholar
Cowell, C.M., Hoalst-Pullen, N. & Jackson, M.T. (2010) The limited role of canopy gaps in the successional dynamics of a mature mixed Quercus forest remnant. Journal of Vegetation Science, 21, 201212.CrossRefGoogle Scholar
Cowles, H.C. (1901) The physiographic ecology of Chicago and vicinity; a study of the origin, development, and classification of plant societies. Botanical Gazette, 31, 145182.Google Scholar
Cowling, R.M., Mustart, P.J., Laurie, H. & Richards, M.B. (1994) Species diversity, functional diversity and functional redundancy in fynbos communities. South African Journal of Science, 90, 333337.Google Scholar
Cowling, R.M., Ojeda, F., Lamont, B.B., Rundel, P.W. & Lechmere-Oertel, R. (2005) Rainfall reliability, a neglected factor in explaining convergence and divergence of plant traits in fire-prone mediterranean-climate ecosystems. Global Ecology and Biogeography, 14, 509519.Google Scholar
Cowling, R.M. & Witkowski, E.T.F. (1994) Convergence and non-convergence of plant traits in climatically and edaphically matched sites in Mediterranean Australia and South Africa. Australian Journal of Ecology, 19, 220232.Google Scholar
Craine, J.M. (2005) Reconciling plant strategy theories of Grime and Tilman. Journal of Ecology, 93, 10411052.Google Scholar
Crawley, M.J. (1986) The structure of plant communities. In Plant Ecology (ed. Crawley, M.J.), pp. 150. Blackwell, Oxford.Google Scholar
Crawley, M.J. (1997) Life history and environment. In Plant Ecology (ed. Crawley, M.J.), pp. 73131. Blackwell, Oxford.Google Scholar
Crawley, M.J., Brown, S.L., Heard, M.S. & Edwards, G.R. (1999) Invasion-resistance in experimental grassland communities: species richness or species identity? Ecology Letters, 2, 140148.Google Scholar
Crepy, M.A. & Casal, J.J. (2015) Photoreceptor-mediated kin recognition in plants. New Phytologist, 205, 329338.Google Scholar
Crews, T.E., Kitayama, K., Fownes, J.H., Riley, R.H., Herbert, D.A., Mueller-Dombois, D. & Vitousek, P.M. (1995) Changes in soil phosphorus fractions and ecosystem dynamics across a long chronosequence in Hawaii. Ecology, 76, 14071424.Google Scholar
Csotonyi, J.T. & Addicott, J.F. (2004) Influence of trampling-induced microtopography on growth of the soil crust bryophyte Ceratodon purpureus in Jasper National Park. Canadian Journal of Botany, 82, 13821392.Google Scholar
Cuesta, B., Villar-Salvador, P., Puertolas, J., Benayas, J.M.R. & Michalet, R. (2010) Facilitation of Quercus ilex in Mediterranean shrubland is explained by both direct and indirect interactions mediated by herbs. Journal of Ecology, 98, 687696.Google Scholar
Culmsee, H., Leuschner, C., Moser, G. & Pitopang, R. (2010) Forest aboveground biomass along an elevational transect in Sulawesi, Indonesia, and the role of Fagaceae in tropical montane rain forests. Journal of Biogeography, 37, 960974.Google Scholar
Cupper, M.L., Drinnan, A.N. & Thomas, I. (2000) Holocene palaeoenvironments of salt lakes in the Darling Anabranch region, south-western New South Wales, Australia. Journal of Biogeography, 27, 10791094.Google Scholar
Curran, L.M. & Leighton, M. (2000) Vertebrate responses to spatiotemporal variation in seed production of mast-fruiting Dipterocarpaceae. Ecological Monographs, 70, 101128.Google Scholar
Curtis, J.T. & McIntosh, R.P. (1951) An upland forest continuum in the prairie-forest border region of Wisconsin. Ecology, 32, 476496.Google Scholar
Dahlin, A.S. & Stenberg, M. (2010) Transfer of N from red clover to perennial ryegrass in mixed stands under different cutting strategies. European Journal of Agronomy, 33, 149156.Google Scholar
Dale, M.B. (2002) Models, measures and messages: an essay on the role for induction. Community Ecology, 3, 191204.Google Scholar
Dale, M.P. & Causton, D.R. (1992) The ecophysiology of Veronica chamaedrys, V. montana and V. officinalis. I. Light quality and light quantity. Journal of Ecology, 80, 483492.Google Scholar
Dale, M.R.T. (1984) The contiguity of upslope and downslope boundaries of species in a zoned community. Oikos, 42, 9296.Google Scholar
Dale, M.R.T. (1985) Graph theoretical methods for comparing phytosociological structures. Vegetatio, 63, 7988.Google Scholar
Dallimore, W. (1917) Natural grafting of branches and roots. Bulletin of Miscellaneous Information, 1917, 303306.Google Scholar
Dalling, J.W., Swaine, M.D. & Garwood, N.C. (1998) Dispersal patterns and seed bank dynamics of pioneer trees in moist tropical forest. Ecology, 79, 564578.Google Scholar
Dansereau, P. (1964) Six problems in New Zealand vegetation. Bulletin of the Torrey Botanical Club, 91, 114140.Google Scholar
Darnell, R.M. (1970) Evolution and the ecosystem. American Zoologist, 10, 915.Google Scholar
Darwin, C. (1859) On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life, 1st edn. John Murray, London.Google Scholar
Davies, K.F., Cavender-Bares, J. & Deacon, N. (2011) Native communities determine the identity of exotic invaders even at scales at which communities are unsaturated. Diversity and Distributions, 17, 3542.Google Scholar
Davis, M.A., Grime, J.P. & Thompson, K. (2000) Fluctuating resources in plant communities: a general theory of invasibility. Journal of Ecology, 88, 528534.Google Scholar
Davis, M.B. (1981) Quaternary history and the stability of forest communities. Forest Succession: Concepts and Applications (eds West, D.C., Shugart, H.H. & Botkin, D.B.), pp. 132153. Springer, New York.Google Scholar
Dawson, T.E. (1993) Hydraulic lift and water use by plants – implications for water balance, performance and plant-plant interactions. Oecologia, 95, 565574.Google Scholar
de Bello, F., Price, J.N., Münkemüller, T., Liira, J., Zobel, M., Thuiller, W., Gerhold, P., Götzenberger, L., Lavergne, S., Lepš, J., Zobel, K. & Pärtel, M. (2012) Functional species pool framework to test for biotic effects on community assembly. Ecology, 93, 22632273.Google Scholar
de Brouwer, J.F.C. & Stal, L.J. (2002) Daily fluctuations of exopolymers in cultures of the benthic diatoms Cylindrotheca closterium and Nitschia sp. (Bacillariophyceae). Journal of Phycology, 38, 464472.Google Scholar
de Kroon, H. & Kalliola, R. (1995) Shoot dynamics of the giant grass Gynerium sagittatum in Peruvian Amazon floodplains, a clonal plant that does show self-thinning. Oecologia, 101, 124131.Google Scholar
de Kroon, H. & van Groenendael, J. (1997) The Ecology and Evolution of Clonal Plants. Backhuys, Leyden.Google Scholar
de Lange, P.J. & Stockley, J.E. (1987) The flora of the Lost World Cavern, Mangapu Caves System, Waitomo, Te Kuiti. Bulletin of the Wellington Botanical Society, 43, 36.Google Scholar
de Rooij-van der Goes, P.C.E.M. (1995) The role of plant-parasitic nematodes and soil-borne fungi in the decline of Ammophila arenaria (L.) Link. New Phytologist, 129, 661669.Google Scholar
del Moral, R. (1983) Competition as a control mechanism in subalpine meadows. American Journal of Botany, 70, 232245.Google Scholar
del Moral, R. (2007) Limits to convergence of vegetation during early primary succession. Journal of Vegetation Science, 18, 479488.Google Scholar
del Moral, R. (2009) Increasing deterministic control of primary succession on Mount St. Helens, Washington. Journal of Vegetation Science, 20, 11451154.Google Scholar
del Moral, R., Saura, J.M. & Emenegger, J.N. (2010) Primary succession trajectories on a barren plain, Mount St. Helens, Washington. Journal of Vegetation Science, 21, 857867.Google Scholar
DeLonge, M., D’Odorico, P. & Lawrence, D. (2008) Feedbacks between phosphorus deposition and canopy cover: the emergence of multiple stable states in tropical dry forests. Global Change Biology, 14, 154160.Google Scholar
DeMalach, N., Zaady, E., Weiner, J. & Kadmon, R. (2016) Size asymmetry of resource competition and the structure of plant communities. Journal of Ecology, 104, 899910.Google Scholar
Den Hartog, C. (1970) Sea Grasses of the World. North Holland, Amsterdam.Google Scholar
Deng, J.M., Zuo, W.Y., Wang, Z.Q., Fan, Z.X., Ji, M.F., Wang, G.X., Ran, J.Z., Zhao, C.M., Liu, J.Q., Niklas, K.J., Hammond, S.T. & Brown, J.H. (2012) Insights into plant size-density relationships from models and agricultural crops. Proceedings of the National Academy of Sciences of the U.S.A., 109, 86008605.Google Scholar
DeWoody, Y.D., Swihart, R.K., Craig, B.A. & Goheen, J.R. (2003) Diversity and stability in communities structured by asymmetric resource allocation. American Naturalist, 162, 514527.Google Scholar
Diamond, J.M. (1975) Assembly of species communities. In Ecology and Evolution of Communities (eds Cody, M.L. & Diamond, J.M.), pp. 342444. Harvard University Press, Cambridge, MA.Google Scholar
Díaz Barradas, M.C., García Novo, F., Collantes, M. & Zunzunegu, M. (2001) Vertical structure of wet grasslands under grazed and non-grazed conditions in Tierra del Fuego. Journal of Vegetation Science, 12, 385390.Google Scholar
Díaz, S., Cabido, M. & Casanoves, F. (1998) Plant functional traits and environmental filters at a regional scale. Journal of Vegetation Science, 9, 113122.Google Scholar
Diaz-Martin, Z., Swamy, V., Terborgh, J., Alvarez-Loayza, P. & Cornejo, F. (2014) Identifying keystone plant resources in an Amazonian forest using a long-term fruit-fall record. Journal of Tropical Ecology, 30, 291301.Google Scholar
Dice, L.R. (1952) Natural Communities. University of Michigan Press, Ann Arbor.Google Scholar
Dickie, I.A., Fukami, T., Wilkie, J.P., Allen, R.B. & Buchanan, P.K. (2012) Do assembly history effects attenuate from species to ecosystem properties? A field test with wood-inhabiting fungi. Ecology Letters, 15, 133141.Google Scholar
Diekmann, M. & Lawesson, J.E. (1999) Shifts in ecological behaviour of herbaceous forest species along a transect from northern Central to North Europe. Folia Geobotanica, 34, 127141.Google Scholar
Diez, J.M., Dickie, I., Edwards, G., Hulme, P.E., Sullivan, J.J. & Duncan, R.P. (2010) Negative soil feedbacks accumulate over time for non-native plant species. Ecology Letters, 13, 803809.Google Scholar
DiMichele, W.A., Phillips, T.L. & Nelson, W.J. (2002) Place vs. time and vegetational persistence: a comparison of four tropical mires from the Illinois Basin during the height of the Pennsylvanian ice age. International Journal of Coal Geology, 50, 4372.Google Scholar
Dixon, A.F.G. & Kundu, R. (1994) Ecology of host alternation in aphids. European Journal of Entomology, 91, 6370.Google Scholar
Dodd, A.P. (1940) The Biological Campaign Against Prickly-Pear. Commonwealth Prickly Pear Board, Brisbane.Google Scholar
Dodd, J., Heddle, E.M., Pate, J.S. & Dixon, K.W. (1984) Rooting patterns of sandplain plants and their functional significance. In Kongwan: Plant Life of the Sandplain (eds Pate, J.S. & Beard, J.S.), pp. 146177. University of Western Australia Press, Nedlands.Google Scholar
Dodd, M.E., Silvertown, J., McConway, K., Potts, J. & Crawley, M.J. (1995) Community stability: a 60-year record of trends and outbreaks in the occurrence of species in the Park Grass Experiment. Journal of Ecology, 83, 277285.Google Scholar
Dornbush, M.E. & Wilsey, B.J. (2010) Experimental manipulation of soil depth alters species richness and co-occurrence in restored tallgrass prairie. Journal of Ecology, 98, 117125.Google Scholar
Dostal, P. (2011) Plant Competitive Interactions and Invasiveness: Searching for the Effects of Phylogenetic Relatedness and Origin on Competition Intensity. American Naturalist, 177, 655667.Google Scholar
Drake, J.A. (1990) The mechanics of community assembly and succession. Journal of Theoretical Biology, 147, 213233.Google Scholar
Drake, J.A. (1991) Community-assembly mechanics and the structure of an experimental species ensemble. American Naturalist, 137, 126.Google Scholar
Drake, J.A., Zimmermann, C.R., Purucker, T. & Rojo, C. (1999) of the assembly trajectory. Assembly Rules and Restoration Ecology. Bridging the Gap Between Theory and Practice. (eds Temperton, V.M., Hobbs, R.J., Nuttle, T. & Halle, S.), p. 464. Island Press, Washington DC.Google 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, 479490.Google Scholar
Drude, O. (1885) Die Vertheilung und zusammensetzung östlicher pflanzengenossenschaften in der umgebung von Dresden. In Festschrift der Isis in Dresden, pp. 75107. Warnatz and Lehmann, Dresden.Google Scholar
Dublin, H.T. (1995) Vegetation dynamics in the Serengeti-Mara ecosystem: the role of elephants, fire and other factors. In Serengeti II: Dynamics, Management and Conservation of an Ecosystem (eds Sinclair, A.R.E. & Arcese, P.), pp. 7190. University of Chicago Press, Chicago, IL.Google Scholar
Dublin, H.T., Sinclair, A.R.E. & McGlade, J. (1990) Elephants and fire as causes of multiple stable states in the Serengeti-Mara woodlands. Journal of Animal Ecology, 59, 11471164.Google Scholar
Dudley, J.P. (1999) Seed dispersal of Acacia erioloba by African bush elephants in Hwange National Park, Zimbabwe. African Journal of Ecology, 37, 375385.Google Scholar
Dudley, S.A., Murphy, G.P. & File, A.L. (2013) Kin recognition and competition in plants. Functional Ecology, 27, 898906.Google Scholar
Dukes, J.S. (2001) Biodiversity and invasibility in grassland microcosms. Oecologia, 126, 563568.Google Scholar
Dukes, J.S. (2002) Species composition and diversity affect grassland susceptibility and response to invasion. Ecological Applications, 12, 602617.Google Scholar
Dulloo, M.E., Kell, S.P. & Jones, C.G. (2002) Impact and control of invasive alien species on small islands. International Forestry Review, 4, 277285.Google Scholar
Dumbrell, A.J., Nelson, M., Helgason, T., Dytham, C. & Fitter, A.H. (2010) Relative roles of niche and neutral processes in structuring a soil microbial community. ISME Journal, 4, 337345.Google Scholar
Dunbar, M.J. (1960) The evolution of stability in marine environments: natural selection at the level of the ecosystem. American Naturalist, 94, 129136.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, 403416.Google Scholar
Dunnett, N.P. & Grime, J.P. (1999) Competition as an amplifier of short-term vegetation responses to climate: an experimental test. Functional Ecology, 13, 388395.Google Scholar
Dunnett, N.P. & Willis, A.J. (2004) Monitoring of permanent plots in roadside verge vegetation at Bibury, Gloucestershire. Bulletin of the British Ecological Society, 35, 1821.Google Scholar
Dunnett, N.P., Willis, A.J., Hunt, R. & Grime, J.P. (1998) A 38 year study of relations between the weather and vegetation dynamics in a road verge near Bibury, Gloucestershire. Journal of Ecology, 86, 610623.Google Scholar
During, H.J. & Willems, J.H. (1986) The impoverishment of the bryophyte and lichen flora of the Dutch chalk grasslands in the thirty years 1953–1983. Biological Conservation, 36, 143158.Google Scholar
Dybzinski, R. & Tilman, D. (2007) Resource use patterns predict long-term outcomes of plant competition for nutrients and light. American Naturalist, 170, 305318.Google Scholar
Dyer, A.R., Fenech, A. & Rice, K.J. (2000) Accelerated seedling emergence in interspecific competitive neighbourhoods. Ecology Letters, 3, 523529.Google Scholar
Eckstein, R.L. & Donath, T.W. (2005) Interactions between litter and water availability affect seedling emergence in four familial pairs of floodplain species. Journal of Ecology, 93, 807816.Google Scholar
Edwards, G.R. & Crawley, M.J. (1999) Rodent seed predation and seedling recruitment in mesic grassland. Oecologia, 118, 288296.Google Scholar
Edwards, P.J. (1988) Effects of the fairy ring fungus Agaricus arvensis on nutrient availability in grassland. New Phytologist, 110, 377381.Google Scholar
Egler, F.E. (1954) Vegetation science concepts. I. Initial floristic composition, a factor in old-field vegetation development. Vegetatio, 4, 412417.Google Scholar
Egler, F.E. (1977) The Nature of Vegetation: Its Management and Mismanagement. Aton Forest Connecticut; in cooperation with the Connecticut Conservation Association, Bridgewater.Google Scholar
Ehlers, B.K., David, P., Damgaard, C.F. & Lenormand, T. (2016) Competitor relatedness, indirect soil effects and plant coexistence. Journal of Ecology, 104, 11261135.Google Scholar
Ehrenfeld, J.G., Kourtev, P. & Huang, W. (2001) Changes in soil functions following invasions of exotic understory plants in deciduous forests. Ecological Applications, 11, 12871300.Google Scholar
Ehrenfeld, J.G., Ravit, B. & Elgersma, K. (2005) Feedback in the plant-soil system. Annual Review of Environment and Resources, 30, 75115.Google Scholar
Ehrlén, J. (1996) Spatiotemporal variation in predispersal seed predation intensity. Oecologia, 108, 708713.Google Scholar
Ehrlén, J. & van Groenendael, J.M. (1998) The trade-off between dispersability and longevity – an important aspect of plant species diversity. Applied Vegetation Science, 1, 2936.Google Scholar
El Mehdawi, A.F., Cappa, J.J., Fakra, S.C., Self, J. & Pilon-Smits, E.A.H. (2012) Interactions of selenium hyperaccumulators and nonaccumulators during cocultivation on seleniferous or nonseleniferous soil – the importance of having good neighbors. New Phytologist, 194, 264277.Google Scholar
Eldridge, D.J. & Rath, D. (2002) Hip holes: Kangaroo (Macropus spp.) resting sites modify the physical and chemical environment of woodland soils. Austral Ecology, 27, 527536.Google Scholar
Elmendorf, S.C. & Harrison, S.P. (2011) Is plant community richness regulated over time? Contrasting results from experiments and long-term observations. Ecology, 92, 602609.Google Scholar
Elton, C.S. (1927) Animal Ecology. Sidgwick and Jackson, London.Google Scholar
Emborg, J., Christensen, M. & Heilmann-Clausen, J. (2000) The structural dynamics of Suserup Skov, a near-natural temperate deciduous forest in Denmark. Forest Ecology and Management, 126, 173189.Google Scholar
Emerman, S.H. & Dawson, T.E. (1996) Hydraulic lift and its influence on the water content of the rhizosphere: an example from sugar maple, Acer saccharum. Oecologia, 108, 273278.Google Scholar
Engelhardt, K.A.M. & Kadlec, J.A. (2001) Species traits, species richness and the resilience of wetlands after disturbance. Journal of Aquatic Plant Management, 39, 3639.Google Scholar
Enquist, B.J. & Niklas, K.J. (2001) Invariant scaling relations across tree-dominated communities. Nature, 410, 655660.Google Scholar
Enright, N.J. (1999) Litterfall dynamics in a mixed conifer-angiosperm forest in northern New Zealand. Journal of Biogeography, 26, 149157.Google Scholar
Eriksson, O. (2005) Game theory provides no explanation for seed size variation in grasslands. Oecologia, 144, 98105.Google Scholar
Eschenbach, C., Glauner, R., Kleine, M. & Kappen, L. (1998) Photosynthesis rates of selected tree species in lowland dipterocarp rainforest of Sabah, Malaysia. Trees Structure and Function, 12, 356365.Google Scholar
Esseen, P.-A. (1985) Litter fall of epiphytic macrolichens in two old Picea abies forests in Sweden. Canadian Journal of Botany, 63, 980987.Google Scholar
Evangelista, P.H., Kumar, S., Stohlgren, T.J., Jarnevich, C.S., Crall, A.W., Norman, J.B. & Barnett, D.T. (2008) Modelling invasion for a habitat generalist and a specialist plant species. Diversity and Distributions, 14, 808817.Google Scholar
Evans, M.N. & Barkham, J.P. (1992) Coppicing and natural disturbance in temperate woodlands, a review. In Ecology and Management of Coppiced Woodlands (ed. Buckley, G.P.), pp. 7998. Chapman and Hall, London.Google Scholar
Evans, R.C. & Turkington, R. (1988) Maintenance of morphological variation in a biotically patchy environment. New Phytologist, 109, 369376.Google Scholar
Eviner, V.T. & Chapin, F.S. (2003) Functional matrix: a conceptual framework for predicting multiple plant effects on ecosystem processes. Annual Review of Ecology Evolution and Systematics, 34, 455485.Google Scholar
Facelli, J.M. (1994) Multiple indirect effects of plant litter affect the establishment of woody seedlings in old fields. Ecology, 75, 17271735.Google Scholar
Facelli, J.M. & Temby, A.M. (2002) Multiple effects of shrubs on annual plant communities in arid lands of South Australia. Austral Ecology, 27, 422432.Google Scholar
Facelli, J.M., Williams, R., Fricker, S. & Ladd, B. (1999) Establishment and growth of seedlings of Eucalyptus obliqua: interactive effects of litter, water, and pathogens. Australian Journal of Ecology, 24, 484494.Google Scholar
Fahnestock, J.T., Povirk, K.L. & Welker, J.M. (2000) Ecological significance of litter redistribution by wind and snow in arctic landscapes. Ecography, 23, 623631.Google Scholar
Falcone, P., Keller, R., Le Tacon, F. & Oswald, H. (1986) Facteurs influencant la forme des feuillus en plantations. Revue Forestiere Francaise, 38, 315323.Google Scholar
Falik, O., Reides, P., Gersani, M. & Novoplansky, A. (2003) Self/non-self discrimination in roots. Journal of Ecology, 91, 525531.Google Scholar
Fargione, J. & Tilman, D. (2006) Plant species traits and capacity for resource reduction predict yield and abundance under competition in nitrogen- limited grassland. Functional Ecology, 20, 533540.Google Scholar
Fargione, J.E., Brown, C.S. & Tilman, D. (2003) Community assembly and invasion: an experimental test of neutral versus niche processes. Proceedings of the National Academy of Sciences of the U.S.A., 100, 89168920.Google Scholar
Fargione, J.E. & Tilman, D. (2005a) Diversity decreases invasion via both sampling and complementarity effects. Ecology Letters, 8, 604611.Google Scholar
Fargione, J.E. & Tilman, D. (2005b) Niche differentiation in phenology and rooting depth promote coexistence with a dominant C4 bunchgrass. Oecologia, 143, 598606.Google Scholar
Farley, R.A. & Fitter, A.H. (1999) Temporal and spatial variation in soil resources in a deciduous woodland. Journal of Ecology, 87, 688696.Google Scholar
Fastie, C.L. (1995) Causes and ecosystem consequences of multiple pathways of primary succession at Glacier Bay, Alaska. Ecology, 76, 18991916.Google Scholar
Federle, W., Maschwitz, U. & Hölldobler, B. (2002) Pruning of host plant neighbours as defence against enemy ant invasions: Crematogaster ant partners of Macaranga protected by “wax barriers” prune less than their congeners. Oecologia, 132, 264270.Google Scholar
Felker-Quinn, E., Schweitzer, J.A. & Bailey, J.K. (2013) Meta-analysis reveals evolution in invasive plant species but little support for Evolution of Increased Competitive Ability (EICA). Ecology and Evolution, 3, 739751.Google Scholar
Fellbaum, C.R., Mensah, J.A., Cloos, A.J., Strahan, G.E., Pfeffer, P.E., Kiers, E.T. & Bucking, H. (2014) Fungal nutrient allocation in common mycorrhizal networks is regulated by the carbon source strength of individual host plants. New Phytologist, 203, 646656.Google Scholar
Fiala, B., Maschwitz, U., Pong, T.Y. & Helbig, A.J. (1989) Studies of a South East Asian ant-plant association: protection of Macaranga trees by Crematogaster boreensis. Oecologia, 79, 463470.Google Scholar
Field, J.P., Breshears, D.D., Whicker, J.J. & Zou, C.B. (2012) Sediment capture by vegetation patches: implications for desertification and increased resource redistribution. Journal of Geophysical Research Biogeosciences, 117, 1033.Google Scholar
Fine, P.V.A. (2002) The invasibility of tropical forests by exotic plants. Journal of Tropical Ecology, 18, 687705.Google Scholar
Firestone, J.L. & Jasieniuk, M. (2013) Small population size limits reproduction in an invasive grass through both demography and genetics. Oecologia, 172, 109117.Google Scholar
Firn, R. (2004) Plant intelligence: an alternative point of view. Annals of Botany, 93, 345351.Google Scholar
Flory, S.L., Long, F.R. & Clay, K. (2011) Invasive Microstegium populations consistently outperform native range populations across diverse environments. Ecology, 92, 22482257.Google Scholar
Fonteyn, P.J. & Mahall, B.E. (1978) Competition among desert perennials. Nature, 275, 544545.Google Scholar
Fortin, M.-J. & Dale, M.R.T. (2005) Spatial Data Analysis: A Guide for Ecologists. Cambridge University Press, Cambridge, UK.Google Scholar
Fortunel, C., Paine, C.E.T., Fine, P.V.A., Kraft, N.J.B. & Baraloto, C. (2014) Environmental factors predict community functional composition in Amazonian forests. Journal of Ecology, 102, 145155.Google Scholar
Fowler, N.L. (1981) Competition and coexistence in a North Carolina grassland. II. The effects of the experimental removal of species. Journal of Ecology, 69, 843854.Google Scholar
Fowler, N.L. (1986a) The role of competition in plant communities in arid and semiarid regions. Annual Review of Ecology and Systematics, 17, 89110.Google Scholar
Fowler, N.L. (1986b) Microsite requirements for germination and establishment of tree grass species. American Midland Naturalist, 115, 131145.Google Scholar
Fowler, N.L. (1990) Disorderliness in plant communities: comparisons, causes, and consequences. Perspectives on Plant Competition (eds Grace, J.B. & Tilman, D.), pp. 291306. Academic Press, San Diego.Google Scholar
Franco, M. (1986) The influence of neighbours on the growth of modular organisms with an example from trees. Philosophical Transactions of the Royal Society of London B, 313, 209225.Google Scholar
Fraser, E.C., Lieffers, V.J. & Landhäusser, S.M. (2006) Carbohydrate transfer through root grafts to support shaded trees. Tree Physiology, 26, 10191023.Google Scholar
Freckleton, R.P. & Watkinson, A.R. (2001) Predicting competition coefficients for plant mixtures: reciprocity, transitivity and correlations with life-history traits. Ecology Letters, 348–357Google Scholar
Freckleton, R.P., Watkinson, A.R. & Rees, M. (2009) Measuring the importance of competition in plant communities. Journal of Ecology, 97, 379384.Google Scholar
Fridley, J.D. & Sax, D.F. (2014) The imbalance of nature: revisiting a Darwinian framework for invasion biology. Global Ecology and Biogeography, 23, 11571166.Google Scholar
Fuentes, E.R. (1976) Ecological convergence of lizard communities in Chile and California. Ecology, 57, 317.Google Scholar
Fuentes, M. (2004) Slight differences among individuals and the unified neutral theory of biodiversity. Theoretical Population Biology, 66, 199203.Google Scholar
Fujinuma, R., Bockheim, J. & Balster, N. (2005) Base-cation cycling by individual tree species in old-growth forests of Upper Michigan, USA. Biogeochemistry, 74, 357376.Google Scholar
Fukami, T. (2004) Community assembly along a species pool gradient: implications for multiple-scale patterns of species diversity. Population Ecology, 46, 137147.Google Scholar
Fukami, T., Bezemer, T.M., Mortimer, S.R. & van der Putten, W.H. (2005) Species divergence and trait convergence in experimental plant community assembly. Ecology Letters, 8, 12831290.Google Scholar
Fukami, T., Dickie, I.A., Wilkie, J.P., Paulus, B.C., Park, D., Roberts, A., Buchanan, P.K. & Allen, R.B. (2010) Assembly history dictates ecosystem functioning: evidence from wood decomposer communities. Ecology Letters, 13, 675684.Google Scholar
Fuller, J.L. (1998) Ecological impact of the mid-holocene hemlock decline in southern Ontario, Canada. Ecology, 79, 23372351.Google Scholar
Fuller, M.M. & Enquist, B.J. (2012) Accounting for spatial autocorrelation in null models of tree species association. Ecography, 35, 510518.Google Scholar
Furness, N.H. & Upadhyaya, M.K. (2002) Differential susceptibility of agricultural weeds to ultraviolet-B radiation. Canadian Journal of Plant Science, 82, 789796.Google Scholar
Futuyma, D.J. & Wasserman, S.S. (1980) Resource concentration and herbivory in oak forests. Science, 210, 920922.Google Scholar
Fynn, R.W.S., Morris, C.D. & Krikman, K.P. (2005) Plant strategies and trait trade-offs influence trends in competitive ability along gradients of soil fertility and disturbance. Journal of Ecology, 93, 384394.Google Scholar
García-Palacios, P., Shaw, E.A., Wall, D.H. & Hättenschwiler, S. (2016) Temporal dynamics of biotic and abiotic drivers of litter decomposition. Ecology Letters, 19, 554563.Google Scholar
Gartner, T.B. & Cardon, Z.G. (2004) Decomposition dynamics in mixed species leaf litter. Oikos, 104, 230246.Google Scholar
Gause, G.F. (1934) The Struggle for Existence. Williams and Wilkins, Baltimore.Google Scholar
Gentry, A.H. (1988) Changes in plant community diversity and floristic composition on environmental and geographical gradients. Annals of the Missouri Botanic Garden, 75, 134.Google Scholar
Genung, M.A., Bailey, J.K. & Schweitzer, J.A. (2012) Welcome to the neighbourhood: interspecific genotype by genotype interactions in Solidago influence above- and belowground biomass and associated communities. Ecology Letters, 15, 6573.Google Scholar
Germino, M.J. & Smith, W.K. (1999) Sky exposure, crown architecture, and low-temperature photoinhibition in conifer seedlings at alpine treeline. Plant Cell and Environment, 22, 407415.Google Scholar
Gigon, A. (1997) Fluctuations of dominance and the coexistence of plant species in limestone grasslands. Phytocoenologia, 27, 275287.Google Scholar
Gilbert, B. & Lechowicz, M.J. (2004) Neutrality, niches, and dispersal in a temperate forest understory. Proceedings of the National Academy of Sciences of the U.S.A., 101, 76517656.Google Scholar
Gilbert, G.S. (2002) Evolutionary ecology of plant diseases in natural ecosystems. Annual Review of Phytopathology, 40, 1343.Google Scholar
Gilbert, G.S. & Parker, I.M. (2010) Rapid evolution in a plant-pathogen interaction and the consequences for introduced host species. Evolutionary Applications, 3, 144156.Google Scholar
Gill, D.E., Chao, L., Perkins, S.L. & Wolf, J.B. (1995) Genetic mosaicism in plants and clonal animals. Annual Review of Ecology and Systematics, 26, 423444.Google Scholar
Gillett, J.B. (1962) Pest pressure, an underestimated factor in evolution. Systematics Association Publications No. 4, pp. 37–46. Oxford.Google Scholar
Gillman, L.N. (2016) Seedling mortality from litterfall increases with decreasing latitude. Ecology, 97, 530535.Google Scholar
Gilmore, R.G. & Snedaker, S.C. (1993) Mangrove forests. Biodiversity of the Southeastern United States (ed. Martin, W.H.), pp. 165199. Wiley, New York.Google Scholar
Gilpin, M.E. & Case, T.J. (1976) Multiple domains of attraction in competition communities. Nature, 261, 3942.Google Scholar
Gitay, H. & Wilson, J.B. (1995) Post-fire changes in community structure of tall tussock grasslands: a test of alternative models of succession. Journal of Ecology, 83, 775782.Google Scholar
Gleason, H.A. (1917) The structure and development of the plant association. Bulletin of the Torrey Botanical Club, 44, 463481.Google Scholar
Gleason, H.A. (1926) The individualistic concept of the plant association. Bulletin of the Torrey Botanical Club, 53, 726.Google Scholar
Gleason, H.A. (1927) Further views on the succession-concept. Ecology, 8, 299326.Google Scholar
Gleason, H.A. (1936) Is the synusia an association? Ecology, 17, 444451.Google Scholar
Gleason, H.A. (1939) The individualistic concept of the plant association. American Midland Naturalist, 21, 92110.Google Scholar
Goldberg, D.E. & Landa, K. (1991) Competitive effect and response – hierarchies and correlated traits in the early stages of competition. Journal of Ecology, 79, 10131030.Google Scholar
Goldberg, D.E., Rajaniemi, T.K., Gurevitch, J. & Stewart-Oaten, A. (1999) Empirical approaches to quantifying interaction intensity: competition and facilitation along productivity gradients. Ecology, 80, 11181131.Google Scholar
Goldberg, D.E. & Werner, P.A. (1983) Equivalence of competitors in plant communities: a null hypothesis and a field experimental approach. American Journal of Botany, 70, 10981104.Google Scholar
Goldblum, D. (1997) The effects of treefall gaps on understory vegetation in New York State. Journal of Vegetation Science, 8, 125132.Google Scholar
Golubski, A.J., Gross, K.L. & Mittelbach, G.G. (2010) Recycling-mediated facilitation and co-existence based on plant size. American Naturalist, 176, 588600.Google Scholar
Gómez‐Aparicio, L., Canham, C.D. & Martin, P.H. (2008) Neighbourhood models of the effects of the invasive Acer platanoides on tree seedling dynamics: linking impacts on communities and ecosystems. Journal of Ecology, 96, 7890.Google Scholar
Gómez-Aparicio, L., Gómez, J.M., Zamora, R. & Boettinger, J.L. (2005) Canopy vs. soil effects of shrubs facilitating tree seedlings in Mediterranean montane ecosystems. Journal of Vegetation Science, 16, 191198.Google Scholar
Goodall, D.W. (1954) Vegetational classification and vegetational continua. Angewandte Pflanzensoziologie, 1, 168182.Google Scholar
Goodall, D.W. (1963) The continuum and the individualistic association. Vegetatio, 11, 297316.Google Scholar
Goodall, D.W. (1966) The nature of the mixed community. Proceedings of the Ecological Society of Australia, 1, 8496.Google Scholar
Gotelli, N.J. & Graves, G.R. (1996) Null Models in Ecology. Smithsonian Institution Press, Washington.Google Scholar
Grace, J. (1983) Plant-Atmosphere Relationships. Chapman and Hall, London.Google Scholar
Grace, J.B. (1991) A clarification of the debate between Grime and Tilman. Functional Ecology, 5, 583587.Google Scholar
Grace, J.B. & Wetzel, R.G. (1981) Habitat partitioning and competitive displacement in cattails (Typha): experimental field studies. American Naturalist, 118, 463474.Google Scholar
Grady, K.C., Wood, T.E., Kolb, T.E., Hersch-Green, E., Shuster, S.M., Gehring, C.A., Hart, S.C., Allan, G.J. & Whitham, T.G. (2016) Local biotic adaptation of trees and shrubs to plant neighbors. Oikos, 583–593.Google Scholar
Graham, R.W. & Grimm, E.C. (1990) Effects of global climate change on the patterns of terrestrial biological communities. Trends in Ecology and Evolution, 5, 289292.Google Scholar
Grant, K., Kreyling, J., Heilmeier, H., Beierkuhnlein, C. & Jentsch, A. (2014) Extreme weather events and plant plant interactions: shifts between competition and facilitation among grassland species in the face of drought and heavy rainfall. Ecological Research, 29, 9911001.Google Scholar
Gravel, D., Canham, C.D., Beaudet, M. & Messier, C. (2010) Shade tolerance, canopy gaps and mechanisms of coexistence of forest trees. Oikos, 119, 475484.Google Scholar
Graves, J.D., Press, M.C., Smith, S. & Stewart, G.R. (1992) The carbon canopy economy of the association between cowpea and the parasitic angiosperm Striga gesnerioides. Plant Cell and Environment, 15, 283288.Google Scholar
Grayston, S.J., Wang, S., Campbell, C.D. & Edwards, A.C. (1998) Selective influence of plant species on microbial diversity in the rhizosphere. Soil Biology and Biochemistry, 30, 369378.Google Scholar
Green, P.T., O’Dowd, D.J. & Lake, P.S. (2008) Recruitment dynamics in a rainforest seedling community: context-independent impact of a keystone consumer. Oecologia, 156, 373385.Google Scholar
Greig-Smith, P. (1952) Ecological observations on degraded and secondary forest in Trinidad, British West Indies. Journal of Ecology, 40, 283330.Google Scholar
Greig-Smith, P. (1983) Quantitative Plant Ecology, 3rd edn. Blackwell, Oxford.Google Scholar
Grigg, A.H. & Mulligan, D.R. (1999) Litterfall from two eucalypt woodlands in central Queensland. Australian Journal of Ecology, 24, 662664.Google Scholar
Griggs, R.F. (1914) Observations on the behavior of some species at the edges of their ranges. Bulletin of the Torrey Botanical Club, 41, 2549.Google Scholar
Grime, J.P. (1973) Control of species density in herbaceous vegetation. Journal of Environmental Management, 1, 151167.Google Scholar
Grime, J.P. (1974) Vegetation classification by reference to strategies. Nature, 250, 2631.Google Scholar
Grime, J.P. (1977) Evidence for the existence of three primary strategies in plants and its relevance to ecological and evolutionary theory. American Naturalist, 111, 11691194.Google Scholar
Grime, J.P. (1979) Plant Strategies and Vegetation Processes. Wiley, Chichester.Google Scholar
Grime, J.P. (1988) The C-S-R model of primary plant strategies – origins, implications and tests. In Plant Evolutionary Biology (eds Gottlieb, L.D. & Jain, S.K.), pp. 371393. Chapman and Hall, London.Google Scholar
Grime, J.P. (2001) Plant Strategies, Vegetation Processes, and Ecosystem Properties, 2nd edn. Wiley, Chichester.Google Scholar
Grime, J.P. (2006) Trait convergence and trait divergence in herbaceous plant communities: mechanisms and consequences. Journal of Vegetation Science, 17, 255260.Google Scholar
Grime, J.P., Hodgson, J.G. & Hunt, R. (2007) Comparative Plant Ecology: A Functional Approach to Common British Species, 2nd edn. Castlepoint Press, Dalbeattie.Google Scholar
Grime, J.P. & Hunt, R. (1975) Relative growth rate: its range and adaptive significance in a local flora. Journal of Ecology, 63, 393422.Google Scholar
Grime, J.P., Hunt, R. & Krzanowski, W.J. (1987) Evolutionary physiological ecology of plants. Evolutionary Physiological Ecology (ed. Calow, P.), pp. 105125. Cambridge University Press, Cambridge, UK.Google Scholar
Grime, J.P., Thompson, K., Hunt, R., Hodgson, J.G., Cornelissen, J.H.C., Rorison, I.H., Hendry, G.A.F., Ashenden, T.W., Askew, A.P., Band, S.R., Booth, R.E., Bossard, C.C., Campbell, B.D., Cooper, J.E.L., Davison, A.W., Gupta, P.L., Hall, W., Hand, D.W., Hannah, M.A., Hillier, S.H., Hodkinson, D.J., Jalili, A., Liu, Z., Mackey, J.M.L., Matthews, N., Mowforth, M.A., Neal, A.M., Reader, R.J., Reiling, K., Ross-Fraser, W., Spencer, R.E., Sutton, F., Tasker, D.E., Thorpe, P.C. & Whitehouse, J. (1997) Integrated screening validates primary axes of specialisation in plants. Oikos, 79, 259281.Google Scholar
Grime, J.P., Willis, A.J., Hunt, R. & Dunnett, N.P. (1994) Climate-vegetation relationships in the Bibury road verge experiments. In Long-Term Experiments in Agricultural and Ecological Sciences (eds Leigh, R.A. & Johnston, A.E.), pp. 271285. CAB International, Wallingford.Google Scholar
Grinnell, J. (1904) The origin and distribution of the chestnut-backed chickadee. Auk, 21, 375377.Google Scholar
Groom, M.J. (1998) Allee effects limit population viability of an annual plant. American Naturalist, 151, 487496.Google Scholar
Gross, C.L., Nelson, P.A., Haddadchi, A. & Fatemi, M. (2012) Somatic mutations contribute to genotypic diversity in sterile and fertile populations of the threatened shrub, Grevillea rhizomatosa (Proteaceae). Annals of Botany, 109, 331342.Google Scholar
Gross, K., Cardinale, B.J., Fox, J.W., Gonzalez, A., Loreau, M., Polley, H.W., Reich, P.B. & van Ruijven, J. (2014) Species richness and the temporal stability of biomass production: a new analysis of recent biodiversity experiments. American Naturalist, 183, 112.Google Scholar
Grostal, P. & O’Dowd, D.J. (1994) Plants, mites and mutualism – leaf domatia and the abundance and reproduction of mites on Viburnum tinus (Caprifoliaceae). Oecologia, 97, 308315.Google Scholar
Grover, J.P. (1994) Assembly rules for communities of nutrient-limited plants and specialist herbivores. American Naturalist, 143, 258282.Google Scholar
Grubb, P.J. (1982) Control of relative abundance in roadside Arrhenatheretum: results of a long-term garden experiment. Journal of Ecology, 70, 845861.Google Scholar
Grubb, P.J. (1985) Plant populations and vegetation in relation to habitat, disturbance and competition: problems of generalization. In The Population Structure of Vegetation (ed. White, J.), pp. 595621. Junk, Dordrecht.Google Scholar
Grubb, P.J. (1987) Some generalizing ideas about colonisation and succession in green plants and fungi. In Colonization, Succession and Stability (eds Gray, A.J., Crawley, M.J. & Edwards, P.J.), pp. 81102. Blackwell, Oxford.Google Scholar
Gruntman, M. & Novoplansky, A. (2004) Physiologically mediated self/non-self discrimination in roots. Proceedings of the National Academy of Sciences of the U.S.A., 101, 38633867.Google Scholar
Gulmon, S., Rundel, P.W., Ehleringer, J.R. & Mooney, H.A. (1979) Spatial relationships and competition in a Chilean desert cactus. Oecologia, 44, 4043.Google Scholar
Gulmon, S.L. (1977) A comparative study of the grassland of California and Chile. Flora, 166, 261278.Google Scholar
Gundale, M.J., Kardol, P., Nilsson, M.C., Nilsson, U., Lucas, R.W. & Wardle, D.A. (2014) Interactions with soil biota shift from negative to positive when a tree species is moved outside its native range. New Phytologist, 202, 415421.Google Scholar
Gurevitch, J. & Unnasch, R.S. (1989) Experimental removal of a dominant species at two levels of soil fertility. Canadian Journal of Botany, 67, 34703477.Google Scholar
Haase, P. (1990) Environmental and floristic gradients in Westland, New Zealand, and the discontinuous distribution of Nothofagus. New Zealand Journal of Botany, 28, 2540.Google Scholar
Hacker, S.D. & Bertness, M.D. (1999) Experimental evidence for factors maintaining plant species diversity in a New England salt marsh. Ecology, 80, 20642073.Google Scholar
Hairston, N.G. (1981) An experimental test of a guilds: salamander competition. Ecology, 62, 6572.Google Scholar
Halkett, J.C. (1991) The Native Forests of New Zealand. GP Publications, Wellington, New Zealand.Google Scholar
Hall, S.J. & Raffaelli, D.G. (1993) Food webs: theory and reality. Advances in Ecological Research, 24, 187239.Google Scholar
Hallett, J.G. (1982) Habitat selection and the community matrix of a desert small-mammal fauna. Ecology, 63, 14001410.Google Scholar
Handa, I.T., Harmsen, R. & Jefferies, R.L. (2002) Patterns of vegetation change and the recovery potential of degraded areas in a coastal marsh system of the Hudson Bay lowlands. Journal of Ecology, 90, 8699.Google Scholar
Hanley, M.E. & Sykes, R.J. (2009) Impacts of seedling herbivory on plant competition and implications for species coexistence. Annals of Botany, 103, 13471353.Google Scholar
Hansen, G.J.A., Ives, A.R., Vander Zanden, M.J. & Carpenter, S.R. (2013) Are rapid transitions between invasive and native species caused by alternative stable states, and does it matter? Ecology, 94, 22072219.Google Scholar
Hanya, G. (2005) Comparisons of dispersal success between the species fruiting prior to and those at the peak of migrant frugivore abundance. Plant Ecology, 181, 167177.Google Scholar
Hara, T. & Srutek, M. (1995) Shoot growth and mortality patterns of Urtica dioica, a clonal forb. Annals of Botany, 76, 235243.Google Scholar
Harberd, D.J. (1962) Some observations on natural clones in Festuca ovina. New Phytologist, 61, 85100.Google Scholar
Hargeby, A., Blindow, I. & Hansson, L.-A. (2004) Shifts between clear and turbid states in a shallow lake: multi-causal stress from climate, nutrients and biotic interactions. Archiv für Hydrobiologie, 161, 433454.Google Scholar
Harper, J.L. (1967) A darwinian approach to plant ecology. Journal of Ecology, 55, 247270.Google Scholar
Harper, J.L. (1977) Population Biology of Plants. Academic Press, London.Google Scholar
Harper, J.L. & 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, 681698.Google Scholar
Harpole, W.S. & Tilman, D. (2006) Non-neutral patterns of species abundance in grassland communities. Ecology Letters, 9, 1523.Google Scholar
Harries, J.H. & Norrington-Davies, J. (1977) Competition studies in diploid and tetraploid varieties of Lolium perenne. II. The inhibition of germination. Journal of Agricultural Science, 88, 411415.Google Scholar
Harrison, K.A. & Bardgett, R.D. (2010) Influence of plant species and soil conditions on plant-soil feedback in mixed grassland communities. Journal of Ecology, 98, 384395.Google Scholar
Hart, S.P., Schreiber, S.J. & Levine, J.M. (2016) How variation between individuals affects species coexistence. Ecology Letters, 19, 825838.Google Scholar
Hautier, Y., Hector, A., Vojtech, E., Purves, D. & Turnbull, L.A. (2010) Modelling the growth of parasitic plants. Journal of Ecology, 98, 857866.Google Scholar
Hautier, Y., Niklaus, P.A. & Hector, A. (2009) Competition for light causes plant biodiversity after eutrophication. Science, 324, 636638.Google Scholar
Haydon, D. (1994) Pivotal assumptions determining the relationship between stability and complexity – an analytical synthesis of the stability-complexity debate. American Naturalist, 144, 1429.Google Scholar
He, Q. & Bertness, M.D. (2014) Extreme stresses, niches, and positive species interactions along stress gradients. Ecology, 95, 14371443.Google Scholar
He, Q., Bertness, M.D. & Altieri, A.H. (2013) Global shifts towards positive species interactions with increasing environmental stress. Ecology Letters, 16, 695706.Google Scholar
He, W.M., Feng, Y.L., Ridenour, W.M., Thelen, G.C., Pollock, J.L., Diaconu, A. & Callaway, R.M. (2009) Novel weapons and invasion: biogeographic differences in the competitive effects of Centaurea maculosa and its root exudates (±)-catechin. Oecologia, 159, 803815.Google Scholar
He, X.H., Critchley, C., Ng, H. & Bledsoe, C. (2004) Reciprocal N (15NH4+ or 15NO3) transfer between non-N2-fixing Eucalyptus maculata and N2-fixing Casuarina cunninghamiana linked by the ectomycorrhizal fungus Pisolithus sp. New Phytologist, 163, 629640.Google Scholar
Heads, M.J. (1989) Integrating earth and life sciences in New Zealand natural history: the parallel arcs model. New Zealand Journal of Zoology, 16, 549585.Google Scholar
Hector, A. (2006) Overyielding and stable species coexistence. New Phytologist, 172, 13.Google Scholar
Hector, A., Hautier, Y., Saner, P., Wacker, L., Bagchi, R., Joshi, J., Schererlorenzen, M., Spehn, E.M., Bazeleywhite, E., Weilenmann, M., Caldeira, M.C., Dimitrakopoulos, P.G., Finn, J.A., Hussdanell, K., Jumpponen, A., Mulder, C.P.H., Palmborg, C., Pereira, J.S., Siamantziouras, A.S.D., Terry, A.C., Troumbis, A.Y., Schmid, B. & Loreau, M. (2010) General stabilizing effects of plant diversity on grassland productivity through population asynchrony and overyielding. Ecology, 91, 22132220.Google Scholar
Hegarty, E.E. (1991) Vine-host interactions. In The Biology of Vines (eds Putz, F.E. & Mooney, H.A.), pp. 357375. Cambridge University Press, Cambridge, UK.Google Scholar
Hellström, G.B. & Lubke, R.A. (1993) Recent changes to a climbing-falling dune system on the Robberg Peninsula, Southern Cape Coast, South Africa. Journal of Coastal Research, 9, 647653.Google Scholar
Herben, T. & Goldberg, D.E. (2014) Community assembly by limiting similarity vs. competitive hierarchies: testing the consequences of dispersion of individual traits. Journal of Ecology, 102, 156166.Google Scholar
Herben, T., Krahulec, F., Hadincová, V. & Kovářová, M. (1990) Fine scale dynamics in a mountain grassland. In Spatial Processes in Plant Communities (eds Krahulec, F., Agnew, A.D.Q., Agnew, S. & Willems, J.H.), pp. 173184. SPB Academic Publishing, The Hague.Google Scholar
Herben, T., Krahulec, F., Hadincová, V., Pecháčková, S. & Wildová, R. (2003) Year-to-year variation in plant competition in a mountain grassland. Journal of Ecology, 91, 103113.Google Scholar
Herbert, D.A., Fownes, J.H. & Vitousek, P.M. (1999) Hurricane damage to a Hawaiian forest: nutrient supply rate affects resistance and resilience. Ecology, 80, 908920.Google Scholar
Herrera, C.M., Jordano, P., Lopez-Soria, L. & Amat, J.A. (1994) Recruitment of a mast-fruiting, bird-dispersed tree-bridging frugivore activity and seedling establishment. Ecological Monographs, 64, 315344.Google Scholar
Hill, S.B. & Kotanen, P.M. (2009) Evidence that phylogenetically novel non-indigenous plants experience less herbivory. Oecologia, 161, 581590.Google Scholar
Himanen, S.J., Blande, J.D., Klemola, T., Pulkkinen, J., Heijari, J. & Holopainen, J.K. (2010) Birch (Betula spp.) leaves adsorb and re-release volatiles specific to neighbouring plants – a mechanism for associational herbivore resistance? New Phytologist, 186, 722732.Google Scholar
Hirose, T. & Werger, M.J.A. (1995) Canopy structure and photon flux partitioning among species in a herbaceous plant community. Ecology, 76, 466474.Google Scholar
Hirota, M., Holmgren, M., VanNes, E.H. & Scheffer, M. (2011) Global resilience of tropical forest and savanna to critical transitions. Science, 334, 232235.Google Scholar
Hoagland, B.W. & Collins, S.L. (1997) Gradient models, gradient analysis, and hierarchical structure in plant communities. Oikos, 78, 2330.Google Scholar
Hochberg, M.E., Menaut, J.C. & Gignoux, J. (1994) Influences of tree biology and fire in the spatial structure of the West African savannah. Journal of Ecology, 82, 217226.Google Scholar
Hodgkinson, K.C. & Baas Becking, H.G. (1977) Effect of defoliation on root growth of some arid zone perennial plants. Australian Journal of Agricultural Research, 29, 3142.Google Scholar
Hodgson, J.G., Wilson, P.J., Hunt, R., Grime, J.P. & Thompson, K. (1999) Allocating C-S-R plant functional types: a soft approach to a hard problem. Oikos, 85, 282294.Google Scholar
Hoffmann, W.A., Adasme, R., Haridasan, M., de Carvalho, M.T., Geiger, E.L., Pereira, M.A.B., Gotsch, S.G. & Franco, A.C. (2009) Tree topkill, not mortality, governs the dynamics of savanna-forest boundaries under frequent fire in central Brazil. Ecology, 90, 13261337.Google Scholar
Hofgaard, A. (1993) Structure and regeneration patterns in a virgin Picea abies forest in northern Sweden. Journal of Vegetation Science, 4, 601608.Google Scholar
Holdredge, C. & Bertness, M.D. (2011) Litter legacy increases the competitive advantage of invasive Phragmites australis in New England wetlands. Biological Invasions, 13, 423433.Google Scholar
Holt, G. & Chesson, P. (2014) Variation in moisture duration as a driver of coexistence by the storage effect in desert annual plants. Theoretical Population Biology, 92, 3650.Google Scholar
Holzapfel, C. & Mahall, B.E. (1999) Bidirectional facilitation and interference between shrubs and annuals in the Mojave desert. Ecology, 80, 17471761.Google Scholar
Hooper, D.U. & Dukes, J.S. (2004) Overyielding among plant functional groups in a long-term experiment. Ecology Letters, 7, 95105.Google Scholar
Hooper, D.U. & Dukes, J.S. (2010) Functional composition controls invasion success in a California serpentine grassland. Journal of Ecology, 98, 764777.Google Scholar
Houlahan, J.E., Currie, D.J., Cottenie, K., Cumming, G.S., Ernest, S.K., Findlay, C.S., Fuhlendorf, S.D., Gaedke, U., Legendre, P., Magnuson, J.J., McArdle, B.H., Muldavin, E.H., Noble, D., Russell, R., Stevens, R.D., Willis, T.J., Woiwod, I.P. & Wondzell, S.M. (2007) Compensatory dynamics are rare in natural ecological communities. Proceedings of the National Academy of Sciences of the U.S.A., 104, 32733277.Google Scholar
Howard, K.S.C., Eldridge, D.J. & Soliveres, S. (2012) Positive effects of shrubs on plant species diversity do not change along a gradient in grazing pressure in an arid shrubland. Basic and Applied Ecology, 13, 159168.Google Scholar
Hubbell, S.P. (2001) The Unified Neutral Theory of Biodiversity and Biogeography. Princeton University Press, Princeton, NJ.Google Scholar
Hubbell, S.P. (2005) Neutral theory in community ecology and the hypothesis of functional equivalence. Functional Ecology, 19, 166172.Google Scholar
Hubbell, S.P. & Foster, R.B. (1986) Biology, chance, and history and the structure of tropical rain forest tree communities. In Community Ecology (eds Diamond, J.M. & Case, T.J.), pp. 314329. Harper and Row, New York.Google Scholar
Hughes, J.B. & Roughgarden, J. (1998) Aggregate community properties and the strength of species’ interactions. Proceedings of the National Academy of Sciences of the U.S.A., 95, 68376842.Google Scholar
Hughes, J.B. & Roughgarden, J. (2000) Species diversity and biomass stability. American Naturalist, 155, 618627.Google Scholar
Hunt, R. & Hope-Simpson, J.F. (1990) Growth of Pyrola rotundfolia ssp. maritima in relation to shade. New Phytologist, 114, 129137.Google Scholar
Hunter, A.F. & Aarssen, L.W. (1988) Plants helping plants. BioScience, 38, 3440.Google Scholar
Hutchinson, G.E. (1941) Ecological aspects of succession in natural populations. American Naturalist, 75, 406418.Google Scholar
Hutchinson, G.E. (1951) Copepodology for the ornithologist. Ecology, 32, 571577.Google Scholar
Hutchinson, G.E. (1957) Concluding remarks. Cold Spring Harbour Symposium on Quantitative Biology, 22, 415427.Google Scholar
Hutchinson, G.E. (1959) Homage to Santa Rosalia, or why are there so many kinds of animals. American Naturalist, 93, 145159.Google Scholar
Hutchinson, G.E. (1961) The paradox of the plankton. American Naturalist, 95, 137145.Google Scholar
Hyatt, L.A., Rosenberg, M.S., Howard, T.G., Bole, G., Fang, W., Anastasia, J., Brown, K., Grella, R., Hinman, K., Kurdziel, J. & Gurevitch, J. (2003) The distance dependence prediction of the Janzen-Connell hypothesis: a meta-analysis. Oikos, 103, 590602.Google Scholar
Inderjit, , Dakshini, K.M.M. & Einhellig, F.A. (1994) Allelopathy: Organisms, Processes, and Applications. American Chemical Society, Washington.Google Scholar
Inderjit, & Mallik, A.U. (1997) Effects of Ledum groenlandicum amendments on black spruce seedling growth. Plant Ecology, 133, 2936.Google Scholar
Innes, L., Hobbs, P.J. & Bardgett, R.D. (2004) The impacts of individual plant species on rhizosphere microbial communities in soils of different fertility. Biology and Fertility of Soils, 40, 713.Google Scholar
Inouye, R.S., Allison, T.D. & Johnson, N.C. (1987) Old-Field succession on a Minnesota sand plain. Ecology, 68, 1226.Google Scholar
Isbell, F.I., Polley, H.W. & Wilsey, B.J. (2009) Biodiversity, productivity and the temporal stability of productivity: patterns and processes. Ecology Letters, 12, 443451.Google Scholar
Ives, A.R. & Hughes, J.B. (2002) General relationships between species diversity and stability in competitive systems. American Naturalist, 159, 388395.Google Scholar
Jabot, F. & Chave, J. (2011) Analyzing tropical forest tree species abundance distributions using a nonneutral model and through approximate Bayesian inference. American Naturalist, 178, 3747.Google Scholar
Jackson, S.T. & Williams, J.W. (2004) Modern analogs in quaternary paleoecology: here today, gone yesterday, gone tomorrow? Annual Review of Earth and Planetary Science, 32, 495537.Google Scholar
Jacobsen, A.L., Esler, K.J., Pratt, R.B. & Ewers, F.W. (2009) Water stress tolerance of shrubs in mediterranean-type climate regions: convergence of fynbos and succulent karoo communities with California shrub communities. American Journal of Botany, 96, 14451453.Google Scholar
Janse, J.H. (1997) A model of nutrient dynamics in shallow lakes in relation to multiple stable states. Hydrobiologia, 342, 18.Google Scholar
Jogesh, T., Carpenter, D. & Cappuccino, N. (2008) Herbivory on invasive exotic plants and their non-invasive relatives. Biological Invasions, 10, 797804.Google Scholar
Johnson, D.J., Beaulieu, W.T., Bever, J.D. & Clay, K. (2012) Conspecific negative density dependence and forest diversity. Science, 336, 904907.Google Scholar
Johnson, E.A. (1992) Fire and Vegetation Dynamics: Studies from the North American Boreal Forest. Cambridge University Press, Cambridge, UK.Google Scholar
Jokinen, K. (1991) Yield and competition in barley variety mixtures. Journal of Agricultural Science in Finland, 63, 287305.Google Scholar
Jones, C.G., Lawton, J.H. & Shachak, M. (1994) Organisms as ecosystem engineers. Oikos, 69, 373386.Google Scholar
Jones, J.I. & Sayer, C.D. (2003) Does the fish-invertebrate-periphyton cascade precipitate plant loss in shallow lakes? Ecology, 84, 21552167.Google Scholar
Jones, M.G. (1933) Grassland management and its effect on the sward. Journal of the Royal Agricultural Society of England, 94, 2141.Google Scholar
Jonzén, N., Nolet, B.A., Santamaría, J. & Svensson, M. (2002) Seasonal herbivory and mortality compensation in a swan-pondweed system. Ecological Modelling, 147, 209219.Google Scholar
Jordán, F., Takács-Sánta, A. & Molnár, I. (1999) A reliability theoretical quest for keystones. Oikos, 86, 453462.Google Scholar
Joshi, J. & Vrieling, K. (2005) The enemy release and EICA hypothesis revisited: incorporating the fundamental difference between specialist and generalist herbivores. Ecology Letters, 8, 704714.Google Scholar
Jump, A.S. & Woodward, E.I. (2003) Seed production and population density decline approaching the range-edge of Cirsium species. New Phytologist, 160, 349358.Google Scholar
Kadmon, R. & Shmida, A. (1990) Competition in a variable environment: an experimental study in a desert annual plant population. Israel Journal of Botany, 39, 403412.Google Scholar
Kaneko, N. & Salamanca, E.F. (1999) Mixed leaf litter effects on decomposition rates and soil microarthropod communities in an oak-pine stand in Japan. Ecological Research, 14, 131138.Google Scholar
Karban, R., Shiojiri, K., Ishizaki, S., Wetzel, W.C. & Evans, R.Y. (2013) Kin recognition affects plant communication and defence. Proceedings of the Royal Society of London B, 280, 20123062.Google Scholar
Kazmierczak, E., van der Maarel, E. & Noest, V. (1995) Plant communities in kettle-holes in central Poland: chance occurrence of species? Journal of Vegetation Science, 6, 863874.Google Scholar
Keddy, P.A. (1983) Shoreline vegetation in Axe Lake, Ontario: effects of exposure on vegetation patterns. Ecology, 64, 331344.Google Scholar
Keddy, P.A. (1992) Assembly and response rules: two goals for predictive community ecology. Journal of Vegetation Science, 3, 157164.Google Scholar
Keddy, P.A. (2007) Plants and Vegetation. Origins. Processes. Consequences. Cambridge University Press, Cambridge, UK.Google Scholar
Keddy, P.A., Fraser, L.H. & Wisheu, I.C. (1998) A comparative approach to examine competitive response of 48 wetland plant species. Journal of Vegetation Science, 9, 777786.Google Scholar
Keeley, J.E. & Fotheringham, C.J. (1997) Trace gas emissions and smoke-induced seed germination. Science, 276, 12481250.Google Scholar
Keith, S.A., Newton, A.C., Morecroft, M.D., Golicher, D.J. & Bullock, J.M. (2011) Plant metacommunity structure remains unchanged during biodiversity loss in English woodlands. Oikos, 120, 302310.Google Scholar
Keller, E. & Steffen, K.L. (1995) Increased chilling tolerance and altered carbon metabolism in tomato leaves following application of mechanical stress. Physiologia Plantarum, 93, 519525.Google Scholar
Kellman, M. & Tackaberry, R. (1993) Disturbance and tree species coexistence in tropical riparian forest fragments. Global Ecology and Biogeography Letters, 3, 19.Google Scholar
Kelly, B.J., Wilson, J.B. & Mark, A.F. (1989) Causes of the species/area relation: a study of islands in Lake Manapouri, New Zealand. Journal of Ecology, 71, 10211028.Google Scholar
Kelly, C.K., Blundell, S.J., Bowler, M.G., Fox, G.A., Harvey, P.H., Lomas, M.R. & Woodward, F.I. (2011) The statistical mechanics of community assembly and species distribution. New Phytologist, 191, 819827.Google Scholar
Kelly, C.K., Venable, D.L. & Zimmerer, K. (1988) Host specialization in Cuscuta costaricensis: an assessment of host use relative to host availability. Oikos, 53, 315320.Google Scholar
Kenkel, N.C., McIlraith, A.L., Burchill, C.A. & Jones, G. (1991) Competition and the response of three plant species to a salinity gradient. Canadian Journal of Botany, 69, 24972502.Google Scholar
Kielland, K. (1994) Amino acid absorption by arctic plants: implications for plant nutrition and nitrogen cycling. Ecology, 75, 23732383.Google Scholar
Kikvidze, Z., Khetsuriani, L. & Kikodze, D. (2005) Small-scale guild proportions and niche complementarity in a Caucasian subalpine hay meadow. Journal of Vegetation Science, 16, 565570.Google Scholar
Kikvidze, Z., Khetsuriani, L., Kikodze, D. & Callaway, R.M. (2006) Seasonal shifts in competition and facilitation in subalpine plant communities of the central Caucasus. Journal of Vegetation Science, 17, 7782.Google Scholar
King, T.J. (1977) The plant ecology of ant-hills in calcareous grasslands. I. Patterns of species in relation to ant-hills in southern England. Journal of Ecology, 65, 235256.Google Scholar
King, W.M. & Wilson, J. (2006) Differentiation between native and exotic plant species from a dry grassland: fundamental responses to resource availability, and growth rates. Austral Ecology, 31, 9961004.Google Scholar
Kissel, R., Wilson, J.B., Bannister, P. & Mark, A.F. (1987) Water relations of some native and exotic shrubs of New Zealand. New Phytologist, 107, 2937.Google Scholar
Kitajima, K. (1994) Relative importance of photosynthetic traits and allocation patterns as correlates of seedling shade tolerance of 13 tropical trees. Oecologia, 98, 419428.Google Scholar
Klekowski, E.J. & Godfrey, P.J. (1989) Ageing and mutation in plants. Nature, 340, 389391.Google Scholar
Kleyer, M. (2002) Validation of plant functional types across two contrasting landscapes. Journal of Vegetation Science, 13, 167178.Google Scholar
Klimeš, L., Jongepier, J.W. & Jongepierová, I. (1995) Variability in species richness and guild structure in two species-rich grasslands. Folia Geobotanica et Phytotaxonomica, 30, 243253.Google Scholar
Klinger, L.F. (1996) The myth of the classic hydrosere model of bog succession. Arctic and Alpine Research, 28, 19.Google Scholar
Kloeppel, B.D. & Abrams, M.D. (1995) Ecophysiological attributes of the native Acer saccharum and the exotic Acer platanoides in urban oak forests in Pennsylvania, USA. Tree Physiology, 15, 739746.Google Scholar
Knight, B., Zhao, F.J., McGrath, S.P. & Shen, Z.G. (1997) Zinc and cadmium uptake by the hyperaccumulator Thlaspi caerulescens in contaminated soils and its effects on the concentration and chemical speciation of metals in soil solution. Plant and Soil, 197, 7178.Google Scholar
Knoop, W.T. & Walker, B.H. (1985) Interactions of woody and herbaceous vegetation in a southern African savanna. Journal of Ecology, 73, 235253.Google Scholar
Kohls, S.J., Baker, D.D., van Kessel, C. & Dawson, J.O. (2003) An assessment of soil enrichment by actinorhizal N2 fixation using delta-15N values in a chronosequence of deglaciation at Glacier Bay, Alaska. Plant and Soil, 254, 1117.Google Scholar
Kohyama, T. & Takada, T. (2009) The stratification theory for plant coexistence promoted by one-sided competition. Journal of Ecology, 97, 463471.Google Scholar
Konno, M., Iwamoto, S. & Seiwa, K. (2011) Specialization of a fungal pathogen on host tree species in a cross-inoculation experiment. Journal of Ecology, 99, 13941401.Google Scholar
Kooijman, A.M. & Bakker, C. (1994) The acidification capacity of wetland bryophytes as influenced by simulated clean and polluted rain. Aquatic Botany, 48, 133144.Google Scholar
Korhola, A. (1996) Initiation of a sloping mire complex in southwestern Finland: autogenic versus allogenic controls. Ecoscience, 3, 216222.Google Scholar
Körner, C. (2003a) Alpine Plant Life: Functional Plant Ecology of High Mountain Ecosystems, 2nd edn. Springer, Berlin.Google Scholar
Körner, C. (2003b) Limitation and stress: always or never? Journal of Vegetation Science, 14, 141143.Google Scholar
Körner, C., Stöcklin, J., Reuther-Thiébaud, L. & Pelaez-Riedl, S. (2008) Small differences in arrival time influence composition and productivity of plant communities. New Phytologist, 177, 698705.Google Scholar
Körner, S. (2001) Development of submerged macrophytes in shallow Lake Muggelsee (Berlin, Germany) before and after its switch to the phytoplankton-dominated state. Archiv für Hydrobiologie, 152, 395409.Google Scholar
Korstian, C.F. & Stickel, P.W. (1927) The natural replacement of blight-killed chestnut in the hardwood forests of the Northeast. Journal of Agricultural Research, 34, 631648.Google Scholar
Kraft, N.J.B. & Ackerly, D.D. (2010) Functional trait and phylogenetic tests of community assembly across spatial scales in an Amazonian forest. Ecological Monographs, 80, 401422.Google Scholar
Kraft, N.J.B., Valencia, R. & Ackerly, D.D. (2008) Functional traits and niche-based tree community assembly in an Amazonian forest. Science, 322, 580582.Google Scholar
Král, K., McMahon, S.M., Janik, D., Adam, D. & Vrska, T. (2014) Patch mosaic of developmental stages in central European natural forests along vegetation gradient. Forest Ecology and Management, 330, 1728.Google Scholar
Kraus, M., Fusseder, A. & Beck, E. (1987) Development and replenishment of the P-depletion zone around the primary root of maize during the vegetation period. Plant and Soil, 101, 247255.Google Scholar
Krebs, C.J. (1978) Ecology, the Experimental Analysis of Distribution and Abundance. Harper and Row, New York.Google Scholar
Kuang, J.J. & Chesson, P. (2010) Interacting coexistence mechanisms in annual plant communities: frequency-dependent predation and the storage effect. Theoretical Population Biology, 77, 5670.Google Scholar
Kubota, Y. (2000) Spatial dynamics of regeneration in a conifer/broad-leaved forest in northern Japan. Journal of Vegetation Science, 11, 633640.Google Scholar
Kubota, Y., Konno, Y. & Hiura, T. (1994) Stand structure and growth patterns of understorey trees in a coniferous forest, Taisetsuzan National Park, northern Japan. Ecological Research, 9, 333341.Google Scholar
Kucbel, S., Jaloviar, P., Saniga, M., Vencurik, J. & Klimaš, V. (2010) Canopy gaps in an old-growth fir-beech forest remnant of Western Carpathians. European Journal of Forest Research, 129, 249259.Google Scholar
Kueffer, C., Kronauer, L. & Edwards, P.J. (2009) Wider spectrum of fruit traits in invasive than native floras may increase the vulnerability of oceanic islands to plant invasions. Oikos, 118, 13271234.Google Scholar
Kuhry, P., Nicholson, B.J., Gignac, L.D. & Bayley, S.E. (1993) Development of Sphagnum dominated peatlands in boreal continental Canada. Canadian Journal of Botany, 71, 1022.Google Scholar
Kuiters, A.T. & Slim, P.A. (2003) Tree colonisation of abandoned arable land after 27 years of horse-grazing: the role of bramble as a facilitator of oak wood regeneration. Forest Ecology and Management, 181, 239251.Google Scholar
Kunin, W.E. (1998) Biodiversity at the edge: a test of the importance of spatial “mass effects” in the Rothamsted Park Grass experiments. Proceedings of the National Academy of Sciences of the U.S.A., 95, 207212.Google Scholar
La Peyre, M.K.G., Grace, J.B., Hahn, E. & Mendelssohn, I.A. (2001) The importance of competition in regulating plant species abundance along a salinity gradient. Ecology, 82, 6269.Google Scholar
Laird, R.A. & Schamp, B.S. (2006) Competitive intransitivity promotes species coexistence. American Naturalist, 168, 182193.Google Scholar
Laland, K.N., Odling-Smee, F.J. & Feldman, M.W. (1996) The evolutionary consequences of niche construction: a theoretical investigation using two-locus theory. Journal of Evolutionary Biology, 9, 293316.Google Scholar
Laland, K.N., Odling-Smee, F.J. & Feldman, M.W. (1999) Evolutionary consequences of niche construction and their implications for ecology. Proceedings of the National Academy of Sciences of the U.S.A., 96, 1024210247.Google Scholar
Laliberte, E., Turner, B.L., Costes, T., Pearse, S.J., Wyrwoll, K.H., Zemunik, G. & Lambers, H. (2012) Experimental assessment of nutrient limitation along a 2-million-year dune chronosequence in the south-western Australia biodiversity hotspot. Journal of Ecology, 100, 631642.Google Scholar
Lambers, J.H.R., Harpole, W.S., Tilman, D. & Knops, J. (2004) Mechanisms responsible for the positive diversity-productivity relationship in Minnesota grasslands. Ecology Letters, 7, 661668.Google Scholar
Lambers, J.H.R., Yelenik, S.G., Colman, B.P. & Levine, J.M. (2010) California annual grass invaders: the drivers or passengers of change? Journal of Ecology, 98, 11471156.Google Scholar
Lamont, B.B., Klinkhamer, P.G.L. & Witkowski, E.T.F. (1993) Population fragmentation may reduce fertility to zero in Banksia goodii – a demonstration of the Allee effect. Oecologia, 94, 446450.Google Scholar
Lamprey, H. (1964) Estimation of the large mammal densities, biomass and energy exchange in the Tarangire Game Reserve and the Maasai Steppe in Tanzania. East African Wildlife Journal, 2, 146.Google Scholar
Landres, P.B. & MacMahon, J.A. (1980) Guilds and community organization: analysis of an oak woodland avifauna in Sonora, Mexico. Auk, 97, 351365.Google Scholar
Lankau, R.A. (2012) Coevolution between invasive and native plants driven by chemical competition and soil biota. Proceedings of the National Academy of Sciences of the U.S.A., 109, 1124011245.Google Scholar
Lankau, R.A. & Strauss, S.Y. (2007) Mutual feedbacks maintain both genetic and species diversity in a plant community. Science, 317, 15611563.Google Scholar
Larson, J.E. & Funk, J.L. (2016) Regeneration: an overlooked aspect of trait-based plant community assembly models. Journal of Ecology, 104, 12841298.Google Scholar
Law, R. & Morton, R.D. (1996) Permanence and the assembly of ecological communities. Ecology, 77, 762775.Google Scholar
Lawton, J.H. (1991) Ecology as she is done, and could be done. Oikos, 61, 289290.Google Scholar
Lawton, J.H. (1999) Are there general laws in ecology? Oikos, 84, 177192.Google Scholar
Leake, J.R. (1994) The biology of myco-heterotrophic (‘saprophytic’) plants. New Phytologist, 127, 171216.Google Scholar
Lee, M.R., Flory, S.L. & Phillips, R.P. (2012) Positive feedbacks to growth of an invasive grass through alteration of nitrogen cycling. Oecologia, 170, 457465.Google Scholar
Lee, W.G., Wilson, J.B., Meurk, C.D. & Kennedy, P.C. (1991) Invasion of the subantarctic Auckland Islands, New Zealand, by the asterad tree Olearia lyallii and its interaction with a resident myrtaceous tree Metrosideros umbellata. Journal of Biogeography, 18, 493508.Google Scholar
Leeflang, L. (1999) Are stoloniferous plants able to avoid neighbours in response to low R:FR ratios in reflected light? Plant Ecology, 141, 5965.Google Scholar
Leger, E.A. & Rice, K.J. (2003) Invasive Californian poppies (Eschscholzia californica Cham.) grow larger than native individuals under reduced competition. Ecology Letters, 6, 257264.Google Scholar
Lehman, C.L. & Tilman, D. (2000) Biodiversity, stability, and productivity in competitive communities. American Naturalist, 156, 534552.Google Scholar
Leibold, M.A. & Mikkelson, G.M. (2002) Coherence, species turnover, and boundary clumping: elements of meta-community structure. Oikos, 97, 237250.Google Scholar
Lenton, T.M. & van Oijen, M. (2002) Gaia as a complex adaptive system. Philosophical Transactions of the Royal Society of London B, 357, 683695.Google Scholar
Lepik, A., Abakumova, M., Zobel, K. & Semchenko, M. (2012) Kin recognition is density-dependent and uncommon among temperate grassland plants. Functional Ecology, 26, 12141220.Google Scholar
Leschen, R.A.B., Buckley, T.R., Harman, H.M. & Shulmeister, J. (2008) Determining the origin and age of the Westland beech (Nothofagus) gap, New Zealand, using fungus beetle genetics. Molecular Ecology, 17, 12561276.Google Scholar
Levin, S.A. (1974) Dispersion and population interactions. American Naturalist, 108, 207228.Google Scholar
Levine, J.M. (2001) Local interactions, dispersal, and native and exotic plant diversity along a California stream. Oikos, 95, 397408.Google Scholar
Levins, R. & Culver, D. (1971) Regional coexistence of species and competition between rare species. Proceedings of the National Academy of Sciences of the U.S.A., 68, 12461248.Google Scholar
Levins, R. & Lewontin, R. (1985) The Dialectical Biologist. Harvard University Press, Cambridge, MA.Google Scholar
Lewis, M.W., Leslie, M.E. & Liljegren, S.J. (2006) Plant separation: 50 ways to leave your mother. Current Opinion in Plant Biology, 9, 5965.Google Scholar
Lewontin, R.C. (1969) The meaning of stability. Diversity and Stability in Ecological Systems, Brookhaven Symposia in Biology No. 22, pp. 1323. Brookhaven National Laboratory, New York.Google Scholar
Li, L., Li, S.-M., Sun, J.-H., Zhou, L.-L., Bao, X.-G., Zhang, H.-G. & Zhang, F.-S. (2007) Diversity enhances agricultural productivity via rhizosphere phosphorus facilitation on phosphorus-deficient soils. Proceedings of the National Academy of Sciences of the U.S.A., 104, 1119211196.Google Scholar
Liancourt, P., Callaway, R.M. & Michalet, R. (2005) Stress tolerance and competitive-response ability determine the outcome of biotic interactions. Ecology, 86, 16111618.Google Scholar
Lichtenthaler, H.K. (1996) Vegetation stress: an introduction to the stress concept in plants. Journal of Plant Physiology, 148, 414.Google Scholar
Lieberman, M., Peralta, R. & Hartshorn, G.S. (1995) Canopy closure and the distribution of tropical forest tree species at La-Selva, Costa-Rica. Journal of Tropical Ecology, 11, 161178.Google Scholar
Lind, E.M. & Morrison, M.E.S. (1974) East African Vegetation. Longman, London.Google Scholar
Lippmaa, T. (1939) The unistratal concept of plant communities (the unions). American Midland Naturalist, 21, 111145.Google Scholar
Liu, X.B., Liang, M.X., Etienne, R.S., Wang, Y.F., Staehelin, C. & Yu, S.X. (2012) Experimental evidence for a phylogenetic Janzen-Connell effect in a subtropical forest. Ecology Letters, 15, 111118.Google Scholar
Lloyd, K.M., Lee, W.G., Fenner, M. & Loughnan, A.E. (2003) Vegetation change after artificial disturbance in an alpine Chionochloa pallens grassland in New Zealand. New Zealand Journal of Ecology, 27, 3136.Google Scholar
Lloyd, K.M., McQueen, A.A.M., Lee, B.J., Wilson, R.C.B., Walker, S. & Wilson, J.B. (2000) Evidence on ecotone concepts from switch, environmental and anthropogenic ecotones. Journal of Vegetation Science, 11, 903910.Google Scholar
Lockwood, J.A. & Lockwood, D.R. (1993) Catastrophe theory: a unified paradigm for rangeland ecosystem dynamics. Journal of Range Management, 46, 282288.Google Scholar
Long, J.N. & Smith, F.W. (1992) Volume increment in Pinus contorta var. latifolia: the influence of stand development and crown dynamics. Forest Ecology and Management, 53, 5364.Google Scholar
Long, Z.T., Carson, W.P. & Peterson, C.J. (1998) Can disturbance create refugia from herbivores? An example with hemlock regeneration on treefall mounds. Journal of the Torrey Botanical Society, 125, 165168.Google Scholar
Long, Z.T., Mohler, C.L. & Carson, W.P. (2003) Extending the resource concentration hypothesis to plant communities: effects of litter and herbivores. Ecology, 84, 652665.Google Scholar
Longworth, J.B., Mesquita, R.C., Bentos, T.V., Moreira, M.P., Massoca, P.E. & Williamson, G.B. (2014) Shifts in dominance and species assemblages over two decades in alternative successions in central amazonia. Biotropica, 46, 529537.Google Scholar
Lord, J.M., Wilson, J.B., Steel, J.B. & Anderson, B.J. (2000) Community reassembly: a test using limestone grassland in New Zealand. Ecology Letters, 3, 213218.Google Scholar
Loreau, M. & Behera, N. (1999) Phenotypic diversity and stability of ecosystem processes. Theoretical Population Biology, 56, 2947.Google Scholar
Loreau, M. & Hector, A. (2001) Partitioning selection and complementarity in biodiversity experiments. Nature, 412, 7276.Google Scholar
Loreau, M., Mouquet, N. & Gonzalez, A. (2003) Biodiversity as spatial insurance in heterogeneous landscapes. Proceedings of the National Academy of Sciences of the U.S.A., 100, 1276512770.Google Scholar
Loreau, M., Sapijanskas, J., Isbell, F. & Hector, A. (2012) Niche and fitness differences relate the maintenance of diversity to ecosystem function: comment. Ecology, 93, 14821487.Google Scholar
Lortie, C.J., Brooker, R.W., Choler, P., Kikvidze, Z., Michalet, R., Pugnaire, F.I. & Callaway, R.M. (2004) Rethinking plant community ecology. Oikos, 107, 433438.Google Scholar
Lovelock, J.E. (1979) Gaia: A New Look at Life on Earth. Oxford University Press, Oxford.Google Scholar
Luh, H.-K. & Pimm, S.L. (1993) The assembly of ecological communities: a minimalist approach. Journal of Animal Ecology, 62, 749765.Google Scholar
Luo, W.B., Xie, Y.H., Chen, X.S., Li, F. & Qin, X.Y. (2010) Competition and facilitation in three marsh plants in response to a water-level gradient. Wetlands, 30, 525530.Google Scholar
Lüscher, A., Connolly, J. & Jacquard, P. (1992) Neighbour specificity between Lolium perenne and Trifolium repens from a natural pasture. Oecologia, 91, 404409.Google Scholar
Lusk, C. & Ogden, J. (1992) Age structure and dynamics of a podocarp broadleaf forest in Tongariro-National-Park, New-Zealand. Journal of Ecology, 80, 379393.Google Scholar
Lusk, C.H., Chazdon, R.L., Hofmann, G. & Memmott, J. (2006) A bounded null model explains juvenile tree community structure along light availability gradients in a temperate rain forest. Oikos, 112, 131137.Google Scholar
Lutz, H.J. (1943) Injuries to trees caused by Celastrus and Vitis. Bulletin of the Torrey Botanical Club, 70, 436439.Google Scholar
Lyons, K.G. & Schwartz, M.W. (2001) Rare species loss alters ecosystem function – invasion resistance. Ecology Letters, 4, 358365.Google Scholar
Maalouf, J.P., Le Bagousse-Pinguet, Y., Marchand, L., Touzard, B. & Michalet, R. (2012) The interplay of stress and mowing disturbance for the intensity and importance of plant interactions in dry calcareous grasslands. Annals of Botany, 110, 821828.Google Scholar
MacArthur, R. & Levins, R. (1967) The limiting similarity, convergence, and divergence of coexisting species. American Naturalist, 101, 377385.Google Scholar
MacArthur, R.H. (1957) On the relative abundance of bird species. Proceedings of the National Academy of Sciences of the U.S.A., 43, 293295.Google Scholar
MacArthur, R.H. (1972) Geographical Ecology: Patterns in the Distribution of Species. Harper and Row, New York.Google Scholar
MacArthur, R.H. & Wilson, E.O. (1963) An equilibrium theory of insular zoogeography. Evolution, 17, 373387.Google Scholar
MacArthur, R.H. & Wilson, E.O. (1967) The Theory of Island Biogeography. Princeton University Press, Princeton, NJ.Google Scholar
MacDonald, I.A.W. & Cooper, J. (1995) Insular Lessons for Global Biodiversity Conservation with Particular Reference to Alien Invasions. In Islands: Biological Diversity and Ecosystem Function (eds Vitousek, P., Loope, L.L. & Andersen, H.), pp. 189202. Springer-Verlag, Berlin.Google Scholar
MacNally, R. (2000) Coexistence of a locally undifferentiated foraging guild: avian snatchers in a southeastern Australian forest. Austral Ecology, 25, 6982.Google Scholar
Madon, O. & Médail, F. (1997) The ecological significance of annuals on a Mediterranean grassland (Mt Ventoux, France). Plant Ecology, 129, 189199.Google Scholar
Madritch, M., Donaldson, J.R. & Lindroth, R.L. (2006) Genetic identity of Populus tremuloides litter influences decomposition and nutrient release in a mixed forest stand. Ecosystems, 9, 528537.Google Scholar
Maestre, F.T. & Cortina, J. (2004) Do positive interactions increase with abiotic stress? A test from a semi-arid steppe. Biology Letters, 271, 331333.Google Scholar
Maestre, F.T., Valladares, F. & Reynolds, J.F. (2005) Is the change of plant–plant interactions with abiotic stress predictable? A meta-analysis of field results in arid environments. Journal of Ecology, 93, 748757.Google Scholar
Mahall, B.E. & Callaway, R.M. (1991) Root communication among desert shrubs. Proceedings of the National Academy of Sciences of the U.S.A., 88, 874876.Google Scholar
Mahall, B.E. & Callaway, R.M. (1992) Root communication mechanisms and intracommunity distributions of two Mojave Desert shrubs. Ecology, 73, 21452151.Google Scholar
Makita, A. (1996) Density regulation during the regeneration of two monocarpic bamboos: self-thinning or intra clonal regulation? Journal of Vegetation Science, 7, 281288.Google Scholar
Mallik, A.U. (2003) Conifer regeneration problems in boreal and temperate forests with ericaceous understory: role of disturbance, seedbed limitation, and keystone species change. Critical Reviews in Plant Sciences, 22, 341366.Google Scholar
Malloch, A.J.C. (1997) Influence of salt spray on dry coastal vegetation. In Dry Coastal Ecosystems: General Aspects (ed. van der Maarel, E.), pp. 411420. Elsevier, Amsterdam.Google Scholar
Malmer, N., Albinsson, C., Svensson, B.M. & Wallén, B. (2003) Interferences between Sphagnum and vascular plants: effects on plant community structure and peat formation. Oikos, 100, 469482.Google Scholar
Malmer, N., Svensson, B.M. & Wallén, B. (1994) Interactions between Sphagnum mosses and field layer vascular plants in the development of peat-forming systems. Folia Geobotanica et Phytotaxonomica, 29, 483496.Google Scholar
Malmstrom, C.M., McCullough, A.J., Johnson, H.A., Newton, L.A. & Borer, E.T. (2005) Invasive annual grasses indirectly increase virus incidence in California native perennial bunchgrasses. Oecologia, 145, 153164.Google Scholar
Manders, P.T. & Richardson, D.M. (1992) Colonisation of Cape Fynbos communities by forest species. Forest Ecology and Management, 48, 277293.Google Scholar
Manders, P.T. & Smith, R.E. (1992) Effects of watering regime on growth and competitive ability of nursery grown Cape Fynbos and forest plants. South African Journal of Botany, 58, 188194.Google Scholar
Mangan, S.A., Herre, E.A. & Bever, J.D. (2010) Specificity between Neotropical tree seedlings and their fungal mutualists leads to plant-soil feedback. Ecology, 91, 25942603.Google Scholar
Mangla, S., Inderjit, & Callaway, R.M. (2008) Exotic invasive plant accumulates native soil pathogens which inhibit native plants. Journal of Ecology, 96, 5867.Google Scholar
Mark, A.F. & Wilson, J.B. (2005) Tempo and mode of vegetation dynamics over 50 years in a New Zealand alpine cushion/tussock community. Journal of Vegetation Science, 16, 227236.Google Scholar
Marks, C.O. & Muller-Landau, H.C. (2007) Comment on “From plant traits to plant communities: a statistical mechanistic approach to biodiversity”. Science, 316, 1425.Google Scholar
Maron, J.L., Marler, M., Klironomos, J.N. & Cleveland, C.C. (2011) Soil fungal pathogens and the relationship between plant diversity and productivity. Ecology Letters, 14, 3641.Google Scholar
Marquard, E., Weigelt, A., Temperton, V.M., Roscher, C., Schumacher, J., Buchmann, N., Fischer, M., Weisser, W.W. & Schmid, B. (2009) Plant species richness and functional composition drive overyielding in a six-year grassland experiment. Ecology, 90, 32903302.Google Scholar
Marr, J.W. (1977) The development and movement of tree islands near the upper limit of tree growth in the southern Rocky Mountains. Ecology, 58, 11591164.Google Scholar
Marrs, R.H. & Hicks, M.J. (1986) Study of vegetation change at Lakenheath Warren: a re-examination of A.S. Watt’s theories of bracken dynamics in relation to succession and management. Journal of Applied Ecology, 23, 10291046.Google Scholar
Marshall, C. (1996) Sectoriality and physiological organisation in herbaceous plants: an overview. Vegetatio, 127, 916.Google Scholar
Martin, M.M. & Harding, J. (1981) Evidence for the evolution of competition between two species of annual plants. Evolution, 35, 975987.Google Scholar
Martin, P.H., Canham, C.D. & Kobe, R.K. (2010) Divergence from the growth-survival trade-off and extreme high growth rates drive patterns of exotic tree invasions in closed-canopy forests. Journal of Ecology, 98, 778789.Google Scholar
Martin, P.H. & Marks, P.L. (2006) Intact forests provide only weak resistance to a shade-tolerant invasive Norway maple (Acer platanoides L.). Journal of Ecology, 94, 10701079.Google Scholar
Mason, N.W.H. & Wilson, J.B. (2006) Mechanisms of species coexistence in a lawn community: mutual corroboration between two independent assembly rules. Community Ecology, 7, 109116.Google Scholar
Mason, N.W.H., Wilson, J.B. & Steel, J.B. (2007) Are alternative stable states more likely in high stress environments? Logic and available evidence do not support Didham et al. 2005. Oikos, 116, 353357.Google Scholar
Matista, A.A. & Silk, W.K. (1997) An electronic device for continuous, in vivo measurement of forces exerted by twining vines. American Journal of Botany, 84, 11641168.Google Scholar
Matsui, T., Dougherty, N.J., Loughnan, A.E., Swaney, J.K., Laurence, B.L., Lloyd, K.M. & Wilson, J.B. (2002) Local texture convergence within three communities in Fiordland, New Zealand. New Zealand Journal of Ecology, 26, 1522.Google Scholar
May, F., Giladi, I., Ristow, M., Ziv, Y. & Jeltsch, F. (2013) Metacommunity, mainland-island system or island communities? Assessing the regional dynamics of plant communities in a fragmented landscape. Ecography, 36, 842853.Google Scholar
May, R.M. (1972) Will a large complex system be stable? Nature, 238, 413414.Google Scholar
May, R.M. (1973) Stability and Complexity in Model Ecosystems. Princeton University Press, Princeton, NJ.Google Scholar
May, R.M. (1975) Patterns of species abundance and diversity. In Ecology and Evolution of Communities (eds Cody, M.L. & Diamond, J.M.), pp. 81120. Harvard University Press, Cambridge, MA.Google Scholar
Maynard Smith, J. & Price, G.R. (1973) The logic of animal conflict. Nature, 246, 1518.Google Scholar
Mazza, C.A., Gimenez, P.I., Kantolic, A.G. & Ballare, C.L. (2013) Beneficial effects of solar UV-B radiation on soybean yield mediated by reduced insect herbivory under field conditions. Physiologia Plantarum, 147, 307315.Google Scholar
McCarthy-Neumann, S. & Kobe, R.K. (2010) Conspecific plant-soil feedbacks reduce survivorship and growth of tropical tree seedlings. Journal of Ecology, 98, 396407.Google Scholar
McCune, B. & Allen, T.F.H. (1985) Will similar forests develop on similar sites? Canadian Journal of Botany, 63, 367376.Google Scholar
McDonald, A.K., Kinucan, R.J. & Loomis, L.E. (2009) Ecohydrological interactions within banded vegetation in the northeastern Chihuahuan Desert, USA. Ecohydrology, 2, 6671.Google Scholar
McGinley, M.A., Dhillion, S.S. & Neumann, J.C. (1994) Environmental heterogeneity and seedling establishment – ant-plant-microbe interactions. Functional Ecology, 8, 607615.Google Scholar
McInerny, G.J. & Etienne, R.S. (2012) Ditch the niche – is the niche a useful concept in ecology or species distribution modelling? Journal of Biogeography, 39, 20962102.Google Scholar
McKee, A., La Roi, G. & Frankilin, J.F. (1982) Structure, composition and reproductive behaviour of terrace forests, South Fork Hoh River, Olympic National Park. In Ecological Research in National Parks of the Pacific Northwest (eds Starkey, E.E., Franklin, J.F. & Matthews, J.W.), pp. 2228. National Park Service Cooperative Park Studies Unit, Corvallis.Google Scholar
McNeilly, T. & Roose, M.L. (1996) Co-adaptation between neighbours? A case study with Lolium perenne genotypes. Euphytica, 92, 121128.Google Scholar
McPherson, G.R., Wright, H.A. & Wester, D.B. (1988) Patterns of shrub invasion in semiarid Texas grasslands. American Midland Naturalist, 120, 391397.Google Scholar
McSorley, R. & Dickson, D.W. (1995) Effect of tropical rotation crops on Meloidogyne incognita and other plant-parasitic nematodes. Journal of Nematology, 27, 535544.Google Scholar
Meier, C.L., Keyserling, K. & Bowman, W.D. (2009) Fine root inputs to soil reduce growth of a neighbouring plant via distinct mechanisms dependent on root carbon chemistry. Journal of Ecology, 97, 941949.Google Scholar
Meijer, W. (1976) A note on Podostemon ceratophyllum Michx. as an indicator of clean streams in and around the Appalachian Mountains. Castanea, 41, 319324.Google Scholar
Meiners, S.J., Cadenasso, M.L. & Pickett, S.T.A. (2007) Succession on the Piedmont of New Jersey and its implications for ecological restoration. In Old fields: Dynamics and Restoration of Abandoned Farmland (eds Cramer, V.A. & Hobbs, R.J.), pp. 145161. Island Press, Washington DC.Google Scholar
Mensah, R.K. & Khan, M. (1997) Use of Medicago sativa (L.) interplantings/trap crops in the management of the green mirid, Creontiades dilutus (Stål) in commercial cotton in Australia. International Journal of Pest Management, 43, 197202.Google Scholar
Mergeay, J., De Meester, L., Eggermont, H. & Verschuren, D. (2011) Priority effects and species sorting in a long paleoecological record of repeated community assembly through time. Ecology, 92, 22672275.Google Scholar
Merritt, D.M., Nilsson, C. & Jansson, R. (2010) Consequences of propagule dispersal and river fragmentation for riparian plant community diversity and turnover. Ecological Monographs, 80, 609626.Google Scholar
Michalet, R., Le Bagousse-Pinguet, Y., Maalouf, J.P. & Lortie, C.J. (2014) Two alternatives to the stress-gradient hypothesis at the edge of life: the collapse of facilitation and the switch from facilitation to competition. Journal of Vegetation Science, 25, 609613.Google Scholar
Middendorp, R.S., Vlam, M., Rebel, K.T., Baker, P.J., Bunyavejchewin, S. & Zuidema, P.A. (2013) Disturbance history of a seasonal tropical forest in Western Thailand: a spatial dendroecological analysis. Biotropica, 45, 578586.Google Scholar
Midgley, J.J., Cameron, M.C. & Bond, W.J. (1995) Gap characteristics and replacement patterns in the Knysna Forest, South Africa. Journal of Vegetation Science, 6, 2936.Google Scholar
Midgley, J.J., Cowling, R.M., Hendricks, H., Desmet, P.G., Esler, K. & Rundel, P.W. (1997) Population ecology of tree succulents (Aloe and Pachypodium) in the arid western Cape: decline of keystone species. Biodiversity and Conservation, 6, 869876.Google Scholar
Mihail, J.D., Alexander, H.M. & Taylor, S.J. (1998) Interactions between root-infecting fungi and plant density in an annual legume, Kummerowia stipulacea. Journal of Ecology, 86, 739748.Google Scholar
Miller, A.D. & Chesson, P. (2009) Coexistence in disturbance-prone communities: how a resistance-resilience trade-off generates coexistence via the storage effect. American Naturalist, 173, 3043.Google Scholar
Miller, A.E., Bowman, W.D. & Suding, K.N. (2007) Plant uptake of inorganic and organic nitrogen: neighbour identity matters. Ecology, 88, 18321840.Google Scholar
Miller, G., Mangan, J., Pollard, D., Thompson, S., Felzer, B. & Magee, J. (2005) Sensitivity of the Australian monsoon to insolation and vegetation: implications for human impact on continental moisture balance. Geology, 33, 6568.Google Scholar
Miller, M.F. (1995) Acacia seed survival, seed germination and seedling growth following pod consumption by large herbivores and seed chewing rodents. African Journal of Ecology, 33, 194210.Google Scholar
Miller, T.E. & Werner, P.A. (1987) Competitive effects and responses between plant species in a first-year old-field community. Ecology, 68, 12011210.Google Scholar
Millington, W.F. & Chaney, W.R. (1973) Shedding of shoots and branches. In Shedding of Plant Parts (ed. Kozlowski, T.T.), pp. 149204. Academic Press, New York.Google Scholar
Minchinton, T.E., Simpson, J.C. & Bertness, M.D. (2006) Mechanisms of exclusion of native coastal marsh plants by an invasive grass. Journal of Ecology, 94, 342354.Google Scholar
Mirams, R.V. (1957) Aspects of the natural regeneration of the kauri (Agathis australis Salisb.). Transactions of the Royal Society of New Zealand, 84, 661680.Google Scholar
Mirkin, B.M. (1994) Which plant communities do exist? Journal of Vegetation Science, 5, 283284.Google Scholar
Mitamura, M., Yamamura, Y. & Nakano, T. (2009) Large-scale canopy opening causes decreased photosynthesis in the saplings of shade-tolerant conifer, Abies veitchii. Tree Physiology, 29, 137145.Google Scholar
Mitchell, C., Tilman, D. & Groth, J. (2002) Effects of grassland plant species diversity, abundance, and composition on foliar fungal disease. Ecology, 83, 17131726.Google Scholar
Mitchell, C.E. (2003) Trophic control of grassland production and biomass by pathogens. Ecology Letters, 6, 147155.Google Scholar
Mitchell, C.E. & Power, A.G. (2003) Release of invasive plants from fungal and viral pathogens. Nature, 421, 625627.Google Scholar
Mitchell, C.E. & Power, A.G. (2006) Disease dynamics in plant communities. Disease Ecology: Community Structure and Pathogen Dynamics (eds Collinge, S.K. & Ray, C.), pp. 5872. Oxford University Press, Oxford.Google Scholar
Mitchell, C.E., Reich, P.B., Tilman, D. & Groth, J.V. (2003) Effects of elevated CO2, nitrogen deposition, and decreased species diversity on foliar fungal plant disease. Global Change Biology, 9, 438451.Google Scholar
Mitchell, S. (1989) Primary production in a shallow eutrophic lake dominated alternately by phytoplankton and by submerged macrophytes. Aquatic Botany, 33, 101110.Google Scholar
Mitchley, J. & Grubb, P. (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, 11391166.Google Scholar
Mittelbach, G., Steiner, C., Scheiner, S., Gross, K., Reynolds, H., Waide, R., Willig, M., Dodson, S. & Gough, L. (2001) What is the observed relationship between species richness and productivity? Ecology, 82, 23812396.Google Scholar
Moeller, D. (2004) Facilitative interactions among plants via shared pollinators. Ecology, 85, 32893301.Google Scholar
Mohler, C. (1990) Co-occurrence of oak subgenera: implications for niche differentiation. Bulletin of the Torrey Botanical Club, 117, 247255.Google Scholar
Molloy, B., Partridge, T. & Thomas, W. (1991) Decline of tree lupin (Lupinus arboreus) on Kaitorete Spit, Canterbury, New-Zealand, 1984–1990. New Zealand Journal of Botany, 29, 349352.Google Scholar
Moog, D., Kahmen, S. & Poschlod, P. (2005) Application of C-S-R- and LHS-strategies for the distinction of differently managed grasslands. Basic and Applied Ecology, 6, 133143.Google Scholar
Moore, J.C., de Ruiter, P.C. & Hunt, H.W. (1993) Influence of productivity on the stability of real and modelled ecosystems. Science, 261, 906908.Google Scholar
Moore, J.L., Mouquet, N., Lawton, J.H. & Loreau, M. (2001) Coexistence, saturation and invasion resistance in simulated plant assemblages. Oikos, 94, 303314.Google Scholar
Moore, R.M. & Williams, J.D. (1976) A study of a subalpine woodland-grassland boundary. Australian Journal of Ecology, 1, 145153.Google Scholar
Morrison, J.A. & Mauck, K. (2007) Experimental field comparison of native and non-native maple seedlings: natural enemies, ecophysiology, growth and survival. Journal of Ecology, 95, 10361049.Google Scholar
Motzkin, G., Orwig, D.A. & Foster, D.R. (2002) Vegetation and disturbance history of a rare dwarf pitch pine community in western New England. USA Journal of Biogeography, 29, 14551467.Google Scholar
Mouillot, D., Mason, N.W.H. & Wilson, J. (2007) Is the abundance of species determined by their functional traits? A new method with a test using plant communities. Oecologia, 152, 729737.Google Scholar
Mouillot, D. & Wilson, J.B. (2002) Can we tell how a community was constructed? A comparison of five indices for their ability to identify theoretical models of community construction. Theoretical Population Biology, 61, 141151.Google Scholar
Mouquet, N., Leadley, P., Mériguet, J. & Loreau, M. (2004) Immigration and local competition in herbaceous plant communities: a three-seed-sowing experiment. Oikos, 104, 7790.Google Scholar
Mueller-Dombois, D. & Ellenberg, H. (1974) Aims and Methods of Vegetation Ecology. John Wiley and Sons, New York.Google Scholar
Mukhortova, L.V., Kirdyanov, A.V., Myglan, V.S. & Guggenberger, G. (2009) Wood transformation in dead-standing trees in the forest-tundra of Central Siberia. Biology Bulletin [Russian Academy of Sciences], 36, 5865.Google Scholar
Muko, S. & Iwasa, Y. (2003) Incomplete mixing promotes species coexistence in a lottery model with permanent spatial heterogeneity. Theoretical Population Biology, 64, 359368.Google Scholar
Muler, A.L., Oliveira, R.S., Lambers, H. & Veneklaas, E.J. (2014) Does cluster-root activity benefit nutrient uptake and growth of co-existing species? Oecologia, 174, 2331.Google Scholar
Munday, P.L. (2004) Competitive coexistence of coral-dwelling fishes: the lottery hypothesis revisited. Ecology, 85, 623628.Google Scholar
Murillo, N., Laterra, P. & Monterubbianesi, G. (2007) Post-dispersal granivory in a tall-tussock grassland: a positive feedback mechanism of dominance? Journal of Vegetation Science, 18, 799806.Google Scholar
Muro-Pérez, G., Jurado, E., Flores, J., Sánchez-Salas, J., García-Pérez, J. & Estrada, E. (2012) Positive effects of native shrubs on three specially protected cacti species in Durango, Mexico. Plant Species Biology, 27, 5358.Google Scholar
Murphy, G.P. & Dudley, S.A. (2009) Kin recognition: competition and cooperation in Impatiens (Balsaminaceae). American Journal of Botany, 96, 19901996.Google Scholar
Murray, B.R. (1998) Density-dependent germination and the role of seed leachate. Australian Journal of Ecology, 23, 411418.Google Scholar
Murrell, D.J. & Law, R. (2003) Heteromyopia and the spatial coexistence of similar competitors. Ecology Letters, 6, 4859.Google Scholar
Musselman, L.J. & Press, M.C. (1995) Introduction to parasitic plants. In Parasitic Plants (eds Press, M.C. & Graves, J.D.), pp. 113. Chapman & Hall, London.Google Scholar
Muth, C.C. & Bazzaz, F.A. (2002) Tree seedling canopy responses to conflicting photosensory cues. Oecologia, 132, 197204.Google Scholar
Myers, B.J. & Talsma, T. (1992) Site water-balance and tree water status in irrigated and fertilised stands of Pinus radiata. Forest Ecology and Management, 52, 1742.Google Scholar
Myers, J.A., Chase, J.M., Jimenez, I., Jorgensen, P.M., Araujo-Murakami, A., Paniagua-Zambrana, N. & Seidel, R. (2013) Beta-diversity in temperate and tropical forests reflects dissimilar mechanisms of community assembly. Ecology Letters, 16, 151157.Google Scholar
Myers, J.A. & Harms, K.E. (2009) Local immigration, competition from dominant guilds, and the ecological assembly of high-diversity pine savannas. Ecology, 90, 27452754.Google Scholar
Nadkarni, N.M. & Matelson, T.J. (1992) Biomass and nutrient dynamics of epiphytic litterfall in a neotropical montane forest, Costa Rica. Biotropica, 24, 2430.Google Scholar
Naeem, S. & Li, S. (1997) Biodiversity enhances ecosystem reliability. Nature, 390, 507509.Google Scholar
Nagashima, H. & Hikosaka, K. (2011) Plants in a crowded stand regulate their height growth so as to maintain similar heights to neighbours even when they have potential advantages in height growth. Annals of Botany, 108, 207214.Google Scholar
Nantel, P. & Gagnon, D. (1999) Variability in the dynamics of northern peripheral versus southern populations of two clonal plant species, Helianthus divaricatus and Rhus aromatica. Journal of Ecology, 87, 748760.Google Scholar
Nason, J.D., Herre, E.A. & Hamrick, J.L. (1998) The breeding structure of a tropical keystone plant resource. Nature, 391, 685687.Google Scholar
Nee, S. & May, R.M. (1992) Dynamics of metapopulations: habitat destruction and competitive coexistence. Journal of Animal Ecology, 61, 3740.Google Scholar
Newman, E.I. (1978) Allelopathy: adaptation or accident? In Biochemical Aspects of Plant and Animal Coevolution (ed. Harborne, J.B.), pp. 327342. Academic Press, London.Google Scholar
Newman, E.I. & Eason, W.R. (1993) Rates of phosphorus transfer within and between ryegrass (Lolium perenne) plants. Functional Ecology, 7, 242248.Google Scholar
Newman, E.I. & Ritz, K. (1986) Evidence on the pathways of phosphorus transfer between vesicular-arbuscular mycorrhizal plants. New Phytologist, 104, 7787.Google Scholar
Newman, E.I. & Rovira, A.D. (1975) Allelopathy among some British grassland species. Journal of Ecology, 63, 727737.Google Scholar
Niinemets, U. (1996) Changes in foliage distribution with relative irradiance and tree size: differences between the saplings of Acer platanoides acid Quercus robur. Ecological Research, 11, 269281.Google Scholar
Niinemets, U. (1997) Role of foliar nitrogen in light harvesting and shade tolerance of four temperate deciduous woody species. Functional Ecology, 11, 518531.Google Scholar
Niklas, K.J. (1998) Effects of vibration on mechanical properties and biomass allocation pattern of Capsella bursa-pastoris (Cruciferae). Annals of Botany, 82, 147156.Google Scholar
Niklas, K.J. & Hammond, S.T. (2013) Biophysical effects on plant competition and coexistence. Functional Ecology, 27, 854864.Google Scholar
Nippert, J.B. & Knapp, A.K. (2007) Linking water uptake with rooting patterns in grassland species. Oecologia, 153, 261272.Google Scholar
Nobel, P.S. (1997) Root distribution and seasonal production in the northwestern Sonoran Desert for a C3 subshrub, a C4 bunchgrass, and a CAM leaf succulent. American Journal of Botany, 84, 949955.Google Scholar
Nord, E.A., Zhang, C.C. & Lynch, J.P. (2011) Root responses to neighbouring plants in common bean are mediated by nutrient concentration rather than self/non-self recognition. Functional Plant Biology, 38, 941952.Google Scholar
Northup, R.R., Dahlgren, R.A. & McColl, J.G. (1998) Polyphenols as regulators of plant-litter-soil interactions in northern California’s pygmy forest: a positive feedback? Biogeochemistry, 42, 189220.Google Scholar
Norton, D.A. (1997) Host specificity and spatial distribution patterns of mistletoes. In New Zealand’s Loranthaceous Mistletoes (eds de Lange, P.J. & Norton, D.A.), pp. 105109. Department of Conservation, Wellington, NZ.Google Scholar
Novoplansky, A., Cohen, D. & Sachs, T. (1990) How Portulaca seedlings avoid their neighbours. Oecologia, 82, 490493.Google Scholar
Novotny, V. & Basset, Y. (2005) Host specificity of insect herbivores in tropical forests. Proceedings of the Royal Society of London B, 272, 10831090.Google Scholar
O’Brien, K.L. (1996) Tropical deforestation and climate change. Progress in Physical Geography, 20, 311335.Google Scholar
O’Brien, T.A., Moorby, J. & Whittington, W.J. (1967) The effect of management and competition on the uptake of 32phosphorus by ryegrass, meadow fescue and their natural hybrid. Journal of Applied Ecology, 4, 513520.Google Scholar
O’Connor, I. & Aarssen, L.W. (1988) Species association patterns in abandoned sand quarries. Vegetatio, 73, 101109.Google Scholar
Odling-Smee, F.J. (1988) Niche-constructing phenotypes. In The Role of Behaviour in Evolution (ed. Plotkin, H.C.), pp. 73132. MIT Press, Cambridge, MA.Google Scholar
Odling-Smee, F.J., Laland, K.N. & Feldman, M.W. (1996) Niche construction. American Naturalist, 147, 641648.Google Scholar
Odling-Smee, F.J., Laland, K.N. & Feldman, M.W. (2003) Niche Construction: The Neglected Process in Evolution. Princeton University Press, Princeton, NJ.Google Scholar
Odum, H.T. (1971) Environment, Power and Society. Wiley-Interscience, New York.Google Scholar
Ogden, J. (1985) Past, present and future: studies on the population dynamics of some long-lived trees. In Studies on Plant Demography (ed. White, J.), pp. 316. Academic Press, London.Google Scholar
Okuyama, T. & Holland, J.N. (2008) Network structural properties mediate the stability of mutualistic communities. Ecology Letters, 11, 208216.Google Scholar
Olff, H., Hoorens, B., de Goede, R.G.M., van der Putten, W.H. & Gleichman, J.M. (2000) Small-scale shifting mosaics of two dominant grassland species: the possible role of soil-borne pathogens. Oecologia, 125, 4554.Google Scholar
Oliver, L.R. & Schreiber, M.V. (1974) Competition for CO2 in a heteroculture. Weed Science, 22, 125130.Google Scholar
Onipchenko, V.G., Blinnikov, M.S., Gerasimova, M.A., Volkova, E.V. & Cornelissen, J.H.C. (2009) Experimental comparison of competition and facilitation in alpine communities varying in productivity. Journal of Vegetation Science, 20, 718727.Google Scholar
Orians, G.H. (1975) Diversity, stability and maturity in natural ecosystems. In Unifying Concepts in Ecology (eds van Dobben, W.H. & Lowe-McConnell, R.H.), pp. 139150. Junk, The Hague.Google Scholar
Orians, G.H., Dirzo, R. & Cushman, J.H. (1996) Biodiversity and Ecosystem Processes in Tropical Forests. Springer, Berlin.Google Scholar
Orians, G.H. & Paine, R.T. (1983) Convergent evolution at the community level. In Coevolution (eds Futuyma, D.J. & Slatkin, M.), pp. 431458. Sinauer, Sunderland.Google Scholar
Orians, G.H. & Solbrig, O.T. (1977) A cost-income model of leaves and roots with special reference to arid and semiarid areas. American Naturalist, 111, 677690.Google Scholar
Orrock, J.L. & Christopher, C.C. (2010) Density of intraspecific competitors determines the occurrence and benefits of accelerated germination. American Journal of Botany, 97, 694699.Google Scholar
Orrock, J.L., Witter, M.S. & Reichman, O.J. (2008) Apparent competition with an exotic plant reduces native plant establishment. Ecology, 89, 11681174.Google Scholar
Otto-Bliesner, B.L. & Upchurch, G.R. (1997) Vegetation-induced warming of high-latitude regions during the late Cretaceous period. Nature, 385, 804807.Google Scholar
Ozinga, W.A., Schaminée, J.H.J., Bekker, R.M., Bonn, S., Poschlod, P., Tackenberg, O., Bakker, J. & van Groenendael, J. (2005) Predictability of plant species composition from environmental conditions is constrained by dispersal limitation. Oikos, 108, 555561.Google Scholar
Padovan, A., Keszei, A., Foley, W.J. & Külheim, C. (2013) Differences in gene expression within a striking phenotypic mosaic Eucalyptus tree that varies in susceptibility to herbivory. BMC Plant Biology, 13, 112.Google Scholar
Page, R.R., da Vinha, S.G. & Agnew, A.D.Q. (1985) The reaction of some sand dune plant species to experimentally imposed environmental change. Vegetatio, 61, 105114.Google Scholar
Paine, R.T. (1969) A note on trophic complexity and community stability. American Naturalist, 103, 9193.Google Scholar
Pake, C.E. & Venable, D.L. (1996) Seed banks in desert annuals: implications for persistence and coexistence in variable environments. Ecology, 77, 14271435.Google Scholar
Palmer, M.W. (1994) Variation in species richness: towards a unification of hypotheses. Folia Geobotanica et Phytotaxonomica, 29, 511530.Google Scholar
Paquette, A., Fontaine, B., Berninger, F., Dubois, K., Lechowicz, M.J., Messier, C., Posada, J.M., Valladares, F. & Brisson, J. (2012) Norway maple displays greater seasonal growth and phenotypic plasticity to light than native sugar maple. Tree Physiology, 32, 13391347.Google Scholar
Parfitt, R.L., Ross, D.J., Coomes, D.A., Richardson, S.J., Smale, M.C. & Dahlgren, R.A. (2005) N and P in New Zealand soil chronosequences and relationships with foliar N and P. Biogeochemistry, 75, 305328.Google Scholar
Parker, J.D. & Hay, M.E. (2005) Biotic resistance to plant invasions? Native herbivores prefer non-native plants. Ecology Letters, 8, 959967.Google Scholar
Parmesan, C. (2000) Unexpected density-dependent effects of herbivory in a wild population of the annual Collinsia torreyi. Journal of Ecology, 88, 392400.Google Scholar
Parrish, J.A.D. & Bazzaz, F.A. (1976) Underground niche separation in successional plants. Ecology, 57, 12811288.Google Scholar
Parrish, J.A.D. & Bazzaz, F.A. (1979) Difference in pollination niche relationships in early and late successional plant communities. Ecology, 60, 597610.Google Scholar
Paterson, D.M. (1989) Short-term changes in the erodibility of intertidal cohesive sediments related to the migratory behavior of epipelic diatoms. Limnology and Oceanography, 34, 223234.Google Scholar
Pauchard, A., García, R.A., Peña, E., González, C., Cavieres, L.A. & Bustamante, R.O. (2008) Positive feedbacks between plant invasions and fire regimes: Teline monspessulana (L.) K. Koch (Fabaceae) in central Chile. Biological Invasions, 10, 547553.Google Scholar
Pavlick, R., Drewry, D.T., Bohn, K., Reu, B. & Kleidon, A. (2013) The Jena Diversity-Dynamic Global Vegetation Model (JeDi-DGVM): a diverse approach to representing terrestrial biogeography and biogeochemistry based on plant functional trade-offs. Biogeosciences, 10, 41374177.Google Scholar
Pearse, I.S., Hughes, K., Shiojiri, K., Ishizaki, S. & Karban, R. (2013) Interplant volatile signaling in willows: revisiting the original talking trees. Oecologia, 172, 869875.Google Scholar
Peco, B. (1989) Modelling Mediterranean pasture dynamics. Vegetatio, 83, 269276.Google Scholar
Peh, K.S.H., Lewis, S.L. & Lloyd, J. (2011) Mechanisms of monodominance in diverse tropical tree-dominated systems. Journal of Ecology, 99, 891898.Google Scholar
Pelletier, B., Fyles, J.W. & Dutilleul, P. (1999) Tree species control and spatial structure of forest floor properties in a mixed-species stand. Ecoscience, 6, 7991.Google Scholar
Pellino, M., Hojsgaard, D., Schmutzer, T., Scholz, U., Horandl, E., Vogel, H. & Sharbel, T.F. (2013) Asexual genome evolution in the apomictic Ranunculus auricomus complex: examining the effects of hybridization and mutation accumulation. Molecular Ecology, 22, 59085921.Google Scholar
Peltzer, D.A., Wardle, D.A., Allison, V.J., Baisden, W.T., Bardgett, R.D., Chadwick, O.A., Condron, L.M., Parfitt, R.L., Porder, S., Richardson, S.J., Turner, B.L., Vitousek, P.M., Walker, J. & Walker, L.R. (2010) Understanding ecosystem retrogression. Ecological Monographs, 80, 509529.Google Scholar
Peltzer, D.A. & Wilson, S.D. (2001) Competition and environmental stress in temperate grasslands. In Competition and Succession in Pastures (eds Tow, P.G. & Lazenby, A.), pp. 193212. CAB International, Wallingford.Google Scholar
Peltzer, D.A., Wilson, S.D. & Gerry, A.K. (1998) Competition intensity along a productivity gradient in a low-diversity grassland. American Naturalist, 151, 465485.Google Scholar
Penfould, W.T. (1964) The relation of grazing to plant succession in the tallgrass prairie. Journal of Range Management, 17, 256260.Google Scholar
Pennings, S.C., Grant, M.-B. & Bertness, M.D. (2005) Plant zonation in low-latitude salt marshes: disentangling the roles of flooding, salinity and competition. Journal of Ecology, 93, 159167.Google Scholar
Peppler-Lisbach, C. & Kleyer, M. (2009) Patterns of species richness and turnover along the pH gradient in deciduous forests: testing the continuum hypothesis. Journal of Vegetation Science, 20, 984995.Google Scholar
Perez-Salicrup, D.R. & Barker, M.G. (2000) Effect of liana cutting on water potential and growth of adult Senna multijuga (Caesalpinoideae) trees in a Bolivian tropical forest. Oecologia, 124, 469475.Google Scholar
Petermann, J.S., Fergus, A.J.F., Roscher, C., Turnbull, L.A., Weigelt, A. & Schmid, B. (2010) Biology, chance, or history? The predictable reassembly of temperate grassland communities. Ecology, 91, 408421.Google Scholar
Petermann, J.S., Fergus, A.J.F., Turnbull, L.A. & Schmid, B. (2008) Janzen-Connell effects are widespread and strong enough to maintain diversity in grasslands. Ecology, 89, 23992406.Google Scholar
Peterson, C.J. & Pickett, S.T.A. (1995) Forest reorganization: a case study in an old-growth forest catastrophic blowdown. Ecology, 76, 763774.Google Scholar
Petraitis, P.S. & Hoffman, C. (2010) Multiple stable states and relationship between thresholds in processes and states. Marine Ecology Progress Series, 413, 189200.Google Scholar
Petritan, A.M., Nuske, R.S., Petritan, I.C. & Tudose, N.C. (2013) Gap disturbance patterns in an old-growth sessile oak (Quercus petraea L.)-European beech (Fagus sylvatica L.) forest remnant in the Carpathian Mountains, Romania. Forest Ecology and Management, 308, 6775.Google Scholar
Pfisterer, A.B. & Schmid, B. (2002) Diversity-dependent production can decrease the stability of ecosystem functioning. Nature, 416, 8486.Google Scholar
Phillips, J. (1935) Succession, development, the climax, and the complex organism: an analysis of concepts. III. The complex organism: conclusions. Journal of Ecology, 23, 488508.Google Scholar
Phillips, O.L., Martínez, R.V., Mendoza, A.M., Baker, T.R. & Vargas, N.M. (2005) Large lianas as hyperdynamic elements of the tropical forest canopy. Ecology, 86, 12501258.Google Scholar
Pianka, E.R. (1980) Guild structure in desert lizards. Oikos, 35, 194201.Google Scholar
Pianka, E.R. & Horn, H.S. (2005) Ecology’s legacy from Robert MacArthur. In Ecological Paradigms Lost: Routes of Theory Change (eds Cuddington, K. & Beisner, B.E.), pp. 213232. Elsevier, Amsterdam.Google Scholar
Pickett, S.T.A. (1976) Succession: an evolutionary interpretation. American Naturalist, 110, 107119.Google Scholar
Pickett, S.T.A. & Bazzaz, F.A. (1978) Organization of an assemblage of early successional species on a soil moisture gradient. Ecology, 59, 12481255.Google Scholar
Pielou, E.C. (1975) Ecological Diversity. Wiley, New York.Google Scholar
Pielou, E.C. (1978) Latitudinal overlap of seaweed species: evidence for quasi-sympatric speciation. Journal of Biogeography, 5, 227238.Google Scholar
Pielou, E.C. & Routledge, R.D. (1976) Salt marsh vegetation: latitudinal gradients in the zonation patterns. Oecologia, 24, 311321.Google Scholar
Pierik, R. & de Wit, M. (2014) Shade avoidance: phytochrome signalling and other aboveground neighbour detection cues. Journal of Experimental Botany, 65, 28152824.Google Scholar
Pierik, R., Mommer, L. & Voesenek, L. (2013) Molecular mechanisms of plant competition: neighbour detection and response strategies. Functional Ecology, 27, 841853.Google Scholar
Pieterse, A.H. & Murphy, K. (1993) Aquatic Weeds: the Ecology and Management of Nuisance Aquatic Vegetation. Oxford University Press, Oxford.Google Scholar
Pigliucci, M. & Murren, C.J. (2003) Perspective: genetic assimilation and a possible evolutionary paradox: Can macroevolution sometimes be so fast as to pass us by? Evolution, 57, 14551464.Google Scholar
Pigott, C.D. (1970) The response of plants to climate and climatic change. In The Flora of a Changing Britain (ed. Perring, F.H.), pp. 3244. Classey, Faringdon.Google Scholar
Pigott, C.D. (1980) [Review of Grime (1979), Plant strategies and vegetation processes]. Journal of Ecology, 68, 704706.Google Scholar
Pigott, C.D. & Huntley, J.P. (1981) Factors controlling the distribution of Tilia cordata at the northern limits of its geographical range. III. Nature and causes of seed sterility. New Phytologist, 87, 817839.Google Scholar
Platt, W.J. & Connell, J.H. (2003) Natural disturbances and directional replacement of species. Ecological Monographs, 73, 507522.Google Scholar
Pleasants, J.M. (1980) Competition for bumblebee pollinators in Rocky Mountain plant communities. Ecology, 61, 14461459.Google Scholar
Pleasants, J.M. (1990) Null-model tests for competitive displacement: the fallacy of not focusing on the whole community. Ecology, 71, 10781084.Google Scholar
Ploeg, A.T. & Maris, P.C. (1999) Effect of temperature on suppression of Meloidogyne incognita by Tagetes cultivars. Journal of Nematology, 31, 709714.Google Scholar
Plumptre, A.J. (1994) The effects of trampling damage by herbivores on the vegetation of the Parc National des Volcans, Rwanda. African Journal of Ecology, 32, 115129.Google Scholar
Pocock, M.J.O., Evans, D.M. & Memmott, J. (2012) The robustness and restoration of a network of ecological networks. Science, 335, 973977.Google Scholar
Pokon, R., Novotny, V. & Samuelson, G.A. (2005) Host specialization and species richness of root-feeding chrysomelid larvae (Chrysomelidae, Coleoptera) in a New Guinea rain forest. Journal of Tropical Ecology, 21, 595604.Google Scholar
Poli, C.H.E.C., Hodgson, J., Cosgrove, G.P. & Arnold, G.C. (2006) Selective behaviour in cattle grazing pastures of strips of birdsfoot trefoil and red clover. 1. The effects of relative sward area. Journal of Agricultural Science, 144, 165171.Google Scholar
Poorter, L., Bongers, F., Sterck, F.J. & Woll, H. (2005) Beyond the regeneration phase: differentiation of height-light trajectories among tropical tree species. Journal of Ecology, 93, 256267.Google Scholar
Post, D.M. & Palkovacs, E.P. (2009) Eco-evolutionary feedbacks in community and ecosystem ecology: interactions between the ecological theatre and the evolutionary play. Philosophical Transactions of the Royal Society B, 364, 16291640.Google Scholar
Potts, A.J., Midgley, J.J., Child, M.F., Larsen, C. & Hempson, T. (2011) Coexistence theory in the Cape Floristic Region: revisiting an example of leaf niches in the Proteaceae. Austral Ecology, 36, 212219.Google Scholar
Poulin, B., Wright, S.J., Lefebvre, G. & Calderon, O. (1999) Interspecific synchrony and asynchrony in the fruiting phenologies of congeneric bird-dispersed plants in Panama. Journal of Tropical Ecology, 15, 213227.Google Scholar
Poulson, T.L. & Platt, W.J. (1996) Replacement patterns of beech and sugar maple in Warren Woods, Michigan. Ecology, 77, 12341253.Google Scholar
Pound, R. & Clements, F.E. (1900) The Phytogeography of Nebraska, 2nd edn. Botanical Seminar, Lincoln, NE.Google Scholar
Power, A.G. & Mitchell, C.E. (2004) Pathogen spillover in disease epidemics. American Naturalist, 164, 7989.Google Scholar
Presley, S.J., Higgins, C.L. & Willig, M.R. (2010) A comprehensive framework for the evaluation of metacommunity structure. Oikos, 119, 908917.Google Scholar
Press, M.C. (1998) Dracula or Robin Hood: a functional role for root hemiparasites in nutrient poor ecosystems. Oikos, 82, 609611.Google Scholar
Preston, C.D., Pearman, D.A. & Dines, T.D. (2002) New Atlas of the British & Irish Flora: An Atlas of the Vascular Plants of Britain, Ireland, the Isle of Man and the Channel Islands. Oxford University Press, Oxford.Google Scholar
Preston, F.W. (1948) The commonness and rarity of species. Ecology, 29, 254283.Google Scholar
Preston, F.W. (1962) The canonical distribution of commonness and rarity. Ecology, 43, 185215.Google Scholar
Prider, J., Watling, J. & Facelli, J.M. (2009) Impacts of a native parasitic plant on an introduced and a native host species: implications for the control of an invasive weed. Annals of Botany, 103, 107115.Google Scholar
Prins, H.H.T. & van der Jeugd, H.P. (1993) Herbivore population crashes and woodland structure in East Africa. Journal of Ecology, 81, 305314.Google Scholar
Prokopy, R.J., Diehl, S.R. & Cooley, S.S. (1988) Behavioral evidence for host races in Rhagoletis pomonella flies. Oecologia, 76, 138147.Google Scholar
Putz, F.E. (1984) The natural history of lianas on Barro Colorado Island, Panama. Ecology, 65, 17131724.Google Scholar
Putz, F.E. (1991) Silvicultural effects of lianas. In The Biology of Vines (eds Putz, F.E. & Mooney, H.A.), pp. 493501. Cambridge University Press, Cambridge, UK.Google Scholar
Putz, F.E. (1995) Vines in treetops: consequences of mechanical dependence. In Forest Canopies (eds Lowman, M.D. & Nadkari, N.M.), pp. 311323. Academic Press, San Diego.Google Scholar
Putz, F.E., Parker, G.G. & Archibald, R.M. (1984) Mechanical abrasion and intercrown spacing. American Midland Naturalist, 112, 2428.Google Scholar
Pyne, S.J. (1991) Burning Bush: a Fire History of Australia. Allen and Unwin, Sydney.Google Scholar
Rabinowitz, D., Rapp, J.K. & Dixon, P.M. (1984) Competitive abilities of sparse grass species: means of persistence or cause of abundance? Ecology, 65, 11441154.Google Scholar
Rajaniemi, T.K. (2003) Evidence for size asymmetry of belowground competition. Basic and Applied Ecology, 4, 239247.Google Scholar
Rastetter, E.B., Gough, L., Hartley, A.E., Herbert, D.A., Nadelhoffer, K.J. & Williams, M. (1999) A revised assessment of species redundancy and ecosystem reliability. Conservation Biology, 13, 440443.Google Scholar
Rathcke, B. & Lacey, E.P. (1985) Phenological patterns of terrestrial plants. Annual Review of Ecology and Systematics, 16, 179214.Google Scholar
Ratnadass, A., Fernandes, P., Avelino, J. & Habib, R. (2012) Plant species diversity for sustainable management of crop pests and diseases in agroecosystems: a review. Agronomy for Sustainable Development, 32, 273303.Google Scholar
Raunkiaer, C. (1909) Formationsundersøgelse og Formationsstatistik. Botanisk Tidsskrift, 30, 20132.Google Scholar
Raventos, J., Wiegand, T. & Deluis, M. (2010) Evidence for the spatial segregation hypothesis: a test with nine-year survivorship data in a Mediterranean shrubland. Ecology, 91, 21102120.Google Scholar
Rebele, F. (2000) Competition and coexistence of rhizomatous perennial plants along a nutrient gradient. Plant Ecology, 147, 7794.Google Scholar
Redmond, M.D., Wilbur, R.B. & Wilbur, H.M. (2012) Recruitment and dominance of Quercus rubra and Quercus alba in a previous Oak-Chestnut forest from the 1980s to 2008. American Midland Naturalist, 168, 427442.Google Scholar
Reich, P.B. (2014) The world-wide ‘fast-slow’ plant economics spectrum: a traits manifesto. Journal of Ecology, 102, 275301.Google Scholar
Reich, P.B., Tilman, D., Isbell, F., Mueller, K., Hobbie, S.E., Flynn, D.F.B. & Eisenhauer, N. (2012) Impacts of biodiversity loss escalate through time as redundancy fades. Science, 336, 589592.Google Scholar
Reich, P.B., Wright, I.J., Cavender-Bares, J., Craine, J.M., Oleksyn, J. & Westoby, M. (2003) The evolution of plant functional variation: traits, spectra, and strategies. Journal of Plant Sciences 164, 143164.Google Scholar
Reichle, D.E., O’Neill, R.V. & Harris, W.F. (1975) Principles of material and energy exchange in ecosystems. In Unifying Concepts in Ecology (eds van Dobben, W.H. & Lowe-McConnell, R.H.), pp. 2743. Junk, The Hague.Google Scholar
Reicosky, D.C. (1989) Diurnal and seasonal trends in carbon dioxide concentrations in corn and soybean canopies as affected by tillage and irrigation. Agricultural and Forest Meteorology, 48, 285303.Google Scholar
Reinhart, K.O. & Callaway, R.M. (2004) Soil biota facilitate exotic Acer invasions in Europe and North America. Ecological Applications, 14, 17371745.Google Scholar
Reinhart, K.O., Gurnee, J., Tirado, R. & Callaway, R.M. (2006) Invasion through quantitative effects: intense shade drives native decline and invasive success. Ecological Applications, 16, 18211831.Google Scholar
Reintam, E., Trukmann, K. & Kuht, J. (2008) Effect of Cirsium arvense L. on soil physical properties and crop growth. Agricultural and Food Science, 17, 153164.Google Scholar
Rejmánek, M. (1996) Species richness and resistance to invasions. In Biodiversity and Ecosystem Processes in Tropical Forests (eds Orians, G.H., Dirzo, R. & Cushman, J.H.), pp. 153172. Springer, Berlin.Google Scholar
Rejmánek, M. (1999) Holocene invasions: finally the resolution ecologists were waiting for. Trends in Ecology and Evolution, 14, 810.Google Scholar
Reuschel, D., Mattheck, C. & Althaus, C. (1998) The mechanical effect of climbing plants upon the host tree. Allgemeine Forst und Jagdzeitung, 169, 8791.Google Scholar
Revilla, T. & Weissing, F.J. (2008) Nonequilibrium coexistence in a competition model with nutrient storage. Ecology, 89, 865877.Google Scholar
Richards, P.W. (1996) The Tropical Rain Forest: An Ecological Study, 2nd edn. Cambridge University Press, Cambridge, UK.Google Scholar
Richardson, A.E., Barea, J.M., McNeill, A.M. & Prigent-Combaret, C. (2009) Acquisition of phosphorus and nitrogen in the rhizosphere and plant growth promotion by microorganisms. Plant and Soil, 321, 305339.Google Scholar
Richardson, S.J., Peltzer, D.A., Allen, R.B., McGlone, M.S. & Parfitt, R.L. (2004) Rapid development of phosphorus limitation in temperate rainforest along the Franz Josef soil chronosequence. Oecologia, 139, 267276.Google Scholar
Rietkerk, M., van den Bosch, F. & van de Koppel, J. (1997) Site-specific properties and irreversible vegetation changes in semi-arid grazing systems. Oikos, 80, 241252.Google Scholar
Robberecht, R., Mahall, B.E. & Nobel, P.S. (1983) Experimental removal of intraspecific competitors – effect on water relations and productivity of a desert bunchgrass, Hilaria rigida. Oecologia, 60, 2124.Google Scholar
Robertson, G.P., Hutson, M.A., Evans, F.C. & Tiedje, J.M. (1988) Spatial variability in a successional plant community: patterns of nitrogen availability. Ecology, 69, 15171524.Google Scholar
Robinson, J.V. & Dickerson, J.E. (1987) Does invasion sequence affect community structure? Ecology, 68, 587595.Google Scholar
Rodríguez, M.A. & Gómezal-Sal, A. (1994) Stability may decrease with diversity in grassland communities – empirical-evidence from the 1986 Cantabrian mountains (Spain) drought. Oikos, 71, 177180.Google Scholar
Rodwell, J.S. (1991–2000) British Plant Com0munities. Cambridge University Press, Cambridge, UK.Google Scholar
Rogers, R.S. (1983) Small-area coexistence of vernal forest herbs: does functional similarity of plants matter. American Naturalist, 121, 834850.Google Scholar
Roggy, J.C., Moiroud, A., Lensi, R. & Domenach, A.M. (2004) Estimating N transfers between N2-fixing actinorhizal species and the non-N2-fixing Prunus avium under partially controlled conditions. Biology and Fertility of Soils, 39, 312319.Google Scholar
Root, R.B. (1967) The niche exploitation pattern of the blue-gray gnatcatcher. Ecological Monographs, 37, 317350.Google Scholar
Roques, K.G., O’Connor, T.G. & Watkinson, A.R. (2001) Dynamics of shrub encroachment in an African savannah: relative influences of fire, herbivory, rainfall and density dependence. Journal of Applied Ecology, 38, 268280.Google Scholar
Roscher, C., Weigelt, A., Proulx, R., Marquard, E., Schumacher, J., Weisser, W.W. & Schmid, B. (2011) Identifying population- and community-level mechanisms of diversity-stability relationships in experimental grasslands. Journal of Ecology, 99, 14601469.Google Scholar
Rosindell, J., Hubbell, S.P., He, F.L., Harmon, L.J. & Etienne, R.S. (2012) The case for ecological neutral theory. Trends in Ecology and Evolution, 27, 203208.Google Scholar
Rousseaux, M.C., Hall, A.J. & Sánchez, R.A. (1996) Far-red enrichment and photosynthetically active radiation level influence leaf senescence in field-grown sunflower. Physiologia Plantarum, 96, 217224.Google Scholar
Roxburgh, S.H. & Mokany, K. (2007) Comment on “From plant traits to plant communities: a statistical mechanistic approach to biodiversity”. Science, 316, 1425.Google Scholar
Roxburgh, S.H. & Mokany, K. (2010) On testing predictions of species relative abundance from maximum entropy optimisation. Oikos, 119, 583590.Google Scholar
Roxburgh, S.H., Shea, K. & Wilson, J.B. (2004) The Intermediate Disturbance Hypothesis: patch dynamics and mechanisms of species coexistence. Ecology, 85, 359371.Google Scholar
Roxburgh, S.H., Watkins, A.J. & Wilson, J.B. (1993) Lawns have vertical stratification. Journal of Vegetation Science, 4, 699704.Google Scholar
Roxburgh, S.H. & Wilson, J.B. (2000a) Stability and coexistence in a lawn community: mathematical prediction of stability using a community matrix with parameters derived from competition experiments. Oikos, 88, 395408.Google Scholar
Roxburgh, S.H. & Wilson, J.B. (2000b) Stability and coexistence in a lawn community: experimental assessment of the stability of the actual community. Oikos, 88, 409423.Google Scholar
Russ, G.R. (1982) Overgrowth in a marine epifaunal community: competitive hierarchies and competitive networks. Oecologia, 53, 1219.Google Scholar
Ryel, R.J., Caldwell, M.M., Leffler, A.J. & Yoder, C.K. (2003) Rapid soil moisture recharge to depth by roots in a stand of Artemisia tridentata. Ecology, 84, 757764.Google Scholar
Ryser, P. (1993) Influences of neighbouring plants on seedling establishment in limestone grassland. Journal of Vegetation Science, 4, 195202.Google Scholar
Sage, R.B. (1999) Weed competition in willow coppice crops: the cause and extent of yield losses. Weed Research, 39, 399411.Google Scholar
Sale, P.F. (1977) Maintenance of high diversity in coral reef fish communities. American Naturalist, 111, 337359.Google Scholar
Šamonil, P., Dolezelova, P., Vasickova, I., Adam, D., Valtera, M., Kral, K., Janik, D. & Sebkova, B. (2013a) Individual-based approach to the detection of disturbance history through spatial scales in a natural beech-dominated forest. Journal of Vegetation Science, 24, 11671184.Google Scholar
Šamonil, P., Schaetzl, R.J., Valtera, M., Goliáš, V., Baldrian, P., Vaší?ková, I., Adam, D., Janík, D. & Hort, L. (2013b) Crossdating of disturbances by tree uprooting: can treethrow microtopography persist for 6000 years? Forest Ecology and Management, 307, 123135.Google Scholar
Samuels, C.L. & Drake, J.A. (1997) Divergent perspectives on community convergence. Trends in Ecology and Evolution, 12, 427432.Google Scholar
Sanders, N.J., Gotelli, N.J., Heller, N.E. & Gordon, D.M. (2003) Community disassembly by an invasive species. Proceedings of the National Academy of Sciences of the U.S.A., 100, 24742477.Google Scholar
Sarmiento, L., Llambí, L.D., Escalona, A. & Marquez, N. (2003) Vegetation patterns, regeneration rates and divergence in an old-field succession of the high tropical Andes. Plant Ecology, 166, 6374.Google Scholar
Sartori, F., Lal, R., Ebinger, M.H. & Miller, R.O. (2007) Tree species and wood ash affect soil in Michigan’s Upper Peninsula. Plant and Soil, 298, 125144.Google Scholar
Sato, T. (1994) Stand structure and dynamics of wave-type Abies sachalinensis coastal forest. Ecological Research, 9, 7784.Google Scholar
Sayer, E.J. (2006) Using experimental manipulation to assess the roles of leaf litter in the functioning of forest ecosystems. Biological Reviews, 81, 131.Google Scholar
Scheffer, M., Hosper, S.H., Meijer, M.-L., Moss, B. & Jeppesen, E. (1993) Alternative equilibria in shallow lakes. Trends in Ecology and Evolution, 8, 275279.Google Scholar
Schimper, A.F.W. (1898) Pflanzen-Geographie auf Physiologischer Grundlage. Fischer, Jena.Google Scholar
Schimper, A.F.W. (1903) Plant-Geography upon a Physiological Basis. Oxford University Press, Oxford.Google Scholar
Schlesinger, W.H., Raikes, J.A., Hartley, A.E. & Cross, A.E. (1996) On the spatial pattern of soil nutrients in desert ecosystems. Ecology, 77, 364374.Google Scholar
Schliephake, E., Graichen, K. & Rabenstein, F. (2000) Investigations on the vector transmission of the Beet mild yellowing virus (BMYV) and the Turnip yellows virus (TuYV). Zeitschrift für Pflanzenkrankheiten und Pflanzenschutz, 107, 8187.Google Scholar
Schluter, D. (1990) Species-for-species matching. American Naturalist, 136, 560568.Google Scholar
Schmitt, J. & Wulff, R.D. (1993) Light spectral quality, phytochrome and plant competition. Trends in Ecology and Evolution, 8, 4748.Google Scholar
Schmitz, O.J. (1997) Press perturbations and the predictability of ecological interactions in a food web. Ecology, 78, 5569.Google Scholar
Schmitz, O.J. (2004) Perturbation and abrupt shift in trophic control of biodiversity and productivity. Ecology Letters, 7, 403409.Google Scholar
Schnitzer, S.A. & Carson, W.P. (2000) Have we forgotten the forest because of the trees? Trends in Ecology and Evolution, 15, 375376.Google Scholar
Schnitzer, S.A., Kuzee, M.E. & Bongers, F. (2005) Disentangling above- and below-ground competition between lianas and trees in a tropical forest. Journal of Ecology, 93, 11151125.Google Scholar
Schoener, T.W. (1986) Resource partitioning. In Community Ecology: Pattern and Process (eds Kikkawa, J. & Anderson, D.J.), pp. 91126. Blackwell, Melbourne.Google Scholar
Scholte, K. & Vos, J. (2000) Effects of potential trap crops and planting date on soil infestation with potato cyst nematodes and root-knot nematodes. Annals of Applied Biology, 137, 153164.Google Scholar
Schoonhoven, L.M., Jermy, T. & van Loon, J.J.A. (1998) Insect-Plant Biology: From Physiology to Evolution. Chapman and Hall, London.Google Scholar
Schrage, L.J. & Downing, J.A. (2004) Pathways of increased water clarity after fish removal from Ventura Marsh; a shallow, eutrophic wetland. Hydrobiologia, 511, 215231.Google Scholar
Schröder, A., Persson, L. & De Roos, A.M. (2005) Direct experimental evidence for alternative stable states: a review. Oikos, 110, 319.Google Scholar
Schulz, J.P. (1960) Ecological Studies on Rain Forest in Northern Suriname. Noord-Hollandsche, Amsterdam.Google Scholar
Scofield, D.G. (2006) Medial pith cells per meter in twigs as a proxy for mitotic growth rate in the apical meristem. American Journal of Botany, 93, 17401747.Google Scholar
Scott, P.A. & Hansell, R.I.C. (2002) Development of white spruce tree islands in the shrub zone of the forest-tundra. Arctic, 55, 238246.Google Scholar
Scott-Phillips, T.C., Laland, K.N., Shuker, D.M., Dickins, T.E. & West, S.A. (2014) The niche construction perspective: a critical appraisal. Evolution, 68, 12311243.Google Scholar
Scunthorpe, C.D. (1967) The Biology of Aquatic Vascular Plants. Arnold, London.Google Scholar
Sears, A.L.W. & Chesson, P. (2007) New methods for quantifying the spatial storage effect: an illustration with desert annuals. Ecology, 88, 22402247.Google Scholar
Šebková, B., Šamonil, P., Valtera, M., Adam, D. & Janík, D. (2012) Interaction between tree species populations and windthrow dynamics in natural beech-dominated forest, Czech Republic. Forest Ecology and Management, 280, 919.Google Scholar
Seifan, M., Seifan, T., Ariza, C. & Tielbörger, K. (2010) Facilitating an importance index. Journal of Ecology, 98, 356361.Google Scholar
Seifert, R.P. & Seifert, F.H. (1976) A community matrix analysis of Heliconia insect communities. American Naturalist, 110, 461483.Google Scholar
Semchenko, M., Lepik, M., Gotzenberger, L. & Zobel, K. (2012) Positive effect of shade on plant growth: amelioration of stress or active regulation of growth rate? Journal of Ecology, 100, 459466.Google Scholar
Shea, K. & Chesson, P.L. (2002) Community ecology theory as a framework for biological invasions. Trends in Ecology and Evolution, 17, 170176.Google Scholar
Shea, K., Roxburgh, S.H. & Rauschert, E.S.J. (2004) Moving from pattern to process: coexistence mechanisms under intermediate disturbance regimes. Ecology Letters, 7, 491508.Google Scholar
Shepherd, V.A. (1999) Bioelectricity and the rhythms of sensitive plants – the biophysical research of Jagadis Chandra Bose. Current Science, 77, 189195.Google Scholar
Shimizu, Y. & Tabata, H. (1985) Invasion of Pinus lutchuensis and its influence on the native forest of a Pacific island. Journal of Biogeography, 12, 195207.Google Scholar
Shipley, B. & Keddy, P.A. (1987) The individualistic and community-unit concepts as falsifiable hypotheses. Vegetatio, 69, 4756.Google Scholar
Shipley, B., Vile, D. & Garnier, E. (2006) From plant traits to plant communities: a statistical mechanistic approach to biodiversity. Science, 314, 812814.Google Scholar
Shmida, A. (1981) Mediterranean vegetation in California and Israel: similarities and differences. Israel Journal of Botany, 30, 105123.Google Scholar
Shnerb, N.M., Sarah, P., Lavee, H. & Solomon, S. (2003) Reactive glass and vegetation patterns. Physical Review Letters, 90, 38101.Google Scholar
Shreve, F. (1942) The desert vegetation of North America. Botanical Review, 8, 195246.Google Scholar
Shuker, D.M. (2014) Genetic variation in niche construction: a comment on Satz and Nuzhdin. Trends in Ecology and Evolution, 29, 303304.Google Scholar
Shulski, M.D., Walter-Shea, E., Hubbard, K. & Horst, L.G. (2004) Penetration of photosynthetically active and ultraviolet radiation into alfalfa and tall fescue canopies. Agronomy Journal, 96, 15621571.Google Scholar
Shure, D.J. & Phillips, D.L. (1987) Litterfall patterns within different sized disturbance patches in a Southern Appalachian Mountain forest. American Midland Naturalist, 118, 348357.Google Scholar
Siitonen, J. (2001) Forest management, coarse woody debris and saproxylic organisms: Fennoscandian boreal forests as an example. Ecological Bulletins, 49, 1142.Google Scholar
Siitonen, J., Martikainen, P., Punttila, P. & Rauh, J. (2000) Coarse woody debris and stand characteristics in mature managed and old-growth boreal mesic forests in southern Finland. Forest Ecology and Management, 128, 211225.Google Scholar
Silander, J.A. & Antonovics, J. (1982) Analysis of interspecific interactions in a coastal plant community – a perturbation approach. Nature, 298, 557560.Google Scholar
Silver, M. & Di Paolo, E. (2006) Spatial effects favour the evolution of niche construction. Theoretical Population Biology, 70, 387400.Google Scholar
Silvertown, J. (1987) Ecological stability: a test case. American Naturalist, 130, 807810.Google Scholar
Silvertown, J., Dodd, M.E., Gowing, D.J.G. & Mountford, J.O. (1999) Hydrologically-defined niches reveal a basis for species richness in plant communities. Nature, 400, 6163.Google Scholar
Silvertown, J., Holtier, S., Johnson, J. & Dale, P. (1992) Cellular automaton models of interspecific competition for space: the effect of pattern on process. Journal of Ecology, 80, 527534.Google Scholar
Silvertown, J., Lines, C.E.M. & Dale, M.P. (1994) Spatial competition between grasses – rates of mutual invasion between four species and the interaction with grazing. Journal of Ecology, 82, 3138.Google Scholar
Simard, S.W. & Durall, D.M. (2004) Mycorrhizal networks: a review of their extent, function, and importance. Canadian Journal of Botany, 82, 11401165.Google Scholar
Simberloff, D. (1980) A succession of paradigms in ecology: essentialism to materialism and probabilism. Synthese, 43, 339.Google Scholar
Singh, H.P., Batisha, D.R. & Kohlia, R.K. (1999) Autotoxicity: concept, organisms, and ecological significance. Critical Reviews in Plant Sciences, 18, 757772.Google Scholar
Šizling, A.L., Storch, D., Sizlingova, E., Reif, J. & Gaston, K.J. (2009) Species abundance distribution results from a spatial analogy of central limit theorem. Proceedings of the National Academy of Sciences of the U.S.A., 106, 66916695.Google Scholar
Skellam, J.G. (1951) Random dispersal in theoretical populations. Biometrika, 38, 196218.Google Scholar
Skinner, R.H. & Simmons, S.R. (1993) Modulation of leaf elongation, tiller appearance and tiller senescence in spring barley by far-red light. Plant Cell and Environment, 16, 555562.Google Scholar
Smith, B., Moore, S., Grove, P., Harris, N., Mann, S. & Wilson, J. (1994) Vegetation texture as an approach to community structure: community-level convergence in a New Zealand temperate rainforest. New Zealand Journal of Ecology, 18, 4150.Google Scholar
Smith, B. & Wilson, J.B. (2002) Community convergence: ecological and evolutionary. Folia Geobotanica, 37, 171183.Google Scholar
Smith, T.W. & Lundholm, J.T. (2010) Variation partitioning as a tool to distinguish between niche and neutral processes. Ecography, 33, 648655.Google Scholar
Snaydon, R.W. & Davies, T.M. (1982) Rapid divergence of plant populations in response to recent changes in soil conditions. Evolution, 36, 289297.Google Scholar
Soliveres, S., Smit, C. & Maestre, F. (2014) Moving forward on facilitation research: response to changing environments and effects on the diversity, functioning and evolution of plant communities. Biological Reviews, 90, 297313.Google Scholar
Spatharis, S., Orfanidis, S., Panayotidis, P. & Tsirtsis, G. (2011) Assembly processes in upper subtidal macroalgae: the effect of wave exposure. Estuarine Coastal and Shelf Science, 91, 298305.Google Scholar
Srivastava, D.S. & Jefferies, R.L. (1995) Mosaics of vegetation and soil salinity: a consequence of goose foraging in an arctic salt marsh. Canadian Journal of Botany, 73, 7583.Google Scholar
Stansfield, J.H., Perrow, M.R., Tench, D.L., Jowitt, A.J.D. & Taylor, A.A.L. (1997) Submerged macrophytes as refuges for grazing Cladocera against fish predation: observations on seasonal changes in relation to macrophyte cover and predation pressure. Hydrobiologia, 342, 229240.Google Scholar
Stastny, M., Schaffner, U.R.S. & Elle, E. (2005) Do vigour of introduced populations and escape from specialist herbivores contribute to invasiveness? Journal of Ecology, 93, 2737.Google Scholar
Steel, J.B., Wilson, J.B., Anderson, B.J., Lodge, R.H.E. & Tangney, R.S. (2004) Are bryophyte communities different from higher-plant communities?: Abundance relations. Oikos, 104, 479486.Google Scholar
Steinger, T., Körner, C. & Schmid, B. (1996) Long-term persistence in a changing climate: DNA analysis suggests very old ages of clones of alpine Carex curvula. Oecologia, 105, 9499.Google Scholar
Stenberg, J.A., Heijari, J., Holopainen, J.K. & Ericson, L. (2007) Presence of Lythrum salicaria enhances the bodyguard effects of the parasitoid Asecodes mento for Filipendula ulmaria. Oikos, 116, 482490.Google Scholar
Stern, W.R. & Donald, C.M. (1962) The influence of leaf area and radiation on growth of clover in swards. Australian Journal of Agricultural Research, 13, 615623.Google Scholar
Stewart, W.S. & Bannister, P. (1973) Seasonal changes in carbohydrate content of three Vaccinium spp. with particular reference to V. uliginosum L. and its distribution in the British Isles. Flora, 162, 134155.Google Scholar
Stiles, F.G. (1977) Coadapted competitors: the flowering seasons of hummingbird-pollinated plants in a tropical forest. Science, 198, 11771178.Google Scholar
Stiles, F.G. (1979) Reply to Poole & Rathcke. Science, 203, 471.Google Scholar
Stokes, C.J. & Archer, S.R. (2010) Niche differentiation and neutral theory: an integrated perspective on shrub assemblages in a parkland savanna. Ecology, 91, 11521162.Google Scholar
Stokke, S. (1999) Sex differences in feeding-patch choice in a megaherbivore: elephants in Chobe National Park, Botswana. Canadian Journal of Zoology, 77, 17231732.Google Scholar
Stoll, P. & Newbery, D.M. (2005) Evidence of species-specific neighborhood effects in the Dipterocarpaceae of a Bornean rain forest. Ecology, 86, 30483062.Google Scholar
Stoll, P. & Prati, D. (2001) Intraspecific aggregation alters competitive interactions in experimental plant communities. Ecology, 82, 319327.Google Scholar
Stone, L., Gabric, A. & Berman, T. (1996) Ecosystem resilience, stability and productivity: seeking a relationship. American Naturalist, 148, 892903.Google Scholar
Stone, L. & Roberts, A. (1991) Conditions for a species to gain advantage from the presence of competitors. Ecology, 72, 19641972.Google Scholar
Stoutjesdijk, P.H. & Barkman, J.J. (1992) Microclimate, Vegetation and Fauna. Opulus, Knivsta.Google Scholar
Strauss, S.Y., Webb, C.O. & Salamin, N. (2006) Exotic taxa less related to native species are more invasive. Proceedings of the National Academy of Sciences of the U.S.A., 103, 58415845.Google Scholar
Strong, D.R. Jr. (1977) Epiphytic tree loads, tree falls and perennial forest disruption: a mechanism for maintaining higher tree species richness in the tropics without animals. Journal of Biogeography, 4, 215218.Google Scholar
Strong, D.R. Jr. (1983) Natural variability and the manifold mechanisms of ecological communities. American Naturalist, 122, 636660.Google Scholar
Strong, D.R. Jr., Lzyska, L.A. & Simberloff, D.S. (1979) Tests of community-wide character displacement against null hypotheses. Evolution, 33, 897913.Google Scholar
Stubbs, W.J. & Wilson, J.B. (2004) Evidence for limiting similarity in a sand dune community. Journal of Ecology, 92, 557567.Google Scholar
Stylinski, C.D. & Allen, E.B. (1999) Lack of native species recovery following severe exotic disturbance in southern Californian shrublands. Journal of Applied Ecology, 36, 544554.Google Scholar
Sugihara, G. (1980) Minimal community structure: an explanation of species abundance patterns. American Naturalist, 116, 770787.Google Scholar
Sutherland, J.P. (1974) Multiple stable points in natural communities. American Naturalist, 108, 859873.Google Scholar
Suzuki, S.N. & Suzuki, R.O. (2011) Distance-dependent shifts in net effects by an unpalatable nettle on a palatable plant species. Acta Oecologica, 37, 386392.Google Scholar
Sykes, M.T., van der Maarel, E., Peet, R.K. & Willems, J.H. (1994) High species mobility in species-rich plant communities – an intercontinental comparison. Folia Geobotanica et Phytotaxonomica, 29, 439448.Google Scholar
Symstad, A.J. (2000) A test of the effects of functional group richness and composition on grassland invasibility. Ecology, 81, 99109.Google Scholar
Tahvanainen, J.O. & Root, R.B. (1972) The influence of vegetational diversity on the population ecology of a specialized herbivore, Phyllotreta cruciferae (Coleoptera: Chrysomelidae). Oecologia, 10, 321346.Google Scholar
Takeuchi, Y. & Nakashizuka, T. (2007) Effect of distance and density on seed/seedling fate of two dipterocarp species. Forest Ecology and Management, 247, 167174.Google Scholar
Tamm, C.O. (1964) Growth of Hylocomium splendens in relation to tree canopy. Bryologist, 67, 423426.Google Scholar
Tamme, R., Gazol, A., Price, J., Hiiesalu, I. & Pärtel, M. (2016) Co-occurring grassland species vary in their responses to fine-scale soil heterogeneity. Journal of Vegetation Science, 27, 10121022.Google Scholar
Tan, J.Q., Pu, Z.C., Ryberg, W.A. & Jiang, L. (2012) Species phylogenetic relatedness, priority effects, and ecosystem functioning. Ecology, 93, 11641172.Google Scholar
Tangney, R.S., Wilson, J.B. & Mark, A.F. (1990) Bryophyte island biogeography: a study in Lake Manapouri, New Zealand. Oikos, 59, 2126.Google Scholar
Tansley, A.G. (1917) On competition between Galium saxatile L. (G. hercynicum Weig.) and Galium sylvestre Poll. (G. asperum Schreb.) on different types of soil. Journal of Ecology, 5, 173179.Google Scholar
Tansley, A.G. (1920) The classification of vegetation and the concept of development. Journal of Ecology, 8, 118149.Google Scholar
Tansley, A.G. (1939) The British Islands and Their Vegetation. Cambridge University Press, Cambridge, UK.Google Scholar
Tansley, A.G. & Adamson, R.S. (1925) Studies of the vegetation of the English chalk: III. The chalk grasslands of Hampshire-Sussex border. Journal of Ecology, 13, 177223.Google Scholar
Tarroux, E. & DesRochers, A. (2011) Effect of natural root grafting on growth response of Jack Pine (Pinus banksiana; pinaceae). American Journal of Botany, 98, 967974.Google Scholar
Taylor, P.J. (1988) The construction and turnover of complex community models having generalized Lotka-Volterra dynamics. Journal of Theoretical Biology, 135, 569588.Google Scholar
Temperton, V.M., Hobba, R.J., Nuttle, T., Fattorini, M. & Halle, S. (1999) The search for ecological assembly rules and its relevance to restoration ecology. In Assembly Rules and Restoration Ecology. Bridging the Gap Between Theory and Practice. (eds Temperton, V.M., Hobbs, R.J., Nuttle, T. & Halle, S.), p. 464. Island Press, Washington DC.Google Scholar
terHorst, C.P., Miller, T.E. & Powell, E. (2010) When can competition for resources lead to ecological equivalence? Evolutionary Ecology Research, 12, 843854.Google Scholar
Teste, F.P., Simard, S.W., Durall, D.M., Guy, R.D. & Berch, S.M. (2010) Net carbon transfer between Pseudotsuga menziesii var. glauca seedlings in the field is influenced by soil disturbance. Journal of Ecology, 98, 429439.Google Scholar
Thébaud, C. & Simberloff, D. (2001) Are plants really larger in their introduced ranges? American Naturalist, 157, 231236.Google Scholar
Theimer, T.C. & Gehring, C.A. (1999) Effects of a litter-disturbing bird species on tree seedling germination and survival in an Australian tropical rain forest. Journal of Tropical Ecology, 15, 737749.Google Scholar
Thies, W. & Kalko, E.K.V. (2004) Phenology of neotropical pepper plants (Piperaceae) and their association with their main dispersers, two short-tailed fruit bats, Carollia perspicillata and C. castanea (Phyllostomidae). Oikos, 104, 362376.Google Scholar
Thijs, K.W., Brys, R., Verboven, H.A.F. & Hermy, M. (2012) The influence of an invasive plant species on the pollination success and reproductive output of three riparian plant species. Biological Invasions, 14, 355365.Google Scholar
Thomas, K.R., Munro, C.L., Graeme, C.B., Steel, J.B. & Wilson, J.B. (1999) Application and evaluation of Dale’s non-parametric method for detecting community structure through zonation. Oikos, 84, 261265.Google Scholar
Thomas, S.C. & Bazzaz, F.A. (1999) Asymptotic height as a predictor of photosynthetic characteristics in Malaysian rain forest trees. Ecology, 80, 16071622.Google Scholar
Thomas, W.R. & Pomerantz, M.J. (1981) Feasibility and stability in community dynamics. American Naturalist, 117, 381385.Google Scholar
Thompson, J.N. & Willson, M.F. (1979) Evolution of temperate fruit/bird interactions: phenological strategies. Evolution, 33, 973982.Google Scholar
Thompson, K., Hodgson, J.G., Grime, J.P. & Burke, M.J.W. (2001) Plant traits and temporal scale: evidence from a 5-year invasion experiment using native species. Journal of Ecology, 89, 10541060.Google Scholar
Thompson, L. & Harper, J.L. (1988) The effect of grasses on the quality of transmitted radiation and its influence on the growth of white clover Trifolium repens. Oecologia, 75, 343347.Google Scholar
Thórhallsdóttir, T.E. (1990) The dynamics of five grasses and white clover in a simulated mosaic sward. Journal of Ecology, 78, 909923.Google Scholar
Thorpe, A.S., Thelen, G.C., Diaconu, A. & Callaway, R.M. (2009) Root exudate is allelopathic in invaded community but not in native community: field evidence for the novel weapons hypothesis. Journal of Ecology, 97, 641645.Google Scholar
Thrall, P.H. & Jarosz, A.M. (1994) Host-pathogen dynamics in experimental populations of Silene alba and Ustilago violacea. II. Experimental tests of theoretical models. Journal of Ecology, 82, 561570.Google Scholar
Tilman, D. (1977) Resource competition between planktonic algae: an experimental and theoretical approach. Ecology, 58, 338348.Google Scholar
Tilman, D. (1981) Tests of resource competition theory using four species of Lake Michigan algae. Ecology, 62, 802815.Google Scholar
Tilman, D. (1982) Resource Competition and Community Structure. Princeton University Press, Princeton, NJ.Google Scholar
Tilman, D. (1986) Nitrogen-limited growth in plants from different successional stages. Ecology, 67, 555563.Google Scholar
Tilman, D. (1987) Secondary succession and the pattern of plant dominance along experimental nitrogen gradients. Ecological Monographs, 57, 189214.Google Scholar
Tilman, D. (1988) Plant Strategies and the Dynamics and Structure of Plant Communities. Princeton University Press, Princeton, NJ.Google Scholar
Tilman, D. (1993) Species richness of experimental productivity gradients: how important is colonization limitation? Ecology, 74, 21792191.Google Scholar
Tilman, D. (1994) Competition and biodiversity in spatially structured habitats. Ecology, 75, 216.Google Scholar
Tilman, D. (1997) Community invasibility, recruitment limitation, and grassland biodiversity. Ecology, 78, 8192.Google Scholar
Tilman, D. & Cowan, M.L. (1989) Growth of old field herbs on a nitrogen gradient. Functional Ecology, 3, 425438.Google Scholar
Tilman, D. & Lehman, C. (2001) Human-caused environmental change: impacts on plant diversity and evolution. Proceedings of the National Academy of Sciences of the U.S.A., 98, 54335440.Google Scholar
Tilman, D., Reich, P.B., Knops, J., Wedin, D., Mielke, T. & Lehman, C. (2001) Diversity and productivity in a long-term grassland experiment. Science, 294, 843845.Google Scholar
Tilman, D., Reich, P.B. & Knops, J.M.H. (2006) Biodiversity and ecosystem stability in a decade-long grassland experiment. Nature, 441, 629632.Google Scholar
Tilman, D. & Wedin, D. (1991a) Plant traits and resource reduction for five grasses growing on a nitrogen gradient. Ecology, 72, 685700.Google Scholar
Tilman, D. & Wedin, D. (1991b) Dynamics of nitrogen competition between successional grasses. Ecology, 72, 10381049.Google Scholar
Timberlake, J.R. & Calvert, G.M. (1993) Preliminary root atlas of Zimbabwe. Zimbabwe Bulletin of Forestry Research, 10, 190.Google Scholar
Titman [Tilman], D. (1976) Ecological competition between algae: experimental confirmation of resource-based competition theory. Science, 192, 463465.Google Scholar
Tokeshi, M. (1996) Species coexistence and abundance: patterns and processes. In Biodiversity: An Ecological Perspective (eds Abe, T., Levin, S.A. & Higashi, M.), pp. 3555. Springer, New York.Google Scholar
Tokita, K. & Yasutomi, A. (2003) Emergence of a complex and stable network in a model ecosystem with extinction and mutation. Theoretical Population Biology, 63, 131146.Google Scholar
Tongway, C.H., Valentin, C. & Seghieri, J. (2001) Banded Vegetation Patterning in Arid and Semi-Arid Environments. Springer, New York.Google Scholar
Torti, S.D., Coley, P.D. & Kursar, T.A. (2001) Causes and consequences of monodominance in tropical lowland forests. American Naturalist, 157, 141153.Google Scholar
Touchan, R., Swetnam, T.W. & Grissino-Mayer, H.D. (1995) Effects of livestock grazing on pre-settlement fire regimes in New Mexico. In Proceedings: Symposium on Fire in Wilderness and Park Management (eds Brown, J.K., Mutsch, R.W., Spoon, C.W. & Wakimoto, R.H.), pp. 268272. Intermountain Research Station, Ogden.Google Scholar
Tregonning, K. & Roberts, A. (1979) Complex systems which evolve towards homeostasis. Nature, 281, 563564.Google Scholar
Tucker, C.M. & Fukami, T. (2014) Environmental variability counteracts priority effects to facilitate species coexistence: evidence from nectar microbes. Proceedings of the Royal Society of London B, 281, 32637.Google Scholar
Tuomi, J., Augner, M. & Nilsson, P. (1994) A dilemma of plant defences: is it really worth killing the herbivore? Journal of Theoretical Biology, 170, 427430.Google Scholar
Turkington, R. & Harper, J.L. (1979a) The growth, distribution and neighbour relationships of Trifolium repens in a permanent pasture. I. Ordination, pattern and contact. Journal of Ecology, 67, 201218.Google Scholar
Turkington, R. & Harper, J.L. (1979b) The growth, distribution and neighbour relationships of Trifolium repens in a permanent pasture. IV. Fine-scale biotic differentiation. Journal of Ecology, 67, 245254.Google Scholar
Turkington, R. & Mehrhoff, L.A. (1990) The role of competition in structuring pasture communities. In Perspectives on Plant Competition (eds Grace, J.B. & Tilman, D.), pp. 307340. Academic Press, San Diego.Google Scholar
Turley, M.C. & Ford, E.D. (2011) Detecting bimodality in plant size distributions and its significance for stand development and competition. Oecologia, 167, 9911003.Google Scholar
Turnbull, L.A., Rees, M. & Crawley, M.J. (1999) Seed mass and the competition/colonization trade-off: a sowing experiment. Journal of Ecology, 87, 899912.Google Scholar
Turnbull, M.H. & Yates, D.J. (1993) Seasonal variation in the red far-red ratio and photon flux density in an Australian sub-tropical rainforest. Agricultural and Forest Meteorology, 64, 111127.Google Scholar
Uhl, C. (1982) Tree dynamics in a species rich tierra firme forest in Amazonia, Venezuela. Acta Cientifica Venezolana, 33, 7177.Google Scholar
Ulrich, W. (2004) Species co-occurrences and neutral models: reassessing J. M. Diamond’s assembly rules. Oikos, 107, 603609.Google Scholar
Underwood, G.J.C. & Paterson, D.M. (1993) Recovery of intertidal benthic diatoms after biocide treatment and associated sediment dynamics. Journal of the Marine Biological Association of the United Kingdom, 73, 2545.Google Scholar
Urban, M.C. (2011) The evolution of species interactions across natural landscapes. Ecology Letters, 14, 723732.Google Scholar
Uriarte, M., Condit, R., Canham, C.D. & Hubbell, S.P. (2004) A spatially explicit model of sapling growth in a tropical forest: does the identity of neighbours matter? Journal of Ecology, 92, 348360.Google Scholar
Usher, M.B. (1987) Modelling successional processes in ecosystems. In Colonization, Succession and Stability (eds Gray, A.J., Crawley, M.J. & Edwards, P.J.), pp. 3156. Blackwell, Oxford.Google Scholar
Usinowicz, J., Wright, S.J. & Ives, A.R. (2012) Coexistence in tropical forests through asynchronous variation in annual seed production. Ecology, 93, 20732084.Google Scholar
Vaidyanathan, L.V., Drew, M.C. & Nye, P.H. (1968) The measurement and mechanism of ion diffusion in soils. IV. The concentration dependence of diffusion concentrations of potassium in soils at a range of moisture levels and a method for the estimation of the differential diffusion coefficient at any concentration. Journal of Soil Science, 19, 94107.Google Scholar
Valiente-Banuet, A. & Verdú, M. (2008) Temporal shifts from facilitation to competition occur between closely related taxa. Journal of Ecology, 96, 489494.Google Scholar
Valone, T.J. & Barber, N.A. (2008) An empirical evaluation of the insurance hypothesis in diversity-stability models. Ecology, 89, 522531.Google Scholar
Valone, T.J., Meyer, M., Brown, J.H. & Chew, R.M. (2002) Timescale of perennial grass recovery in desertified arid grasslands following livestock removal. Conservation Biology, 16, 9951002.Google Scholar
van de Koppel, J., Herman, P.M.J., Thoolen, P. & Heip, C.H.R. (2001) Do alternate stable states occur in natural ecosystems? Evidence from a tidal flat. Ecology, 82, 34493461.Google Scholar
van de Koppel, J. & Prins, H.H.T. (1998) The importance of herbivore interactions for the dynamics of African savanna woodlands: an hypothesis. Journal of Tropical Ecology, 14, 565576.Google Scholar
van de Koppel, J., Rietkerk, M., van Langevelde, F., Kumar, L., Klausmeier, C.A., Fryxell, J.M., Hearne, J.W., van Andel, J., de Ridder, N., Skidmore, A., Stroosnijder, L. & Prins, H.H. (2002) Spatial heterogeneity and irreversible vegetation change in semiarid grazing systems. American Naturalist, 159, 209218.Google Scholar
van den Bergh, J.P. & Elberse, W.T. (1962) Competition between Lolium perenne L. and Anthoxanthum odoratum L. at two levels of phosphate and potash. Journal of Ecology, 50, 8795.Google Scholar
van der Heide, T., Smolders, A., Rijkens, B., van Nes, E.H., van Katwijk, M.M. & Roelofs, J. (2008) Toxicity of reduced nitrogen in eelgrass (Zostera marina) is highly dependent on shoot density and pH. Oecologia, 158, 411419.Google Scholar
van der Heide, T., van Nes, E.H., van Katwijk, M.M., Scheffer, M., Hendriks, A.J. & Smolders, A.J.P. (2010) Alternative stable states driven by density-dependent toxicity. Ecosystems, 13, 841850.Google Scholar
van der Maarel, E. & Sykes, M.T. (1997) Rates of small-scale species mobility in alvar limestone grassland. Journal of Vegetation Science, 8, 199208.Google Scholar
Van der Putten, W.H. & Peters, B.A.M. (1997) How soil pathogens may affect plant competition. Ecology, 78, 17851795.Google Scholar
Van der Stoel, C.D., Van Der Putten, W.H. & Duyts, H. (2002) Development of a negative soil plant feedback in the expansion zone of the clonal grass Ammophila arenaria following root formation and nematode colonisation. Journal of Ecology, 90, 978988.Google Scholar
van Donk, E. & van de Bund, W.J. (2002) Impact of submerged macrophytes including charophytes on phyto- and zooplankton communities: allelopathy versus other mechanisms. Aquatic Botany, 72, 261274.Google Scholar
van Gardingen, P. & Grace, J. (1991) Plants and wind. Advances in Botanical Research, 18, 192248.Google Scholar
van Kleunen, M. & Fischer, M. (2009) Release from foliar and floral fungal pathogen species does not explain the geographic spread of naturalized North American plants in Europe. Journal of Ecology, 97, 385392.Google Scholar
van Nes, E.H. & Scheffer, M. (2004) Large species shifts triggered by small forces. American Naturalist, 164, 255266.Google Scholar
van Nes, E.H., Scheffer, M., van den Berg, M.S. & Coops, H. (2002) Dominance of charophytes in eutrophic shallow lakes - when should we expect it to be an alternative stable state? Aquatic Botany, 72, 275296.Google Scholar
van Ruijven, J. & Berendse, F. (2007) Contrasting effects of diversity on the temporal stability of plant populations. Oikos, 116, 13231330.Google Scholar
Van Ruijven, J. & Berendse, F. (2009) Long-term persistence of a positive plant diversity-productivity relationship in the absence of legumes. Oikos, 118, 101106.Google Scholar
van Ruijven, J., De Deyn, G.B. & Berendse, F. (2003) Diversity reduces invasibility in experimental plant communities: the role of plant species. Ecology Letters, 6, 910918.Google Scholar
van Steenis, C.G.G.T. (1981) Rheophytes of the World. Sijhoff and Noordhoff, Alphen.Google Scholar
Vance, R.R. (1984) Interference competition and the coexistence of two competitors on a single limiting resource. Ecology, 65, 13491357.Google Scholar
Vandermeer, J.H. (1969) The competitive structure of communities: an experimental approach with protozoa. Ecology, 50, 362371.Google Scholar
Vandermeer, J.H. (1977) Notes on density dependence in Welfia georgii Wendl. ex Burnett (Palmae): a lowland rainforest species in Costa Rica. Brenesia, 10, 915.Google Scholar
Vandermeer, J.H., de la Cerda, I.G., Perfecto, I., Boucher, D., Ruiz, J. & Kaufmann, A. (2004) Multiple basins of attraction in a tropical forest: evidence for non-equilibrium community structure. Ecology, 85, 575579.Google Scholar
Vanderwall, S.B. (2001) The evolutionary ecology of nut dispersal. Botanical Review, 67, 74117.Google Scholar
Vanelslander, B., Dewever, A., Vanoostende, N., Kaewnuratchadasorn, P., Vanormelingen, P., Hendrickx, F., Sabbe, K. & Vyverman, W. (2009) Complementarity effects drive positive diversity effects on biomass production in experimental benthic diatom biofilms. Journal of Ecology, 97, 10751082.Google Scholar
Vanhinsberg, A. & Vantienderen, P. (1997) Variation in growth form in relation to spectral light quality (red/far-red ratio) in Plantago lanceolata L. in sun and shade populations. Oecologia, 111, 452459.Google Scholar
Vannette, R.L. & Fukami, T. (2014) Historical contingency in species interactions: towards niche-based predictions. Ecology Letters, 17, 115124.Google Scholar
Vasquez, E.C. & Meyer, G.A. (2011) Relationships among leaf damage, natural enemy release, and abundance in exotic and native prairie plants. Biological Invasions, 13, 621633.Google Scholar
Vasseur, D.A., Amarasekare, P., Rudolf, V.H.W. & Levine, J.M. (2011) Eco-evolutionary dynamics enable coexistence via neighbor-dependent selection. American Naturalist, 178, 96109.Google Scholar
Vázquez-Yanes, C., Orozco-Segovia, A., Rincón, E., Sánchez-Coronado, M.E., Huante, P., Toledo, J.R. & Barradas, V.L. (1990) Light beneath the litter in a tropical forest: effect on seed germination. Ecology, 71, 19521958.Google Scholar
Veblen, T.T. & Stewart, G.H. (1980) Comparison of forest structure and regeneration on Bench and Stewart Islands, New Zealand. New Zealand Journal of Ecology, 3, 5068.Google Scholar
Veblen, T.T. & Stewart, G.H. (1982) On the conifer regeneration gap in New Zealand: dynamics of Libocedrus bidwillii stands on South Island. Journal of Ecology, 70, 413436.Google Scholar
Vellend, M. (2010) Conceptual synthesis in community ecology. Quarterly Review of Biology, 85, 183206.Google Scholar
Vellend, M. (2016) The Theory of Ecological Communities. Princeton University Press, Princeton, NJ.Google Scholar
Veloz, S.D., Williams, J.W., Blois, J.L., He, F., OttoBliesner, B. & Liu, Z.Y. (2012) No-analog climates and shifting realized niches during the late quaternary: implications for 21st-century predictions by species distribution models. Global Change Biology, 18, 16981713.Google Scholar
Veresoglou, D.S. & Fitter, A.H. (1984) Spatial and temporal patterns of growth and nutrient uptake of five co-existing grasses. Journal of Ecology, 72, 259272.Google Scholar
Verhulst, J., Montaña, C., Mandujano, M.C. & Franco, M. (2008) Demographic mechanisms in the coexistence of two closely related perennials in a fluctuating environment. Oecologia, 156, 95105.Google Scholar
Vermeulen, P.J., Stuefer, J.F., Anten, N.P.R. & During, H.J. (2009) Carbon gain in the competition for light between genotypes of the clonal herb Potentilla reptans. Journal of Ecology, 97, 508517.Google Scholar
Verrecchia, E., Yair, A., Kidron, G.J. & Verrecchia, K. (1995) Physical properties of the psammophile cryptogamic crust and their consequences to the water regime of sandy soils, north-western Negev Desert, Israel. Journal of Arid Environments, 29, 427437.Google Scholar
Viketoft, M. (2008) Effects of six grassland plant species on soil nematodes: A glasshouse experiment. Soil Biology & Biochemistry, 40, 906915.Google Scholar
Viketoft, M., Palmborg, C., Sohlenius, B., Huss-Danell, K. & Bengtsson, J. (2005) Plant species effects on soil nematode communities in experimental grasslands. Applied Soil Ecology, 30, 90103.Google Scholar
Vilà, M., Maron, J.L. & Marco, L. (2005) Evidence for the enemy release hypothesis in Hypericum perforatum. Oecologia, 142, 474479.Google Scholar
Vile, D., Shipley, B. & Garnier, E. (2006) A structural equation model to integrate changes in functional strategies during old-field succession. Ecology, 87, 504517.Google Scholar
Vitousek, P.M. (2004) Nutrient Cycling and Limitation: Hawai’i as a Model System. Princeton University Press, Princeton, NJ.Google Scholar
Volkov, I., Banavar, J.R., Hubbell, S.P. & Maritan, A. (2009) Inferring species interactions in tropical forests. Proceedings of the National Academy of Sciences of the U.S.A., 106, 1385413859.Google Scholar
Von Holle, B., Delcourt, H.R. & Simberloff, D. (2003) The importance of biological inertia in plant community resistance to invasion. Journal of Vegetation Science, 14, 425432.Google Scholar
Von Holle, B. & Simberloff, D. (2004) Testing Fox’s assembly rule: does plant invasion depend on recipient community structure? Oikos, 105, 551563.Google Scholar
Vos, V.C.A., van Ruijven, J., Berg, M.P., Peeters, E. & Berendse, F. (2013) Leaf litter quality drives litter mixing effects through complementary resource use among detritivores. Oecologia, 173, 269280.Google Scholar
Vranckx, G. & Vandelook, F. (2012) A season- and gap-detection mechanism regulates seed germination of two temperate forest pioneers. Plant Biology, 14, 481490.Google Scholar
Wade, M.J. (1978) A critical review of the models of group selection. Quarterly Review of Biology, 53, 101114.Google Scholar
Walker, B.H. (1992) Biodiversity and ecological redundancy. Conservation Biology, 6, 1823.Google Scholar
Walker, B.H. (1993) Rangeland ecology: understanding and managing change. Ambio, 22, 8087.Google Scholar
Walker, B.H., Kinzig, A. & Langridge, J. (1999) Plant attribute diversity, resilience, and ecosystem function: the nature and significance of dominant and minor species. Ecosystems, 2, 95113.Google Scholar
Walker, D. (1970) Direction and rate in some British postglacial hydroseres. In Studies in the Postglacial History of the British Isles (eds Walker, D. & West, R.G.), pp. 117140. Cambridge University Press, Cambridge, UK.Google Scholar
Walker, J., Thompson, C.H., Fergus, I.F. & Tunstall, B.R. (1981) plant succession and soil development in coastal dunes of Eastern Australia. In Forest Succession: Concepts and Applications (eds West, D.C., Shugart, H.H. & Botkin, D.B.), pp. 107131. Springer, New York.Google Scholar
Walker, L.R. & Chapin, F.S.I. (1986) Physiological controls over seedling growth in primary succession on an Alaskan floodplain. Ecology, 67, 15081523.Google Scholar
Walker, L.R., Clarkson, B.D., Silvester, W.B. & Clarkson, B.R. (2003a) Colonization dynamics and facilitative impacts of a nitrogen-fixing shrub in primary succession. Journal of Vegetation Science, 14, 277290.Google Scholar
Walker, L.R., Thompson, D.B. & Landau, F.H. (2001) Experimental manipulations of fertile islands and nurse plant effects in the Mojave Desert, USA. Western North American Naturalist, 61, 2535.Google Scholar
Walker, N.A., Henry, H.A.L., Wilson, D.J. & Jefferies, R.L. (2003b) The dynamics of nitrogen movement in an Arctic salt marsh in response to goose herbivory: a parameterized model with alternate stable states. Journal of Ecology, 91, 637650.Google Scholar
Walker, S. & Wilson, J.B. (2002) Tests for nonequilibrium, instability and stabilizing processes in semiarid plant communities. Ecology, 83, 809822.Google Scholar
Walker, S., Wilson, J.B., Steel, J.B., Rapson, G.L., Smith, B., King, W.M. & Cottam, Y.H. (2003c) Properties of ecotones: evidence from five coastal ecotones. Journal of Vegetation Science, 14, 579590.Google Scholar
Walley, K.A., Khan, M.S.I. & Bradshaw, A.D. (1974) The potential for evolution of heavy-metal tolerance in plants. I. Copper and zinc tolerance in Agrostis tenuis. Heredity, 32, 309319.Google Scholar
Wang, H.C., Wang, S.F., Lin, K.C., Shaner, P.J. & Lin, T.C. (2013) Litterfall and element fluxes in a natural hardwood forest and a Chinese-fir plantation experiencing frequent typhoon disturbance in central Taiwan. Biotropica, 45, 541548.Google Scholar
Wang, P., Stieglitz, T., Zhou, D.W. & Cahill, J.F. (2010) Are competitive effect and response two sides of the same coin, or fundamentally different? Functional Ecology, 24, 196207.Google Scholar
Wang, Y.-H., He, W.-M., Dong, M., Yu, F.-H., Zhang, L.-L., Cui, Q.-G. & Chu, Y. (2008) Effects of shaking on the growth and mechanical properties of Hedysarum laeve may be independent of water regimes. International Journal of Plant Sciences, 169, 503508.Google Scholar
Ward, C.M. (1988) Marine terraces of the Waitutu district and their relation to the late Cenozoic tectonics of the southern Fiordland region. Journal of the Royal Society of New Zealand, 18, 128.Google Scholar
Wardle, D.A. (2001) Experimental demonstration that plant diversity reduces invasibility. Evidence of a biological mechanism or a consequence of sampling effect? Oikos, 95, 161170.Google Scholar
Wardle, D.A. (2002) Communities and Ecosystems: Linking the Aboveground and Belowground Components. Princeton University Press, Princeton, NJ.Google Scholar
Wardle, D.A., Walker, L.R. & Bardgett, R.D. (2004) Ecosystem properties and forest decline in contrasting long-term chronosequences. Science, 305, 509513.Google Scholar
Wardle, P. (1964) Facets of the distribution of forest vegetation in New Zealand. New Zealand Journal of Botany, 2, 352366.Google Scholar
Warming, E. (1909) Oecology of Plants: an Introduction to the Study of Plant-Communities. Oxford University Press, Oxford.Google Scholar
Waser, N.M. & Real, L.A. (1979) Effective mutualism between sequentially flowering plant-species. Nature, 281, 670672.Google Scholar
Watkins, A.J. & Wilson, J.B. (1992) Fine-scale community structure of lawns. Journal of Ecology, 80, 1524.Google Scholar
Watkins, A.J. & Wilson, J.B. (1994) Plant community structure, and its relation to the vertical complexity of communities: dominance/diversity and spatial rank consistency. Oikos, 70, 9198.Google Scholar
Watkins, A.J. & Wilson, J.B. (2003) Local texture convergence: a new approach to seeking assembly rules. Oikos, 102, 525532.Google Scholar
Watkinson, A.R. (1985) On the abundance of plants along an environmental gradient. Journal of Ecology, 73, 569578.Google Scholar
Watson, I.W., Burnside, D.G. & Holm, A.M. (1996) Event-driven or continuous; which is the better model for managers? Rangeland Journal, 18, 351369.Google Scholar
Watt, A.S. (1947) Pattern and process in the plant community. Journal of Ecology, 35, 122.Google Scholar
Watt, A.S. (1955) Bracken versus heather, a study in plant sociology. Journal of Ecology, 43, 490506.Google Scholar
Watt, A.S. (1960) Population changes in acidophilous grass-heath in Breckland, 1936–1957. Journal of Ecology, 48, 605629.Google Scholar
Watt, A.S. (1981) Further observations on the effects of excluding rabbits from Grassland A in East Anglian Breckland: the pattern of change and factors affecting it (1936–1973). Journal of Ecology, 69, 509536.Google Scholar
Weaver, J.E. & Clements, F.E. (1929) Plant Ecology. McGraw-Hill, New York.Google Scholar
Wedin, D. & Tilman, D. (1993) Competition among grasses along a nitrogen gradient – initial conditions and mechanisms of competition. Ecological Monographs, 63, 199229.Google Scholar
Weiher, E., Clarke, G.D.P. & Keddy, P.A. (1998) Community assembly rules, morphological dispersion, and the coexistence of plant species. Oikos, 81, 309322.Google Scholar
Weiner, J., Wright, D.B. & Castro, S. (1997) Symmetry of below-ground competition between Kochia scoparia individuals. Oikos, 79, 8591.Google Scholar
Weisner, S.E.B. (1993) Long term competitive displacement of Typha latifolia by T. angustifolia in a eutrophic lake. Oecologia, 94, 451456.Google Scholar
Welbank, P.J. (1963) Toxin production during decay of Agropyron repens (couch grass) and other species. Weed Research, 3, 205214.Google Scholar
Went, F.W. (1955) The ecology of desert plants. Scientific American, 192, 6875.Google Scholar
Wheelwright, N.T. (1985) Competition for dispersers, and the timing of flowering and fruiting in a guild of tropical trees. Oikos, 44, 465477.Google Scholar
Whelan, R.J. (1995) The Ecology of Fire. Cambridge University Press, Cambridge, UK.Google Scholar
Whisenant, S.G. & Wagstaff, F.J. (1995) Successional trajectories of a grazed salt desert shrubland. Vegetatio, 94, 133140.Google Scholar
White, A.J., Wratten, S.D., Berry, N.A. & Weigmann, U. (1995) Habitat manipulation to enhance biological control of Brassica pests by hover flies (Diptera: Syrphidae). Journal of Economic Entomology, 88, 11711176.Google Scholar
White, J.A. & Whitham, T.G. (2000) Associational susceptibility of cottonwood to a box elder herbivore. Ecology, 81, 17951803.Google Scholar
White, P.S. (1979) Pattern, process, and natural disturbance in vegetation. Botanical Review, 45, 229299.Google Scholar
Whitford, W.G. (2002) Ecology of Desert Systems. Academic Press, San Diego.Google Scholar
Whitford, W.G., Anderson, J. & Rice, P.M. (1997) Stemflow contribution to the ‘fertile island’ effect in creosotebush, Larrea tridentata. Journal of Arid Environments, 35, 451457.Google Scholar
Whitham, T.G., Young, W.P., Martinsen, G.D., Gehring, C.A., Schweitzer, J.A., Shuster, S.M., Wimp, G.M., Fischer, D.G., Bailey, J.K., Lindroth, R.L., Woolbright, S. & Kuske, C.R. (2003) Community and ecosystem genetics: a consequence of the extended phenotype. Ecology, 84, 559573.Google Scholar
Whittaker, R.H. (1960) Vegetation of the Siskiyou Mountains, Oregon and California. Ecological Monographs, 30, 279338.Google Scholar
Whittaker, R.H. (1965) Dominance and diversity in land plant communities. Science, 147, 250260.Google Scholar
Whittaker, R.H. (1967) Gradient analysis of vegetation. Biological Reviews, 49, 207264.Google Scholar
Whittaker, R.H. (1975a) Communities and Ecosystems, 2nd edn. Macmillan, New York.Google Scholar
Whittaker, R.H. (1975b) The design and stability of some plant communities. In Unifying Concepts in Ecology (eds van Dobben, W.H. & Lowe-McConnell, R.H.), pp. 169181. Junk, The Hague.Google Scholar
Whittaker, R.H. (1977) Evolution of species diversity in land communities. Evolutionary Biology, 10, 167.Google Scholar
Whittaker, R.H. & Levin, S.A. (1975) Niche: Theory and Application. Dowden, Hutchinson and Ross, Stroudsburg, PA.Google Scholar
Whittaker, R.H. & Woodwell, G.M. (1972) Evolution of natural communities. In Ecosystem Structure and Function (ed. Wiens, J.A.), pp. 137155. Oregon State University Press, Corvallis.Google Scholar
Whittington, W.J. & O’Brien, T.A. (1968) A comparison of yields from plots sown with a single species or a mixture of grass species. Journal of Applied Ecology, 5, 209213.Google Scholar
Wiens, J.A. (1989) The Ecology of Bird Communities, Vol. 1, Foundations and Patterns. Cambridge University Press, Cambridge, UK.Google Scholar
Wilkinson, M.J. & Stace, C.A. (1991) A new taxonomic treatment of the Festuca ovina L. aggregate (Poaceae) in the British Isles. Biological Journal of the Linnean Society, 106, 347397.Google Scholar
Willby, N.J., Abernethy, V.J. & Demars, B.O.L. (2000) Attribute-based classification of European hydrophytes and its relationship to habitat utilization. Freshwater Biology, 43, 4374.Google Scholar
Williams, H.T.P. & Lenton, T.M. (2008) Environmental regulation in a network of simulated microbial ecosystems. Proceedings of the National Academy of Sciences of the U.S.A., 105, 1043210437.Google Scholar
Williams, J.W., Shuman, B. & Webb, T. (2001) Dissimilarity analyses of late-Quaternary vegetation and climate in eastern North America. Ecology, 82, 33463362.Google Scholar
Williamson, G.B. (1990) Allelopathy, Koch’s postulates and the neck riddle. In Perspectives on Plant Competition (eds Grace, J.B. & Tilman, D.), pp. 143162. Academic Press, San Diego.Google Scholar
Willig, M.R., Presley, S.J., Bloch, C.P., Castro-Arellano, I., Cisneros, L.M., Higgins, C.L. & Klingbeil, B.T. (2011) Tropical metacommunities along elevational gradients: effects of forest type and other environmental factors. Oikos, 120, 14971508.Google Scholar
Wilsey, B.J. & Polley, H.W. (2002) Reductions in grassland species evenness increase dicot seedling invasion and spittle bug infestation. Ecology Letters, 5, 676684.Google Scholar
Wilson, D.S. (1992) Complex interactions in metacommunities, with implications for biodiversity and higher levels of selection. Ecology, 73, 19842000.Google Scholar
Wilson, E.O. (1961) The nature of the taxon cycle in the Melanesian ant fauna. American Naturalist, 95, 169193.Google Scholar
Wilson, J.B. (1987) Group selection in plant populations. Theoretical and Applied Genetics, 74, 493502.Google Scholar
Wilson, J.B. (1988a) A review of evidence on the control of shoot:root ratio, in relation to models. Annals of Botany, 61, 433449.Google Scholar
Wilson, J.B. (1988b) The effect of initial advantage on the course of competition. Oikos, 51, 1924.Google Scholar
Wilson, J.B. (1988c) Shoot competition and root competition. Journal of Applied Ecology, 25, 279296.Google Scholar
Wilson, J.B. (1988d) Community structure in the flora of islands in Lake Manapouri. Journal of Ecology, 76, 10301042.Google Scholar
Wilson, J.B. (1989a) Relations between native and exotic plant guilds in the Upper Clutha, New Zealand. Journal of Ecology, 77, 223235.Google Scholar
Wilson, J.B. (1989b) A null model of guild proportionality, applied to stratification of a New Zealand temperate rain forest. Oecologia, 80, 263267.Google Scholar
Wilson, J.B. (1990) Mechanisms of species coexistence: twelve explanations for Hutchinson’s ‘Paradox of the Plankton’: evidence from New Zealand plant communities. New Zealand Journal of Ecology, 13, 1742.Google Scholar
Wilson, J.B. (1991) Methods for fitting dominance/diversity curves. Journal of Vegetation Science, 2, 3546.Google Scholar
Wilson, J.B. (1993) Macronutrient (NPK) toxicity and interactions in the grass Festuca ovina. Journal of Plant Nutrition, 16, 11511159.Google Scholar
Wilson, J.B. (1994a) The ‘Intermediate disturbance hypothesis’ of species coexistence is based on patch dynamics. New Zealand Journal of Ecology, 18, 176181.Google Scholar
Wilson, J.B. (1994b) Who makes the assembly rules? Journal of Vegetation Science, 5, 275278.Google Scholar
Wilson, J.B. (1995) Fox and Brown’s ‘random data sets’ are not random. Oikos, 74, 543544.Google Scholar
Wilson, J.B. (1997) An evolutionary perspective on the ‘death hormone’ hypothesis in plants. Physiologia Plantarum, 99, 511516.Google Scholar
Wilson, J.B. (1999a) Assembly rules in plant communities. In Ecological Assembly Rules: Perspectives, Advances and Retreats (eds Weiher, E. & Keddy, P.A.), pp. 130164. Cambridge University Press, Cambridge, UK.Google Scholar
Wilson, J.B. (1999b) Guilds, functional types and ecological groups. Oikos, 86, 507522.Google Scholar
Wilson, J.B. (2002) The ‘emergent property’ of Anand and Li is a mathematical artefact. Emergent properties do not exist. Community Ecology, 3, 4748.Google Scholar
Wilson, J.B. (2003) The deductive method in community ecology. Oikos, 101, 216218.Google Scholar
Wilson, J.B. (2011) Cover plus: ways of measuring plant canopies and the terms used for them. Journal of Vegetation Science, 22, 197206.Google Scholar
Wilson, J.B. & Agnew, A.D.Q. (1992) Positive-feedback switches in plant communities. Advances in Ecological Research, 23, 263336.Google Scholar
Wilson, J.B., Agnew, A.D.Q. & Partridge, T.R. (1994) Carr texture in Britain and New Zealand: community convergence compared with a null model. Journal of Vegetation Science, 5, 109116.Google Scholar
Wilson, J.B., Agnew, A.D.Q. & Sykes, M.T. (2004) Ecology or mythology? Are Whittaker’s “gradient analysis” curves reliable evidence of continuity in vegetation? Preslia, 76, 245253.Google Scholar
Wilson, J.B. & Allen, R.B. (1990) Deterministic versus Individualistic community structure: a test from invasion by Nothofagus menziesii in southern New Zealand. Journal of Vegetation Science, 1, 467474.Google Scholar
Wilson, J.B., Allen, R.B. & Lee, W.G. (1995c) An assembly rule in the ground and herbaceous strata of a New Zealand rainforest. Functional Ecology, 9, 6164.Google Scholar
Wilson, J.B. & Anderson, B.J. (2001) Species-pool relations: like a wooden light bulb? Folia Geobotanica, 36, 3544.Google Scholar
Wilson, J.B., Crawley, M.J., Dodd, M.E. & Silvertown, J. (1996c) Evidence for constraint on species coexistence in vegetation of the Park Grass experiment. Vegetatio, 124, 183190.Google Scholar
Wilson, J.B. & Gitay, H. (1995a) Limitations to species coexistence: evidence for competition from field observations, using a patch model. Journal of Vegetation Science, 6, 369376.Google Scholar
Wilson, J.B. & Gitay, H. (1995b) Community structure and assembly rules in a dune slack: variance in richness, guild proportionality, biomass constancy and dominance/diversity relations. Vegetatio, 116, 93106.Google Scholar
Wilson, J.B. & Gitay, H. (1999) Alternative classifications in the intrinsic guild structure of a New Zealand tussock grassland. Oikos, 86, 566572.Google Scholar
Wilson, J.B., Gitay, H. & Agnew, A.D.Q. (1987) Does niche limitation exist? Functional Ecology, 1, 391397.Google Scholar
Wilson, J.B., Gitay, H., Roxburgh, S.H., King, W.M. & Tangney, R.S. (1992c) Egler’s concept of ‘Initial Floristic Composition’ in succession – ecologists citing it don’t agree what it means. Oikos, 64, 591593.Google Scholar
Wilson, J.B., Gitay, H., Steel, J.B. & King, W.M. (1998) Relative abundance distributions in plant communities: the effects of species richness and of spatial scale. Journal of Vegetation Science, 9, 213220.Google Scholar
Wilson, J.B., Hubbard, J.C.E. & Rapson, G.L. (1988) A comparison of the realised niche relations of species in New Zealand and Britain. Oecologia, 76, 106110.Google Scholar
Wilson, J.B., James, R.E., Newman, J.E. & Myers, T.E. (1992a) Rock pool algae – species composition determined by chance. Oecologia, 91, 150152.Google Scholar
Wilson, J.B. & King, W.M. (1995) Human mediated switches as processes in landscape ecology. Landscape Ecology, 10, 191196.Google Scholar
Wilson, J.B. & Lee, W.G. (1989) Infiltration invasion. Functional Ecology, 3, 379380.Google Scholar
Wilson, J.B. & Lee, W.G. (1994) Niche overlap of congeners: a test using plant altitudinal distribution. Oikos, 69, 469475.Google Scholar
Wilson, J.B. & Lee, W.G. (2000) C-S-R Triangle theory: community-level predictions, tests, evaluation of criticisms, and relation to other theories. Oikos, 91, 7796.Google Scholar
Wilson, J.B. & Lee, W.G. (2012) Is New Zealand vegetation really problematic’? Dansereau’s puzzles revisited. Biological Reviews, 87, 367389.Google Scholar
Wilson, J.B. & Newman, E.I. (1987) Competition between upland grasses: root and shoot competition between Deschampsia flexuosa and Festuca ovina. Acta Oecologia Oecologia Generalis, 8, 501511.Google Scholar
Wilson, J.B., Peet, R.K., Dengler, J. & Partel, M. (2012) Plant species richness: the world records. Journal of Vegetation Science, 23, 796802.Google Scholar
Wilson, J.B. & Roxburgh, S.H. (1994) A demonstration of guild-based assembly rules for a plant community, and determination of intrinsic guilds. Oikos, 69, 267276.Google Scholar
Wilson, J.B. & Roxburgh, S.H. (2001) Intrinsic guild structure: determination from competition experiments. Oikos, 92, 189192.Google Scholar
Wilson, J.B., Roxburgh, S.H. & Watkins, A.J. (1992b) Limitation to plant species coexistence at a point: a study in a New Zealand lawn. Journal of Vegetation Science, 3, 711714.Google Scholar
Wilson, J.B. & Smith, B. (2001) Methods for testing for texture convergence using abundance data: a randomisation test and a method for comparing the shape of distributions. Community Ecology, 2, 5766.Google Scholar
Wilson, J.B., Spijkerman, E. & Huisman, J. (2007b) Is there really insufficient support for Tilman’s R* concept? A comment on Miller et al. American Naturalist, 169, 700706.Google Scholar
Wilson, J.B., Steel, J.B., Dodd, M.E., Anderson, B.J., Ullmann, I. & Bannister, P. (2000b) A test of community re-assembly using the exotic communities of New Zealand roadsides, in comparison to British roadsides. Journal of Ecology, 88, 757764.Google Scholar
Wilson, J.B., Steel, J.B., Newman, J.E. & King, W.M. (2000a) Quantitative aspects of community structure examined in a semi-arid grassland. Journal of Ecology, 88, 749756.Google Scholar
Wilson, J.B., Steel, J.B., Newman, J.E. & Tangney, R.S. (1995b) Are bryophyte communities different? Journal of Bryology, 18, 689705.Google Scholar
Wilson, J.B., Steel, J.B. & Steel, S.-L.K. (2007a) Do plants ever compete for space? Folia Geobotanica, 42, 431436.Google Scholar
Wilson, J.B. & Stubbs, W.J. (2012) Evidence for assembly rules: limiting similarity within a saltmarsh. Journal of Ecology, 100, 210221.Google Scholar
Wilson, J.B. & Sykes, M.T. (1988) Some tests for niche limitation by examination of species diversity in the Dunedin area, New Zealand. New Zealand Journal of Botany, 26, 237244.Google Scholar
Wilson, J.B., Sykes, M.T. & Peet, R.K. (1995a) Time and space in the community structure of a species-rich grassland. Journal of Vegetation Science, 6, 729740.Google Scholar
Wilson, J.B., Ullmann, I. & Bannister, P. (1996b) Do species assemblages ever recur? Journal of Ecology, 84, 471474.Google Scholar
Wilson, J.B. & Watkins, A.J. (1994) Guilds and assembly rules in lawn communities. Journal of Vegetation Science, 5, 591600.Google Scholar
Wilson, J.B., Wells, T.C.E., Trueman, I.C., Jones, G., Atkinson, M.D., Crawley, M.J., Dodd, M.E. & Silvertown, J. (1996a) Are there assembly rules for plant species abundance: an investigation in relation to soil resources and successional trends. Journal of Ecology, 84, 527538.Google Scholar
Wilson, J.B. & Whittaker, R.J. (1995) Assembly rules demonstrated in a saltmarsh community. Journal of Ecology, 83, 801807.Google Scholar
Wilson, S.D. & Keddy, P.A. (1985) Plant zonation on a shoreline gradient: physiological response curves of component species. Journal of Ecology, 73, 851860.Google Scholar
Wilson, S.D. & Tilman, D. (1991) Components of plant competition along an experimental gradient of nitrogen availability. Ecology, 72, 10501065.Google Scholar
Wilson, S.D. & Tilman, D. (1993) Plant competition and resource availability in response to disturbance and fertilization. Ecology, 74, 599611.Google Scholar
Woodell, S.R.J., Mooney, H.A. & Hill, A.J. (1969) The behaviour of Larrea divaricata (Creosote Bush) in response to rainfall in North California. Journal of Ecology, 57, 3744.Google Scholar
Wootton, J.T. (2005) Field parameterization and experimental test of the neutral theory of biodiversity. Nature, 433, 309312.Google Scholar
Wright, I.J., Reich, P.B., Westoby, M., Ackerly, D.D., Baruch, Z., Bongers, F., Cavender-Bares, J., Chapin, T., Cornelissen, J.H.C., Diemer, M., Flexas, J., Garnier, E., Groom, P.K., Gulias, J., Hikosaka, K., Lamont, B.B., Lee, T., Lee, W., Lusk, C., Midgley, J.J., Navas, M.-L., Niinemets, U., Oleksyn, J., Osada, N., Poorter, H., Poot, P., Prior, L., Pyankov, V.I., Roumet, C., Thomas, S.C., Tjoelker, M.G., Veneklaas, E.J. & Villar, R. (2004) The worldwide leaf economics spectrum. Nature, 428, 821827.Google Scholar
Wright, S.J. (2002) Plant diversity in tropical forests: a review of mechanisms of species coexistence. Oecologia, 130, 114.Google Scholar
Wright, S.J. & Calderon, O. (1995) Phylogenetic patterns among tropical flowering phenologies. Journal of Ecology, 83, 937948.Google Scholar
Wright, S.J., Muller-Landau, H.C., Condit, R. & Hubbell, S.P. (2003) Gap-dependent recruitment, realized vital rates, and size distributions of tropical trees. Ecology, 84, 31743185.Google Scholar
Wright, S.J. & van Schaik, C.P. (1994) Light and the phenology of tropical trees. American Naturalist, 143, 192199.Google Scholar
Wurzburger, N. & Hendrick, R.L. (2009) Plant litter chemistry and mycorrhizal roots promote a nitrogen feedback in a temperate forest. Journal of Ecology, 97, 528536.Google Scholar
Wyse, S.V. & Burns, B.R. (2013) Effects of Agathis australis (New Zealand kauri) leaf litter on germination and seedling growth differs among plant species. New Zealand Journal of Ecology, 37, 178183.Google Scholar
Yamazaki, M., Iwamoto, S. & Seiwa, K. (2009) Distance- and density-dependent seedling mortality caused by several diseases in eight tree species co-occurring in a temperate forest. Plant Ecology, 201, 181196.Google Scholar
Yan, X., Wang, Z., Huang, L., Wang, C., Hou, R., Xu, Z. & Quao, X. (2009) Research progress on electrical signals in higher plants. Progress in Natural Science, 19, 531541.Google Scholar
Yeaton, R.I. (1978) A cyclical relationship between Larrea tridentata and Opuntia leptocaulis in the Northern Chihuahuan Desert. Journal of Ecology, 66, 651656.Google Scholar
Yeaton, R.I. & Cody, M.L. (1976) Competition and spacing in plant communities in the northern Mohave Desert. Journal of Ecology, 64, 689696.Google Scholar
Yi, H.-S., Heil, M., Adame-Álvarez, R.M., Ballhorn, D.J. & Ryu, C.-M. (2009) Airborne induction and priming of plant defenses against a bacterial pathogen. Plant Physiology, 151, 21522161.Google Scholar
Yoder, C.K. & Nowak, R.S. (1999a) Hydraulic lift among native plant species in the Mojave Desert. Plant and Soil, 215, 93102.Google Scholar
Yoder, C.K. & Nowak, R.S. (1999b) Soil moisture extraction by evergreen and drought deciduous shrubs in the Mojave Desert during wet and dry years. Journal of Arid Environments, 42, 8196.Google Scholar
Yodzis, P. (1978) Competition for Space and the Structure of Ecological Communities. Springer, Berlin.Google Scholar
Yodzis, P. (1986) Competition, mortality, and community structure. In Community Ecology (eds Diamond, J.M. & Case, T.J.), pp. 480491. Harper and Row, New York.Google Scholar
Young, H.S., McCauley, D.J., Pollock, A. & Dirzo, R. (2014) Differential plant damage due to litterfall in palm-dominated forest stands in a Central Pacific atoll. Journal of Tropical Ecology, 30, 231236.Google Scholar
Young, T.P., Stanton, M.L. & Christian, C.E. (2003) Effects of natural and simulated herbivory on spine lengths of Acacia drepanolobium in Kenya. Oikos, 101, 171179.Google Scholar
Zhang, J., Zhao, H., Zhang, T., Zhao, X. & Drake, S. (2005) Community succession along a chronosequence of vegetation restoration on sand dunes in Horqin Sandy Land. Journal of Arid Environments, 62, 555566.Google Scholar
Zhang, Q.G. & Zhang, D.Y. (2007) Colonization sequence influences selection and complementarity effects on biomass production in experimental algal microcosms. Oikos, 116, 17481758.Google Scholar
Zhang, S.T. & Lamb, E.G. (2012) Plant competitive ability and the transitivity of competitive hierarchies change with plant age. Plant Ecology, 213, 1523.Google Scholar
Zimmerman, G.T. & Neuenschwander, L.F. (1984) Livestock grazing influences on community structure, fire intensity, and fire frequency within the Douglas-fir/Ninebark habitat type. Journal of Range Management, 37, 104110.Google Scholar
Zobel, K., Zobel, M. & Rosén, E. (1994) An experimental test of diversity maintenance mechanisms by a species removal experiment in a species-rich wooded meadow. Folia Geobotanica et Phytotaxonomica, 29, 449457.Google Scholar
Zohary, M. (1973) Geobotanical Foundations of the Middle East. Springer, Stuttgart.Google Scholar
Zonneveld, I.S. (1995) Vicinism and mass effect. Journal of Vegetation Science, 6, 441444.Google Scholar
Zuppinger-Dingley, D., Schmid, B., Petermann, J.S., Yadav, V., De Deyn, G.B. & Flynn, D.F.B. (2014) Selection for niche differentiation in plant communities increases biodiversity effects. Nature, 515, 108112.Google 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.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×