Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-20T01:35:00.082Z Has data issue: false hasContentIssue false

2 - Biogeography of ferns

Published online by Cambridge University Press:  05 June 2012

Michael Kessler
Affiliation:
University of Zürich
Klaus Mehltreter
Affiliation:
Instituto de Ecologia, A.C., Xalapa, Mexico
Lawrence R. Walker
Affiliation:
University of Nevada, Las Vegas
Get access

Summary

Key points

  1. 1. Biogeographical patterns of ferns and angiosperms are the result of a combination of vicariance and long distance dispersal, but due to their more effective dispersal via spores, the latter is more frequent among ferns. Therefore, fern species tend to have wider ranges and the relative number of fern species compared with seed plants is highest on remote, mountainous tropical islands such as Hawaii and the Mascarenes. Also, fern communities on different continents are more similar compositionally than those of seed plants.

  2. 2. Despite their potential for long distance spore dispersal, many fern species have localized ranges as a result of low frequency of successful long distance dispersal, habitat specialization, geographical isolation and competitive interactions between species.

  3. 3. Species richness of ferns follows a latitudinal gradient that peaks in the tropics, where ferns are especially diverse and abundant in wet habitats with moderate temperatures at elevations of about 1000–2500 m. On average, species in tropical mountains have elevational amplitudes of about 1000 m. The peak of endemism is located at higher elevations than that of species richness.

Introduction

Biogeography deals with the distribution patterns of species and communities, and their causal relationships with factors such as climate, soil and evolutionary history (Humboldt, 1805; Lomolino et al., 2006). Specific topics addressed by biogeographers include the sizes of geographical ranges and their spatial placement, the way individual species attain their distribution ranges (dispersal, extinction and vicariance), the distribution of species numbers (alpha diversity), changes in species composition (beta diversity) and the spatial distribution of species traits (macroecology).

Type
Chapter
Information
Fern Ecology , pp. 22 - 60
Publisher: Cambridge University Press
Print publication year: 2010

Access options

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

References

Abrams, P. A. (1995). Monotonic or unimodal diversity-productivity gradients: what does competition theory predict?Ecology, 76, 2019–27.CrossRefGoogle Scholar
Adersen, H. (1988). Null hypotheses and species composition in the Galápagos Islands. In Diversity and Patterns in Plant Communities, ed. During, H. J., Werger, M. J. A. and Willems, J. H.. The Hague: Academic Publishing, pp. 37–46.Google Scholar
Aldasoro, J. J., Cabezas, F. and Aedo, C. (2004). Diversity and distribution of ferns in sub-Saharan Africa, Madagascar and some islands of the South Atlantic. Journal of Biogeography, 31, 1579–604.CrossRefGoogle Scholar
Allen, A. P. and Gillooly, J. F. (2006). Assessing latitudinal gradients in speciation rates and biodiversity at the global scale. Ecology Letters, 9, 947–54.CrossRefGoogle ScholarPubMed
Andersen, M., Thornhill, A. and Koopowitz, H. (1997). Tropical forest disruption and stochastic biodiversity losses. In Tropical Forest Remnants: Ecology, Management, and Conservation of Fragmented Communities, ed. Laurance, W. F. and Bierregaard, R. O.. Chicago, IL, USA: University of Chicago Press, pp. 281–91.Google Scholar
Bach, K. (2004). Vegetationskundliche Untersuchungen zur Höhenzonierung tropischer Bergwälder in den Anden Boliviens. Marburg, Germany: Verlag Görich and Weiershäuser.Google Scholar
Bach, K., Kessler, M. and Gradstein, S. R. (2007). A simulation approach to determine statistical significance of species turnover peaks in a species-rich tropical cloud forest. Diversity and Distributions, 13, 863–70.CrossRefGoogle Scholar
Balmford, A. and Long, A. (1994). Avian endemism and forest loss. Nature, 372, 623–4.CrossRefGoogle Scholar
Barrington, D. S. (1993). Ecological and historical factors in fern biogeography. Journal of Biogeography, 20, 275–80.CrossRefGoogle Scholar
Barthlott, W., Mutke, J., Rafiqpoor, M. D., Kier, G. and Kreft, H. (2005). Global centres of vascular plant diversity. Nova Acta Leopoldina, 92, 61–83.Google Scholar
Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A. and Potts, R. (1992). Terrestrial Ecosystems through Time: Evolutionary Paleoecology of Terrestrial Plants and Animals. Chicago, IL, USA: University of Chicago Press.Google Scholar
Bhattarai, K. R., Vetaas, O. R. and Grytnes, J. A. (2004). Fern species richness along a central Himalayan elevational gradient, Nepal. Journal of Biogeography, 31, 389–400.CrossRefGoogle Scholar
Bickford, S. A. and Laffan, S. W. (2006). Multi-extent analysis of the relationship between pteridophyte species richness and climate. Global Ecology and Biogeography, 15, 588–601.CrossRefGoogle Scholar
Bokma, F. and Mönkkönen, M. (2000). The mid-domain effect and the longitudinal dimension of continents. Trends in Ecology and Evolution, 15, 288–9.CrossRefGoogle ScholarPubMed
Bromham, L. and Penny, D. (2003). The modern molecular clock. Nature Reviews Genetics, 4, 216–24.CrossRefGoogle ScholarPubMed
Brownsey, P. J. (2001). New Zealand's pteridophyte flora: plants of ancient lineage but recent arrival?Brittonia, 53, 284–303.CrossRefGoogle Scholar
Cardillo, M. (1999). Latitude and rates of diversification in birds and butterflies. Proceedings of the Royal Society of London, Series B, 266, 1221–5.CrossRefGoogle Scholar
Carlquist, S. (1974). Island Biology. New York: Columbia University Press.CrossRefGoogle Scholar
Caulton, E., Keddie, S., Carmichael, R. and Sales, J. (2000). A ten year study of the incidence of spores of bracken (Pteridium aquilinum (L.) Kuhn) in an urban rooftop airstream in south-east Scotland. Aerobiologia, 16, 29–33.CrossRefGoogle Scholar
Chapman, A. D. (2006). Numbers of Living Species in Australia and the World. Report for the Department of the Environment and Heritage, Canberra, Australia.Google Scholar
Christ, H. (1910). Die Geographie der Farne. Leipzig, Germany: Gustav Fischer Verlag.Google Scholar
Chust, G., Chave, J., Condit, R., et al. (2006). Determinants and spatial modeling of tree β-diversity in a tropical forest landscape in Panama. Journal of Vegetation Science, 17, 83–92.CrossRefGoogle Scholar
Colwell, R. K. and Hurtt, G. C. (1994). Nonbiological gradients in species richness and a spurious Rapoport effect. American Naturalist, 144, 570–95.CrossRefGoogle Scholar
Colwell, R. K. and Lees, D. C. (2000). The mid-domain effect, geometric constraints on the geography of species richness. Trends in Ecology and Evolution, 15, 70–76.CrossRefGoogle ScholarPubMed
Colwell, R. K., Rahbek, C. and Gotelli, N. J. (2004). The mid-domain effect and species richness patterns, what have we learned so far?American Naturalist, 163, E000–E023.CrossRefGoogle ScholarPubMed
Conant, D. S. (1978). A radioisotope technique to measure spore dispersal of the tree fern Cyathea arborea Sm. Pollen et Spores, 20, 583–93.Google Scholar
Conant, D. S., Raubeson, L. A., Attwood, D. K. and Stein, D. B. (1995). The relationships of Papuasian Cyatheaceae to New World tree ferns. American Fern Journal, 85, 328–40.CrossRefGoogle Scholar
Condit, R., Pitman, N., Leigh, E. G., et al. (2002). Beta-diversity in tropical forest trees. Science, 295, 666–9.CrossRefGoogle ScholarPubMed
Copeland, E. B. (1939). Fern evolution in Antarctica. Philippine Journal of Science, 70, 157–88.Google Scholar
Cousens, M. I. (1981). Blechnum spicant, habitat and vigor of optimal, marginal, and disjunct populations and field observations of gametophytes. Botanical Gazette, 142, 251–8.CrossRefGoogle Scholar
Cousens, M. I. (1988). Reproductive strategies in pteridophytes. In Plant Reproductive Ecology, Patterns and Strategies, ed. Doust, J. L. and Doust, L. L.. New York: Oxford University Press, pp. 307–28.Google Scholar
Cousens, M. I., Lacey, D. G. and Scheller, J. M. (1985). Life-history studies of ferns: a consideration of perspective. Proceedings of the Royal Society of Edinburgh, Series B, 86, 371–80.Google Scholar
Cowling, R. M., Witkowski, E. T. F., Milewski, A. V. and Newbey, K. R. (1994). Taxonomic, edaphic and biological aspects of narrow plant endemism on matched sites in mediterranean South Africa and Australia. Journal of Biogeography, 21, 651–64.CrossRefGoogle Scholar
Cranfill, R. and Kato, M. (2003). Phylogenetics, biogeography, and classification of the woodwardioid ferns (Blechnaceae). In Pteridology in the New Millennium, ed. Chandra, S. and Srivastava, M.. Dordrecht, The Netherlands: Kluwer Academic Publishers, pp. 25–48.CrossRefGoogle Scholar
Currie, D. J., Mittelbach, G. G., Cornell, H. V., et al. (2004). Predictions and tests of climate-based hypotheses of broad-scale variation in taxonomic richness. Ecology Letters, 7, 1121–34.CrossRefGoogle Scholar
Dalling, J. W., Muller-Landau, H. C., Wright, S. J. and Hubbell, S. P. (2002). Role of dispersal in the recruitment limitation of neotropical pioneer species. Journal of Ecology, 90, 714–27.CrossRefGoogle Scholar
Dassler, C. L. and Farrar, D. R. (1997). Significance of form in fern gametophytes, clonal, gemmiferous gametophytes of Callistopteris baueriana (Hymenophyllaceae). International Journal of Plant Sciences, 158, 622–39.CrossRefGoogle Scholar
Dassler, C. L. and Farrar, D. R. (2001). Significance of gametophyte form in long-distance colonization by tropical, epiphytic ferns. Brittonia, 53, 325–69.CrossRefGoogle Scholar
Dubuisson, J.-Y., Hennequin, S., Rakotondrainibe, F. and Schneider, H. (2003). Ecological diversity and adaptive tendencies in the tropical fern Trichomanes L. (Hymenophyllaceae) with special reference to epiphytic and climbing habits. Botanical Journal of the Linnean Society, 142, 41–63.CrossRefGoogle Scholar
Dzwonko, Z. and Kornaś, J. (1994). Patterns of species richness and distribution of pteridophytes in Rwanda (Central Africa), a numerical approach. Journal of Biogeography, 21, 491–501.CrossRefGoogle Scholar
Ellenberg, H. (1975). Vegetationsstufen in perhumiden bis perariden Bereichen der tropischen Anden. Phytocoenologia, 2, 368–78.Google Scholar
Erdtman, G. (1937). Pollen grains recovered from the atmosphere over the Atlantic. Acta Horti Gotoburgensis, 12, 185–96.Google Scholar
Evans, K. L., Warren, P. H. and Gaston, K. J. (2005). Species-energy relationships at the macroecological scale, a review of mechanisms. Biological Review, 80, 1–25.CrossRefGoogle Scholar
Farrar, D. R. (1967). Gametophytes of four tropical fern genera reproducing independently of their sporophytes in the southern Appalachians. Science, 155, 1266–7.CrossRefGoogle ScholarPubMed
Ferrer-Castán, D. and Vetaas, O. R. (2005). Pteridophyte richness, climate and topography in the Iberian Peninsula, comparing spatial and nonspatial models of richness patterns. Global Ecology and Biogeography, 14, 155–65.CrossRefGoogle Scholar
Frahm, J.-P. and Gradstein, S. R. (1991). An altitudinal zonation of tropical rain forests using bryophytes. Journal of Biogeography, 18, 669–78.CrossRefGoogle Scholar
Gaff, D. F. (1977). Desiccation-tolerant vascular plants of Southern Africa. Oecologia, 31, 95–109.CrossRefGoogle Scholar
Geiger, J. M. O., Ranker, T. A., Neale, J. M. R. and Klimas, S. T. (2007). Molecular biogeography and origins of the Hawaiian fern flora. Brittonia, 59, 142–58.CrossRefGoogle Scholar
Gentry, A. H. (1986). Endemism in tropical vs temperate plant communities. In Conservation Biology: The Science of Scarcity and Diversity, ed. Soulé, M.. Sunderland, MA. USA: Sinauer Associates, pp. 153–81.Google Scholar
Gilpin, M. E. and Hanski, I. A. (1991). Metapopulation Dynamics: Empirical and Theoretical Investigations. London: Academic Press.Google Scholar
Given, D. R. (1993). Changing aspects of endemism and endangerment in Pteridophyta. Journal of Biogeography, 20, 293–302.CrossRefGoogle Scholar
Gradstein, S. R. and Zanten, B. O. (2001). High altitude dispersal of spores, an experimental approach. XVI International Botanical Congress, St. Louis. Abstract Number 15.14.13.
Graves, G. L. (1988). Linearity of geographic range and its possible effect on the population structure of Andean birds. Auk, 105, 47–52.Google Scholar
Greer, G. K. and McCarthy, B. C. (1999). Gametophytic plasticity among four species of ferns with contrasting ecological distributions. International Journal of Plant Sciences, 160, 879–86.CrossRefGoogle ScholarPubMed
Gressitt, J. L., Sedlacek, J., Wise, K. A. J. and Yoshimoto, C. M. (1961). A high speed airplane trap for air-borne organisms. Pacific Insects, 3, 549–55.Google Scholar
Grubb, P. J. (1974). Factors controlling the distribution of forest types on tropical mountains, new facts and new perspective. In Altitudinal Zonation in Malaysia, ed. Flenley, J. R.. Trans. 3rd Aberdeen–Hull Symposium on Malaysian Ecology. University of Hull, Department of Geography, Miscellaneous Series, 16, pp. 13–46.Google Scholar
Grytnes, J. A. (2003a). Species-richness patterns of vascular plants along seven altitudinal transects in Norway. Ecography, 26, 291–300.CrossRefGoogle Scholar
Grytnes, J. A. (2003b). Ecological interpretations of the mid-domain effect. Ecology Letters, 6, 883–8.CrossRefGoogle Scholar
Grytnes, J. A. and Vetaas, O. R. (2002). Species richness and altitude, a comparison between null models and interpolated plant species richness along the Himalayan altitudinal gradient, Nepal. American Naturalist, 159, 294–304.CrossRefGoogle ScholarPubMed
Grytnes, J. A., Heegaard, E. and Romdal, T. S. (2008). Can the mass effect explain the mid-altitudinal peak in vascular species richness?Basic and Applied Ecology, 9, 371–82.CrossRefGoogle Scholar
Guo, Q., Kato, M. and Ricklefs, R. E. (2003). Life history, diversity and distribution, a study of Japanese pteridophytes. Ecography, 26, 129–38.CrossRefGoogle Scholar
Harrison, S. and Cornell, H. V. (2007). Introduction, merging evolutionary and ecological approaches to understanding geographic gradients in species richness. American Naturalist, 170, S1–S4.CrossRefGoogle ScholarPubMed
Haufler, C. H. (2007). Genetics, phylogenetics, and biogeography: Considering how shifting paradigms and continents influence fern diversity. Brittonia, 59, 108–14.CrossRefGoogle Scholar
Haufler, C. H., Hooper, E. A. and Therrien, J. P. (2000). Modes and mechanisms of speciation in pteridophytes: Implications of contrasting patterns in ferns representing temperate and tropical habitats. Plant Species Biology, 15, 223–36.CrossRefGoogle Scholar
Hawkins, B. A. and Diniz-Filho, J. A. F. (2002). The mid-domain effect cannot explain the diversity gradient of Nearctic birds. Global Ecology and Biogeography, 11, 419–26.CrossRefGoogle Scholar
Hawkins, B. A., Diniz-Filho, J. A. F. and Weis, A. E. (2005). The mid-domain effect and diversity gradients, is there anything to learn?American Naturalist, 166, E140–E143.CrossRefGoogle Scholar
Hemp, A. (2002). Ecology of the pteridophytes on the southern slopes of Mt. Kilimanjaro. I. Altitudinal distribution. Plant Ecology, 159, 211–39.CrossRefGoogle Scholar
Hemp, A. (2006). Continuum or zonation? Altitudinal diversity patterns in the forests on Mt. Kilimanjaro. Plant Ecology, 184, 27–42.CrossRefGoogle Scholar
Herzog, S. K., Kessler, M. and Bach, K. (2005). The elevational gradient in Andean bird species richness at the local scale, a foothill peak and a high-elevation plateau. Ecography, 28, 209–22.CrossRefGoogle Scholar
Holdridge, L. R., Grenke, W. C., Hatheway, W. H., Liang, T. and Tosi, J. A. (1971). Forest Environments in Tropical Life Zones, a Pilot Study. Oxford, UK: Pergamon Press.Google Scholar
Hoot, S. B., Taylor, W. C. and Napier, N. S. (2006). Phylogeny and biogeography of Isoëtes (Isoëtaceae) based on nuclear and chloroplast DNA sequence data. Systematic Botany, 31, 449–60.CrossRefGoogle Scholar
Hubbell, S. P. (2001). The Unified Neutral Theory of Biodiversity and Biogeography. Princeton, NJ, USA: Princeton University Press.Google Scholar
Humboldt, A. (1805). Essai sur la géographie des plantes accompagné d'un tableau physique des régions équinoxiales, fondé sur des mesures exécutées, depuis le dixième degré de latitude boréale jusqu'au dixième degré de latitude australe, pendant les années 1799, 1800, 1801, 1802 et 1803. Paris: Levrault Schoell.Google Scholar
Humboldt, A., (1808). Ansichten der Natur mit wissenschaftlichen Erläuterungen. Tübingen, Germany: J. G. Gotta.Google Scholar
Humphries, C. J. (1979). Endemism and evolution in Macaronesia. In Plants and Islands, ed. Bramwell, D.. London: Academic Press, pp. 171–99.Google Scholar
Hurlbert, A. H. (2004). Species-energy relationships and habitat complexity in bird communities. Ecology Letters, 7, 714–20.CrossRefGoogle Scholar
Hurlbert, A. H. (2006). Linking species-area and species-energy relationships in Drosophila microcosms. Ecology Letters, 9, 287–94.CrossRefGoogle ScholarPubMed
Hurlbert, A. H. and Haskell, J. P. (2003). The effect of energy and seasonality on avian species richness and community composition. American Naturalist, 161, 83–97.CrossRefGoogle ScholarPubMed
Jacobsen, W. B. G. and Jacobsen, N. H. G. (1989). Comparison of the pteridophyte floras of southern and eastern Africa, with special reference to high-altitude species. Bulletin du Jardin Botanique National de Belgique, 59, 261–317.CrossRefGoogle Scholar
Jácome, J., Kessler, M. and Smith, A. R. (2007). A human-induced skewed elevational frequency distribution of ferns in the Andes. Global Ecology and Biogeography, 16, 313–8.CrossRefGoogle Scholar
James, K. E., Schneider, H., Ansell, S. E., et al. (2008). Diversity arrays technology (DArT) for pan-genomic evolutionary studies of non-model organisms. PLoS ONE, 3, e 1682, 1–11.Google ScholarPubMed
Janssen, T., Kreier, H.-P. and Schneider, H. (2007). Origin and diversification of African ferns with special emphasis on Polypodiaceae. Brittonia, 59, 159–81.CrossRefGoogle Scholar
Jetz, W. and Rahbek, C. (2001). Geometric constraints explain much of the species richness pattern in African birds. Proceedings of the National Academy of Sciences, USA, 98, 5661–6.CrossRefGoogle ScholarPubMed
Jones, M. M., Tuomisto, H., Clark, D. B. and Olivas, P. (2006). Effects of mesoscale environmental heterogeneity and dispersal limitation on floristic variation in rain forest ferns. Journal of Ecology, 94, 181–95.CrossRefGoogle Scholar
Jones, M. M., Olivas-Rojas, P., Tuomisto, H. and Clark, D. B. (2007). Environmental and neighbourhood effects on tree fern distributions in a neotropical lowland rain forest. Journal of Vegetation Science, 18, 13–24.CrossRefGoogle Scholar
Jones, M. M., Tuomisto, H., Clark, D. B. and Olivas, P. (2008). Differences in the degree of environmental control of large and small tropical plants, just a sampling effect?Journal of Ecology, 96, 367–77.CrossRefGoogle Scholar
Kaspari, M., O'Donnell, S. and Kercher, J. R. (2000). Energy, density, and constraints to species richness, ant assemblages along a productivity gradient. American Naturalist, 155, 280–93.CrossRefGoogle ScholarPubMed
Kattan, G. H. and Franco, P. (2004). Bird diversity along elevational gradients in the Andes of Colombia, area and mass effects. Global Ecology and Biogeography, 13, 451–8.CrossRefGoogle Scholar
Kato, M. (1993). Biogeography of ferns; dispersal and vicariance. Journal of Biogeography, 20, 265–74.CrossRefGoogle Scholar
Kawai, H., Kanegae, T., Christensen, S., et al. (2003). Responses of ferns to red light are mediated by an unconventional photoreceptor. Nature, 421, 287–90.CrossRefGoogle ScholarPubMed
Kessler, M. (1999). Plant species richness and endemism during natural landslide succession in a perhumid montane forest in the Bolivian Andes. Ecotropica, 5, 123–36.Google Scholar
Kessler, M. (2000a). Altitudinal zonation of Andean cryptogam communities. Journal of Biogeography, 27, 275–82.CrossRefGoogle Scholar
Kessler, M. (2000b). Elevational gradients in species richness and endemism of selected plant groups in the central Bolivian Andes. Plant Ecology, 149, 181–93.CrossRefGoogle Scholar
Kessler, M. (2001a). Maximum plant community endemism at intermediate intensities of anthropogenic disturbance in Bolivian montane forests. Conservation Biology, 15, 634–41.CrossRefGoogle Scholar
Kessler, M. (2001b). Pteridophyte species richness in Andean forests in Bolivia. Biodiversity and Conservation, 10, 1473–95.CrossRefGoogle Scholar
Kessler, M. (2001c). Patterns of diversity and range size of selected plant groups along an elevational transect in the Bolivian Andes. Biodiversity and Conservation, 10, 1897–1920.CrossRefGoogle Scholar
Kessler, M. (2002a). The elevational gradient of Andean plant endemism, varying influences of taxon-specific traits and topography at different taxonomic levels. Journal of Biogeography, 29, 1159–66.CrossRefGoogle Scholar
Kessler, M. (2002b). Range size and its ecological correlates among the pteridophytes of Carrasco National Park, Bolivia. Global Ecology and Biogeography, 11, 89–102.CrossRefGoogle Scholar
Kessler, M. and Bach, K. (1999). Using indicator families for vegetation classification in species-rich Neotropical forests. Phytocoenologia, 29, 485–502.Google Scholar
Kessler, M. and Siorak, Y. (2007). Desiccation and rehydration experiments on leaves of 43 pteridophyte species. American Fern Journal, 97, 175–85.CrossRefGoogle Scholar
Kessler, M. and Smith, A. R. (2006). Five new species of Asplenium (Aspleniaceae) from Bolivia. Candollea, 61, 305–13.Google Scholar
Kessler, M. and Smith, A. R. (2007). Ten new species and other nomenclatural changes for ferns from Bolivia. Brittonia, 59, 186–97.CrossRefGoogle Scholar
Kessler, M., Parris, B. S. and Kessler, E. (2001). A comparison of the tropical montane pteridophyte communities of Mount Kinabalu, Borneo, and Parque Nacional Carrasco, Bolivia. Journal of Biogeography, 28, 611–22.CrossRefGoogle Scholar
Kessler, M., Siorak, Y., Wunderlich, M. and Wegner, C. (2007). Patterns of morphological leaf traits among pteridophytes along humidity and temperature gradients in the Bolivian Andes. Functional Plant Biology, 34, 963–71.CrossRefGoogle Scholar
Kluge, J. and Kessler, M. (2006). Fern endemism and its correlates: contribution from an elevational transect in Costa Rica. Diversity and Distributions, 12, 535–45.CrossRefGoogle Scholar
Kluge, J. and Kessler, M. (2007). Morphological characteristics of fern assemblages along an elevational gradient, patterns and causes. Ecotropica, 13, 27–43.Google Scholar
Kluge, J., Kessler, M. and Dunn, R. (2006). What drives elevational patterns of diversity? A test of geometric constraints, climate, and species pool effects for pteridophytes on an elevational gradient in Costa Rica. Global Ecology and Biogeography, 15, 358–71.CrossRefGoogle Scholar
Kluge, J., Bach, K. and Kessler, M. (2008). Elevational distribution and zonation of tropical pteridophyte assemblages in Costa Rica. Basic and Applied Ecology, 9, 35–43.CrossRefGoogle Scholar
Korall, P., Pryer, K. M., Metzgar, J. S., Schneider, H. and Conant, D. S. (2006). Tree ferns: monophyletic groups and their relationships as revealed by four protein-coding plastid loci. Molecular Phylogenetics and Evolution, 39, 830–45.CrossRefGoogle ScholarPubMed
Kornaś, J. (1993). The significance of historical factors and ecological preference in the distribution of African pteridophytes. Journal of Biogeography, 20, 281–6.CrossRefGoogle Scholar
Körner, C. (2000). Why are there global gradients in species richness? Mountains might hold the answer. Trends in Ecology and Evolution, 15, 513–14.CrossRefGoogle Scholar
Kramer, K. U. (1993). Distribution patterns in major pteridophyte taxa relative to those of angiosperms. Journal of Biogeography, 20, 287–91.CrossRefGoogle Scholar
Kramer, K. U., Schneller, J. J. and Wollenweber, E. (1995). Farne und Farnverwandte. Stuttgart, Germany: Georg Thieme Verlag.Google Scholar
Kreft, H. and Jetz, W. (2007). Global patterns and determinants of vascular plant diversity. Proceedings of the National Academy of Sciences, USA, 104, 5925–30.CrossRefGoogle ScholarPubMed
Kreier, H.-P. and Schneider, H. (2006). Phylogeny and biogeography of the staghorn fern genus Platycerium (Polypodiaceae, Polypodiidae). American Journal of Botany, 93, 217–25.Google Scholar
Krömer, T., Kessler, M., Gradstein, S. R. and Acebey, A. (2005). Diversity patterns of vascular epiphytes along an elevational gradient in the Andes. Journal of Biogeography, 32, 1799–810.CrossRefGoogle Scholar
Kruckeberg, A. R. and Rabinowitz, D. (1985). Biological aspects of endemism in higher plants. Annual Review of Ecology and Systematics, 16, 447–79.CrossRefGoogle Scholar
Lawton, R. O., Nair, U. S., Pielke, R. A., Sr. and Welch, R. M. (2001). Climatic impact of tropical lowland deforestation on nearby montane cloud forests. Science, 294, 584–7.Google ScholarPubMed
Lehmann, A., Leathwick, J. R. and Overton, J. McC. (2002). Assessing New Zealand fern diversity from spatial predictions of species assemblages. Biodiversity and Conservation, 11, 2217–38.CrossRefGoogle Scholar
Lehnert, M. (2006). The Cyatheaceae and Dicksoniaceae (Pteridophyta) of Bolivia. Brittonia, 58, 229–44.CrossRefGoogle Scholar
Leibold, M. A., Holyoak, M., Mouquet, N., et al. (2004). The metacommunity concept, a framework for multi-scale community ecology. Ecology Letters, 7, 601–13.CrossRefGoogle Scholar
Lester, S. E., Ruttenberg, B. I., Gaines, S. D. and Kinlan, B. P. (2007). The relationship between dispersal ability and geographic range size. Ecology Letters, 10, 745–58.CrossRefGoogle ScholarPubMed
Little, D. P. and Barrington, D. S. (2003). Major evolutionary events in the origin and diversification of the fern genus Polystichum (Dryopteridaceae). American Journal of Botany, 90, 508–14.CrossRefGoogle Scholar
Lloyd, R. M. and Klekowski, E. J. (1970). Spore germination and viability in Pteridophyta: evolutionary significance of chlorophyllous spores. Biotropica, 2, 129–37.CrossRefGoogle Scholar
Lomolino, M. V. (2001). Elevational gradients of species-density, historical and prospective views. Global Ecology and Biogeography, 10, 3–13.CrossRefGoogle Scholar
Lomolino, M. V., Riddle, B. R. and Brown, J. H. (2006). Biogeography. Sunderland, MA, USA: Sinauer Associates.Google Scholar
Lovis, J. D. (1959). The geographical affinities of the New Zealand pteridophyte flora. British Fern Gazette, 10, 1–7.Google Scholar
Luebke, N. T. and Budke, J. M. (2003). Isoëtes tennesseensis (Isoëtaceae), an octoploid quillwort from Tennessee. American Fern Journal, 93, 184–90.CrossRefGoogle Scholar
Lwanga, J. S., Balmford, A. and Badaza, R. (1998). Assessing fern diversity, relative species richness and its environmental correlates in Uganda. Biodiversity and Conservation, 7, 1387–98.CrossRefGoogle Scholar
MacArthur, R. H. and Wilson, E. O. (1967). The Theory of Island Biogeography. Monographs in Population Biology, 1. Princeton, NJ, USA: Princeton University Press.Google Scholar
Magurran, A. E. (2004). Measuring Biological Diversity. Malden, MA: Blackwell Publishing.Google Scholar
McCain, C. M. (2005). Elevational gradients in diversity of small mammals. Ecology, 86, 366–72.CrossRefGoogle Scholar
Mehltreter, K. V. (1995). Species richness and geographical distribution of montane pteridophytes of Costa Rica, Central America. Feddes Repertorium, 106, 563–84.CrossRefGoogle Scholar
Mittelbach, G. G., Schemske, D. W., Cornell, H. V., et al. (2007). Evolution and the latitudinal diversity gradient, speciation, extinction and biogeography. Ecology Letters, 10, 315–31.CrossRefGoogle ScholarPubMed
Mönkkönen, M., Forsman, J. T. and Bokma, F. (2006). Energy availability, abundance, energy-use and species richness in forest bird communities: a test of the species-energy theory. Global Ecology and Biogeography, 15, 290–302.Google Scholar
Moolman, H. J. and Cowling, R. M. (1994). The impact of elephant and goat grazing on the endemic flora of South African succulent thicket. Biological Conservation, 68, 53–61.CrossRefGoogle Scholar
Moran, R. C. (1995). The importance of mountains to pteridophytes, with emphasis on neotropical montane forests. In Biodiversity and Conservation of Neotropical Montane Forests, ed. Churchill, S. P., Balslev, H., Forero, E. and Luteyn, J. L.. Bronx, NY, USA: The New York Botanical Garden, pp. 359–63.Google Scholar
Moran, R. C. (1996). The importance of the Andes as a barrier to migration, as illustrated by the pteridophytes of the Chocó phytogeographic region. In Pteridology in Perspective, ed. Camus, J. M., Gibby, M. and Johns, R. J.. Kew, UK: Royal Botanic Gardens, p. 75.Google Scholar
Moran, R. C. and Smith, A. R. (2001). Phytogeographic relationships between neotropical and African-Madagascan pteridophytes. Brittonia, 53, 304–51.CrossRefGoogle Scholar
Nathan, R. (2006). Long-distance dispersal in plants. Science, 313, 786–8.CrossRefGoogle Scholar
Ohlemüller, R. and Wilson, J. B. (2000). Vascular plant species richness along latitudinal and altitudinal gradients, a contribution from New Zealand temperate rainforests. Ecology Letters, 3, 262–6.CrossRefGoogle Scholar
Paciencia, M. L. B. and Prado, J. (2005). Effects of forest fragmentation on pteridophyte diversity in a tropical rain forest in Brazil. Plant Ecology, 180, 87–104.CrossRefGoogle Scholar
Page, C. N. (2002). Ecological strategies in fern evolution, a neopteridological overview. Review of Palaeobotany and Palynology, 119, 1–33.CrossRefGoogle Scholar
Palmer, D. D. (2003). Hawai‘i's Ferns and Fern Allies. Honolulu, HI, USA: University of Hawaii Press.Google Scholar
Parris, B. S. (2001). Circum-Antarctic continental distribution patterns in pteridophyte species. Brittonia, 53, 270–83.CrossRefGoogle Scholar
Parris, B. S., Beaman, R. S. and Beaman, J. H. (1992). Ferns and Fern Allies. Vol I of The Plants of Mount Kinabalu. Kew, UK: Royal Botanic Gardens.Google Scholar
Pausas, J. G. and Sáez, L. (2000). Pteridophyte richness in the NE Iberian Peninsula, biogeographic patterns. Plant Ecology, 148, 197–207.CrossRefGoogle Scholar
Pautasso, M. and Gaston, K. J. (2005). Resources and global avian assemblage structure in forests. Ecology Letters, 8, 282–9.CrossRefGoogle Scholar
Pautasso, M. and Gaston, K. J. (2006). A test of the mechanisms behind avian generalized individuals–area relationships. Global Ecology and Biogeography, 15, 303–17.Google Scholar
Peck, J. H., Peck, C. J. and Farrar, D. R. (1990). Influences of life history events on formation of local and distant fern populations. American Fern Journal, 80, 126–42.CrossRefGoogle Scholar
Perrie, L. and Brownsey, P. (2007). Molecular evidence for long-distance dispersal in the New Zealand pteridophyte flora. Journal of Biogeography, 34, 2028–38.CrossRefGoogle Scholar
Pickett, F. (1931). Notes on xerophytic ferns. American Fern Journal, 21, 49–57.CrossRefGoogle Scholar
Pole, M. (1994). The New Zealand flora – entirely long-distance dispersal?Journal of Biogeography, 21, 625–35.CrossRefGoogle Scholar
Polunin, N. (1951). Seeking airborne botanical particles about the North Pole. Svensk Botanisk Tidskrift, 45, 320–54.Google Scholar
Potts, M. D., Ashton, P. S., Kaufmann, L. S. and Plotkin, J. B. (2002). Habitat patterns in tropical rain forests: a comparison of 105 plots in northwest Borneo. Ecology, 83, 2782–97.CrossRefGoogle Scholar
Poulsen, A. D., Tuomisto, H. and Balslev, H. (2006). Edaphic and floristic variation within 1-ha plot of lowland Amazonian rain forest. Biotropica, 38, 468–78.CrossRefGoogle Scholar
Pounds, J. A., Fogden, M. P. L. and Campbell, J. H. (1999). Biological response to climate change on a tropical mountain. Nature, 398, 611–14.CrossRefGoogle Scholar
Price, J. P. and Clague, D. A. (2002). How old is the Hawai‘ian biota? Geology and phylogeny suggest recent divergence. Proceedings of the Royal Society of London, Series B, 269, 2429–35.CrossRefGoogle Scholar
Pryer, K. M., Smith, A. R., Hunt, J. and Dubuisson, J.-Y. (2001). rbcL data reveal two monophyletic groups of filmy ferns (Filicopsida, Hymenophyllaceae). American Journal of Botany, 88, 1118–30.CrossRefGoogle Scholar
Pulliam, H. R. (1988). Sources, sinks and population regulation. American Naturalist, 132, 652–61.CrossRefGoogle Scholar
Punetha, N. (1991). Studies on atmospheric fern spores at Pithorgarh (northwest Himalaya) with particular reference to distribution of ferns in the Himalayas. Annual Review of Plant Science, 13, 146–61.Google Scholar
Rahbek, C. (1995). The elevational gradient of species richness: a uniform pattern?Ecography, 18, 200–5.CrossRefGoogle Scholar
Rahbek, C. (1997). The relationship among area, elevation, and regional species richness in neotropical birds. American Naturalist, 149, 875–902.CrossRefGoogle ScholarPubMed
Ranker, T. A., Floyd, S. K. and Trapp, P. G. (1994). Multiple colonizations of Asplenium adiantum-nigrum onto the Hawai‘ian archipelago. Evolution, 48, 1364–70.CrossRefGoogle Scholar
Ranker, T. A., Geiger, J. M. O., Kennedy, S. C., et al. (2003). Molecular phylogenetics and evolution of the endemic Hawai‘ian genus Adenophorus (Grammitidaceae). Molecular Phylogenetics and Evolution, 26, 337–47.CrossRefGoogle Scholar
Ranker, T. A., Smith, A. R., Parris, B. S., et al. (2004). Phylogeny and evolution of grammitid ferns (Grammitidaceae): a case of rampant morphological homoplasy. Taxon, 53, 415–428.CrossRefGoogle Scholar
Renner, S. (2005). Relaxed molecular clocks for dating historical plant dispersal events. Trends in Plant Science, 10, 550–8.CrossRefGoogle ScholarPubMed
Richard, M., Bernhardt, T. and Bell, G. (2000). Environmental heterogeneity and the spatial structure of fern species diversity in one hectare of old-growth forest. Ecography, 23, 231–45.CrossRefGoogle Scholar
Ricklefs, R. E. (2005). Phylogenetic perspectives on patterns of regional and local richness. In Tropical Rainforest, Past, Present, and Future, ed. Bermingham, E., Dick, C. W. and Moritz, C.. Chicago, IL, USA: University of Chicago Press, pp. 16–40.Google Scholar
Ricklefs, R. E. (2007). Estimating diversification rates from phylogenetic information. Trends in Ecology and Evolution, 22, 601–10.CrossRefGoogle ScholarPubMed
Roos, M. (1996). Mapping the world's pteridophyte diversity – systematics and floras. In Pteridology in Perspective, ed. Camus, J. M., Gibby, M. and Johns, R. J.. Kew, UK: Royal Botanic Gardens, pp. 29–42.Google Scholar
Roos, M., Keßler, P. J. A., Gradstein, S. R. and Baas, P. (2004). Species diversity and endemism of five major Malesian islands, diversity-area relationships. Journal of Biogeography, 31, 1893–1908.CrossRefGoogle Scholar
Rosenzweig, M. L. and Abramsky, Z. (1993). How are diversity and productivity related? In Species Diversity in Ecological Communities: Historical and Geographical Perspectives, ed. Ricklefs, R. E. and Schluter, D.. Chicago, IL, USA: University of Chicago Press, pp. 52–65.Google Scholar
Rosenzweig, M. L. and Ziv, Y. (1999). The echo pattern of species diversity, pattern and process. Ecography, 22, 614–28.CrossRefGoogle Scholar
Rouhan, G., Dubuisson, J.-Y., Rakotondrainibe, F., et al. (2004). Molecular phylogeny of the fern genus Elaphoglossum (Elaphoglossaceae) based on chloroplast non-coding DNA sequences: contributions of species from the Indian Ocean area. Molecular Phylogenetics and Evolution, 33, 745–763.CrossRefGoogle ScholarPubMed
Roy, K. and Goldberg, E. E. (2007). Origination, extinction, and dispersal, integrative models for understanding present-day diversity gradients. American Naturalist, 170, S71–S85.Google ScholarPubMed
Ruokolainen, K., Linna, A. and Tuomisto, H. (1997). Use of Melostomataceae and pteridophytes for revealing phytogeographical patterns in Amazonian rain forests. Journal of Tropical Ecology, 13, 243–56.CrossRefGoogle Scholar
Ruokolainen, K., Tuomisto, H., Macía, M. J., Higgins, M. A. and Yli-Halla, M. (2007). Are floristic and edaphic patterns in Amazonian rain forests congruent for trees, pteridophytes and Melastomataceae?Journal of Tropical Ecology, 23, 13–25.CrossRefGoogle Scholar
Salovaara, K. J., Cárdenas, G. G. and Tuomisto, H. (2004). Forest classification in an Amazonian rainforest landscape using pteridophytes as indicator species. Ecography, 27, 689–700.CrossRefGoogle Scholar
Samways, M. J. (1994). Insect Conservation Biology. London: Chapman and Hall.Google Scholar
Sato, T. and Sakai, A. (1980). Freezing resistance of gametophytes of the temperate fern, Polystichum retroso-paleaceum. Canadian Journal of Botany, 58, 1144–8.CrossRefGoogle Scholar
Sato, T. and Sakai, A. (1981). Cold tolerance of gametophytes and sporophytes of some cool temperature ferns native to Hokkaido. Canadian Journal of Botany, 59, 604–8.CrossRefGoogle Scholar
Schneider, H., Schuettpelz, E., Pryer, K. M., et al. (2004a). Ferns diversified in the shadow of angiosperms. Nature, 428, 553–7.CrossRefGoogle ScholarPubMed
Schneider, H., Russell, S. J., Cox, C. J., et al. (2004b). Chloroplast phylogeny of asplenioid ferns based on rbcL and trnL-F spacer sequences (Polypodiidae, Aspleniaceae) and its implications for biogeography. Systematic Botany, 29, 260–74.CrossRefGoogle Scholar
Schneider, H., Ranker, T. A., Russell, S. J., et al. (2005). Origin of the endemic fern genus Diellia coincides with the renewal of Hawai‘ian terrestrial life in the Miocene. Proceedings of the Royal Society of London, Series B, 272, 455–60.CrossRefGoogle Scholar
Schneider-Pötsch, H. A. W., Kolukisaoglu, Ü., Clapham, D. H., Hughes, J. and Lamparter, T. (1998). Non-angiosperm phytochromes and the evolution of vascular plants. Physiologia Plantarum, 102, 612–22.CrossRefGoogle Scholar
Schneller, J. J. and Liebst, B. (2007). Patterns of variation of a common fern (Athyrium filix-femina; Woodsiaceae): population structure along and between altitudinal gradients. American Journal of Botany, 94, 965–71.CrossRefGoogle ScholarPubMed
Schuettpelz, E. and Pryer, K. M. (2006). Reconciling extreme branch length differences, decoupling time and rate through the evolutionary history of filmy ferns. Systematic Botany, 55, 485–502.CrossRefGoogle ScholarPubMed
Schuettpelz, E., Schneider, H., Huiet, L., Windham, M. D. and Pryer, K. M. (2007). A molecular phylogeny of the fern family Pteridaceae, assessing overall relationships and the affinities of previously unsampled genera. Molecular Phylogenetics and Evolution, 44, 1172–85.CrossRefGoogle ScholarPubMed
Sheffield, E. (1996). From pteridophyte spore to sporophyte in the natural environment. In Pteridology in Perspective, ed. Gibby, M. and Johns, R. J.. Kew, UK: Royal Botanic Gardens, pp. 541–9.Google Scholar
Shepherd, L. D., Perrie, L. R. and Brownsey, P. J. (2007). Fire and ice: volcanic and glacial impacts on the phylogeography of the New Zealand forest fern Asplenium hookerianum. Molecular Ecology, 16, 4536–49.CrossRefGoogle ScholarPubMed
Shmida, A. and Wilson, M. W. (1985). Biological determinants of species diversity. Journal of Biogeography, 12, 1–20.CrossRefGoogle Scholar
Smith, A. R. (1972). Comparison of fern and flowering plant distributions with some evolutionary interpretations for ferns. Biotropica, 4, 4–9.CrossRefGoogle Scholar
Smith, A. R. (1993). Phytogeographic principles and their use in understanding fern relationships. Journal of Biogeography, 20, 255–64.CrossRefGoogle Scholar
Smith, A. R. (2006). Floristics in the 21st century: balancing user-needs and phylogenetic information. Fern Gazette, 17, 105–37.Google Scholar
Smith, A. R., Pryer, K. M., Schuettpelz, E., et al. (2006). A classification for extant ferns. Taxon, 55, 705–31.CrossRefGoogle Scholar
Soria-Auza, R. W. and Kessler, M. (2008). The influence of sampling intensity on the perception of the spatial distribution of tropical diversity and endemism, a case study of ferns from Bolivia. Diversity and Distributions, 14, 123–30.CrossRefGoogle Scholar
Stevens, G. C. (1989). The latitudinal gradient in geographical range: how so many species coexist in the tropics. American Naturalist, 133, 240–56.CrossRefGoogle Scholar
Still, C. J., Foster, P. N. and Schneider, S. H. (1999). Simulating the effects of climate change on tropical montane cloud forests. Nature, 398, 608–10.CrossRefGoogle Scholar
Svenning, J.-C., Kinner, D. A., Stallard, R. F., Engelbrecht, B. M. J. and Wright, S. J. (2004). Ecological determinism in plant community structure across a tropical forest landscape. Ecology, 85, 2526–38.CrossRefGoogle Scholar
Trewick, S. A., Morgan-Richards, M., Russell, S. J., et al. (2002). Polyploidy, phylogeography and Pleistocene refugia of the rockfern Asplenium ceterach: evidence from chloroplast DNA. Molecular Ecology, 11, 2003–12.CrossRefGoogle ScholarPubMed
Tryon, A. F. (1957). A revision of the fern genus Pellaea section Pellaea. Annals of the Missouri Botanical Garden, 44, 125–93.CrossRefGoogle Scholar
Tryon, A. F. and Lugardon, B. (1990). Spores of the Pteridophyta. Berlin: Springer-Verlag.Google Scholar
Tryon, R. M. (1970). Development and evolution of fern floras of oceanic islands. Biotropica, 2, 76–84.CrossRefGoogle Scholar
Tryon, R. M. (1972). Endemic areas and geographic speciation in tropical American ferns. Biotropica, 4, 121–31.CrossRefGoogle Scholar
Tryon, R. M. (1976). The biogeography of species, with special reference to ferns. Botanical Review, 52, 116–56.Google Scholar
Tryon, R. M. (1985). Fern speciation and biogeography. Proceedings of the Royal Society of Edinburgh, 86B, 353–60.Google Scholar
Tryon, R. M. (1986). The biogeography of species, with special reference to ferns. Botanical Review, 52, 118–56.CrossRefGoogle Scholar
Tuomisto, H. (1994). Ecological Variation in the Rain Forests of Peruvian Amazonia, Integrating Fern Distribution Patterns with Satellite Imagery. Reports from the Department of Biology, 45. Turku, Finland: University of Turku.Google Scholar
Tuomisto, H. (1998). What satellite imagery and large-scale field studies can tell about biodiversity patterns in Amazonian forests. Annals of the Missouri Botanical Garden, 85, 48–62.CrossRefGoogle Scholar
Tuomisto, H. (2006). Edaphic niche differentiation among Polybotrya ferns in Western Amazonia, implications for coexistence and speciation. Ecography, 29, 273–84.CrossRefGoogle Scholar
Tuomisto, H. and Poulsen, A. D. (1996). Influence of edaphic specialization of pteridophyte distribution in neotropical rain forests. Journal of Biogeography, 23, 283–93.CrossRefGoogle Scholar
Tuomisto, H. and Poulsen, A. D. (2000). Pteridophyte diversity and species composition in four Amazonian rain forests. Journal of Vegetation Science, 11, 383–96.CrossRefGoogle Scholar
Tuomisto, H. and Ruokolainen, K. (1994). Distribution of Pteridophyta and Melastomataceae along an edaphic gradient in an Amazonian rain forest. Journal of Vegetation Science, 5, 25–34.CrossRefGoogle Scholar
Tuomisto, H., Ruokolainen, K., Kalliola, R., et al. (1995). Dissecting Amazonian biodiversity. Science, 269, 63–6.CrossRefGoogle ScholarPubMed
Tuomisto, H., Poulsen, A. D. and Moran, R. C. (1998). Edaphic distribution of some species of the fern genus Adiantum in western Amazonia. Biotropica, 30, 392–9.CrossRefGoogle Scholar
Tuomisto, H., Ruokolainen, K., Poulsen, A. D., et al. (2002). Distribution and diversity of pteridophytes and Melastomataceae along edaphic gradients in Yasuni National Park, Ecuadorian Amazonia. Biotropica, 34, 516–33.Google Scholar
Tuomisto, H., Poulsen, A. D., Ruokolainen, K., et al. (2003a). Linking floristic patterns with soil heterogeneity and satellite imagery in Ecuadorian Amazonia. Ecological Applications, 13, 352–71.CrossRefGoogle Scholar
Tuomisto, H., Ruokolainen, K., Aguilar, M. and Sarmientos, A. (2003b). Floristic patterns along a 43-km long transect in an Amazonian rain forest. Journal of Ecology, 91, 743–56.CrossRefGoogle Scholar
Tuomisto, H., Ruokolainen, K. and Yli-Halla, M. (2003c). Dispersal, environment, and floristic variation of Western Amazonian forests. Science, 299, 241–4.CrossRefGoogle ScholarPubMed
Zanten, B. O. and Gradstein, S. R. (1988). Experimental dispersal geography of neotropical liverworts. Beihefte zur Nova Hedwigia, 90, 41–94.Google Scholar
Vogel, J. C., Barrett, J. A., Rumsey, F. J. and Gibby, M. (1999a). Identifying multiple origins in polyploid homosporous pteridophytes. In Molecular Systematics and Plant Evolution, ed. Hollingsworth, P. M., Bateman, R. M. and Gornall, R. J.. London: Taylor & Francis, pp. 101–17.Google Scholar
Vogel, J. C., Rumsey, F. J., Schneller, J. J., Barrett, J. A. and Gibby, M. (1999b). Where are the glacial refugia in Europe? Evidence from pteridophytes. Botanical Journal of the Linnean Society, 66, 23–37.CrossRefGoogle Scholar
Vormisto, J., Phillips, O. L., Ruokolainen, K., Tuomisto, H. and Vásquez, R. (2000). A comparison of fine-scale distribution patterns of four plant groups in an Amazonian rainforest. Ecography, 23, 349–59.CrossRefGoogle Scholar
Wagner, W. H. (1995). Evolution of Hawai‘ian ferns and fern allies in relation to their conservation status. Pacific Science, 49, 31–41.Google Scholar
Wagner, W. H., Herbst, D. R. and Sohmer, S. H. (1990). Manual of the Flowering Plants of Hawai‘i, Volume 1. Special Publication 83. Honolulu, HI, USA: University of Hawai‘i Press and Bishop Museum Press.Google Scholar
Waide, R. B., Willig, M. R., Steiner, C. F., et al. (1999). The relationship between productivity and species richness. Annual Review of Ecology and Systematics, 30, 257–300.CrossRefGoogle Scholar
Watkins, J. E., Jr., Cardelús, C., Colwell, R. K. and Moran, R. C. (2006). Species richness and distribution of ferns along an elevational gradient in Costa Rica. American Journal of Botany, 93, 73–83.CrossRefGoogle Scholar
Watkins, J. E., Jr., Mack, M. K. and Mulkey, S. S. (2007). Gametophyte ecology and demography of epiphytic and terrestrial tropical ferns. American Journal of Botany, 94, 701–8.CrossRefGoogle ScholarPubMed
Wiens, J. J. and Donoghue, M. J. (2004). Historical biogeography, ecology and species richness. Trends in Ecology and Evolution, 19, 639–44.CrossRefGoogle ScholarPubMed
Wikström, N. and Kenrick, P. (1997). Phylogeny of Lycopodiaceae (Lycopsida) and the relationship of Phylloglossum drummondii Kunze based on rbcL sequence data. International Journal of Plant Sciences, 160, 862–71.CrossRefGoogle Scholar
Wikström, N. and Kenrick, P. (2000). Phylogeny of epiphytic Huperzia (Lycopodiaceae), paleotropical and neotropical clades corroborated by rbcL sequences. Nordic Journal of Botany, 20, 165–71.CrossRefGoogle Scholar
Wikström, N. and Kenrick, P. (2001). Evolution of Lycopodiaceae (Lycopsida), estimating divergence times from rbcL gene sequences by use of nonparametric rate smoothing. Molecular Phylogenetics and Evolution, 19, 177–86.CrossRefGoogle Scholar
Wikström, N., Kenrick, P. and Chase, M. (1999). Epiphytism and terrestrialization in tropical Huperzia (Lycopodiaceae). Plant Systematics and Evolution, 218, 221–43.CrossRefGoogle Scholar
Wild, M. and Gagnon, D. (2005). Does lack of suitable habitat explain the patchy distribution of rare calcicole fern species?Ecography, 28, 191–6.CrossRefGoogle Scholar
Willig, M. R., Kaufman, D. M. and Stevens, R. D. (2003). Latitudinal gradients of biodiversity, pattern, process, scale, and synthesis. Annual Reviews of Ecology, Evolution and Systematics, 34, 273–309.CrossRefGoogle Scholar
Wilson, D. S. (1992). Complex interactions in metacommunities, with implications for biodiversity and higher levels of selection. Ecology, 73, 1984–2000.CrossRefGoogle Scholar
Wolf, P. G., Sheffield, E. and Haufler, C. H. (1991). Estimates of gene flow, genetic substructure and population heterogeneity in bracken (Pteridium aquilinum). Biological Journal of the Linnean Society, 42, 407–23.CrossRefGoogle Scholar
Wolf, P. G., Schneider, H. and Ranker, T. A. (2001). Geographic distributions of homosporous ferns: does dispersal obscure evidence of vicariance?Journal of Biogeography, 28, 263–70.CrossRefGoogle Scholar
Wright, D. H. (1983). Species-energy theory, an extension of species-area theory. Oikos, 41, 496–506.CrossRefGoogle Scholar
Yatabe, Y., Masuyama, S., Darnaedi, D. and Murakami, N. (2001). Molecular systematics of the Asplenium nidus complex from Mt. Halimun National Park, Indonesia: evidence for reproductive isolation among three sympatric rbcL sequence types. American Journal of Botany, 88, 1517–22.CrossRefGoogle ScholarPubMed
Young, K. R. and León, B. (1989). Pteridophyte species diversity in the Central Peruvian Amazon, importance of edaphic specialization. Brittonia, 41, 388–95.CrossRefGoogle Scholar
Zotz, G. (2005). Vascular epiphytes in the temperate zones – a review. Plant Ecology, 176, 173–83.CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

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
×