Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-5nwft Total loading time: 0 Render date: 2024-05-09T11:23:00.667Z Has data issue: false hasContentIssue false

Literature Cited

Published online by Cambridge University Press:  13 October 2017

Tod F. Stuessy
Affiliation:
Ohio State University
Daniel J. Crawford
Affiliation:
Kansas University
Patricio López-Sepúlveda
Affiliation:
Universidad de Concepción, Chile
Carlos M. Baeza
Affiliation:
Universidad de Concepción, Chile
Eduardo A. Ruiz
Affiliation:
Universidad de Concepción, Chile
Get access
Type
Chapter
Information
Plants of Oceanic Islands
Evolution, Biogeography, and Conservation of the Flora of the Juan Fernández (Robinson Crusoe) Archipelago
, pp. 381 - 383
Publisher: Cambridge University Press
Print publication year: 2017

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

Literature Cited

Aagesen, L. 1999. Phylogeny of the tribe Colletieae, Rhamnaceae. Bot. J. Linn. Soc. 131: 143.CrossRefGoogle Scholar
Aagesen, L., Medan, D., Kellermann, J., and Hilger, H. H. 2005. Phylogeny of the tribe Colletieae (Rhamnaceae): a sensitivity analysis of the plastid region trnL-trnF combined with morphology. Plant Syst. Evol. 250: 197214.Google Scholar
Abbott, R., Albach, D., Ansell, S., Arntzen, J., Baird, S. J. E., et al. 2013. Hybridization and speciation. Evol. Biol. 26: 229246.Google Scholar
Abbott, R. J., Hegarty, M. J., Hiscock, S. J., and Brennan, A. C. 2010. Homoploid hybrid speciation in action. Taxon 59: 13751386.Google Scholar
Abe, T. 2006. Threatened pollination systems in native flora of the Ogasawara (Bonin) Islands. Ann. Bot. 98: 317334.CrossRefGoogle ScholarPubMed
Adhikari, B., Milne, R., Pennington, R. T., Sarkinen, T., and Pendry, C. A. 2015. Systematics and biogeography of Berberis s.l. inferred from nuclear ITS and chloroplast ndhF gene sequences. Taxon 64: 3948.CrossRefGoogle Scholar
Adsersen, H. 1995. Research on islands: classic, recent, and prospective approaches. In Vitousek, P. M., Loope, L. L., and Adsersen, H. (eds.), Islands: Biological and Ecosystem Function (pp. 722). New York: Springer.Google Scholar
AFP. 2010. Tsunami kills 5, 11 missing. The Straits Times, Feb. 28 (online).Google Scholar
Agarwal, M., Shrivastava, N., and Padh, H. 2008. Advances in molecular marker techniques in plant science. Plant Cell Rep. 27: 617631.Google Scholar
Aguirre de Maino, I. M. (ed.). 1981. Chile a Color: Biografías, vol. 1. Santiago: Editorial Antártica.Google Scholar
Agurto, B. R. 1943. Sobre el valor económico de las Islas de Juan Fernández. Licenciado Thesis, Facultad de Ciencias Jurídicas y Sociales, Universidad de Chile, Santiago.Google Scholar
Ahrendt, L. 1961. Berberis and Mahonia, a taxonomic revision. Bot. J. Linn. Soc. 57: 1410.Google Scholar
Akin, W. E. 1991. Global patterns: Climate, Vegetation, and Soils. Norman: University of Oklahoma Press.Google Scholar
Allan, R., Lindesay, J., and Parker, D. 1996. El Niño Southern Oscillation and Climatic Variability. Collingwood, Australia: CSIRO Publishing.Google Scholar
Allendorf, F. W., Luikart, G., and Aitken, S. N. 2013. Conservation and the Genetics of Populations (2nd edn.). West Sussex: Wiley-Blackwell.Google Scholar
Al-Shehbaz, I. A. 1988. The genera of Arabideae (Cruciferae, Brassicaceae) in the southeastern United States. J. Arnold Arbor. 69: 85166.Google Scholar
Alston, R. E., and Turner, B. L. 1963. Biochemical Systematics. Englewood Cliffs, NJ: Prentice-Hall.Google Scholar
Amundsen, K., and Warnke, S. 2012. Agrostis species relationships based on trnL-trnF and atpI-atpH intergenic spacer regions. Hort. Sci. 47: 1824.Google Scholar
Anderson, A., Haberle, S., Rojas, G., Seelenfreund, A., Smith, I., et al. 2002. An archaeological exploration of Robinson Crusoe Island, Juan Fernandez Archipelago, Chile. In Bedford, S., Sand, C., and Burley, D. (eds.), Fifty Years in the Field: Essays in Honour and Celebration of Richard Shutler Jr.’s Archaeological Career (pp. 239249). Auckland, Private Bag: New Zealand Archaeological Association (Monograph 25).Google Scholar
Anderson, G. and Bernardello, G. 2012. Robinson Crusoe principle: the pollination biology of the flora of the Juan Fernández Archipelago and how the species “made do” with what they had. Columbus, OH: Botany 2012 (Abstr.).Google Scholar
Anderson, G., Bernardello, G., and Engel, M. S. 2001a. Conservation implications of a newly discovered bee species on Isla Robinson Crusoe, Chile. Conserv. Biol. 15: 803805.Google Scholar
Anderson, G., Bernardello, G., Opel, M., Santos Guerra, A., and Anderson, M. 2006. Reproductive biology of the dioecious Canary Islands endemic Withania aristata (Solanaceae). Am. J. Bot. 93: 12951305.Google Scholar
Anderson, G., Bernardello, G., Stuessy, T., and Crawford, D.J. 2001b. Breeding system and pollination of selected plants endemic to Juan Fernández Islands. Am. J. Bot. 88: 220233.CrossRefGoogle ScholarPubMed
Anderson, G., Crawford, D. J., and Bernardello, G. 2013. The reproductive biology of island plants: the Juan Fernández and Canary Islands (facilitated by the work of Arnoldo Santos-Guerra). Vieraea 41: 109121.Google Scholar
Anderson, G. J. 1979. Dioecious Solanum of hermaphroditic origin is an example of a broad convergence. Nature 282: 836838.CrossRefGoogle Scholar
Anderson, G. J., Anderson, M. K. J., and Patel, N. 2015. The ecology, evolution and biogeography of dioecy in the genus Solanum: with paradigms from the strong dioecy in Solanum polygamum to the unsuspected and cryptic dioecy in Solanum conocarpum. Am. J. Bot. 102: 471486.CrossRefGoogle Scholar
Anderson, G. J., Bernardello, G., López, P., Crawford, D. J., and Stuessy, T. 2000a. Reproductive biology of Wahlenbergia (Campanulaceae) endemic to Robinson Crusoe Island (Chile). Plant Syst. Evol. 223: 109123.Google Scholar
Anderson, G. J., Bernardello, G., López, P., Stuessy, T., and Crawford, D. J. 2000b. Dioecy and wind pollination in Pernettya rigida (Ericaceae) of the Juan Fernández Islands. Bot. J. Linn. Soc. 132: 121141.CrossRefGoogle Scholar
Anderson, G. J., Bernardello, G., and Santos-Guerra, A. 2014. Reproductive biology of Solanum vespertilio (Solanaceae), a zygomorphic, heterantherous, enantiostylous, and andromonoecious rare Canary Islands endemicPlant Syst. Evol301: 11911206.Google Scholar
Andrus, N., Tye, A., Nesom, G., Bogler, D., Lewis, C., et al. 2009. Phylogenetics of Darwiniothamnus (Asteraceae: Astereae): molecular evidence for multiple origins in the endemic flora of the Galápagos Islands. J. Biogeogr. 36: 10551069.Google Scholar
Anonymous. 1978. Isla Robinson Crusoe, from Surveys by the Chilean Navy, map. No. 22492. Defense Mapping Agency Hydrographic Center, Washington, DC.Google Scholar
APG (Angiosperm Phylogeny Group). 1998. An ordinal classification for the families of flowering plants. Ann. Missouri Bot. Gard. 85: 531553.CrossRefGoogle Scholar
APG (Angiosperm Phylogeny Group). 2003. An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG II. Bot. J. Linn. Soc. 141: 399436.Google Scholar
APG (Angiosperm Phylogeny Group). 2009. An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG III. Bot. J. Linn. Soc. 161: 105121.Google Scholar
APG (Angiosperm Phylogeny Group). 2016. An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG IV. Bot. J. Linn. Soc. 181: 120.Google Scholar
Appelhans, M. C., Wen, J., Wood, K. R., Allan, G. J., Zimmer, E. A., et al. 2014. Molecular phylogenetic analysis of Hawaiian Rutaceae (Melicope, Platydesma and Zanthoxylum) and their different colonization patterns. Bot. J. Linn. Soc. 174: 425448.Google Scholar
Arana, P. 1987. Perspectivas históricas y proyecciones de la actividad pesquera realizada en el Archipiélago de Juan Fernández, Chile. In Castilla, J. C. (ed.), Islas Oceánicas Chilenas; Conocimiento Científico y Necesidades de Investigaciones. Santiago: Universidad Católica de Chile.Google Scholar
Arana, P. 2010. La Isla de Robinson Crusoe. Valparaíso: Ediciones Universitarias de Valparaíso.Google Scholar
Arana E., P. 1985. Análisis y recomendaciones sobre medidas de regulación en la pesquería de la langosta de Juan Fernández (Jasus frontalis). In Arana E., P. (ed.), Investigaciones Marinas en el Archipiélago de Juan Fernández (pp. 291300). Valparaíso: Escuela de Ciencias del Mar, Universidad Católica de Valparaíso.Google Scholar
Archibald, J. K., Crawford, D. J., Santos-Guerra, A., and Mort, M. E. 2006. The utility of automated analysis of inter-simple sequence repeat (ISSR) loci for resolving relationships in the Canary Island species of Tolpis (Asteraceae). Am. J. Bot. 93: 11541162.CrossRefGoogle ScholarPubMed
Arditti, J., and Ghani, A. K. A. 2000. Numerical and physical properties of orchid seeds and their biological implications. New Phytol. 145: 367421.Google Scholar
Arellano, G. 2012. Evaluación de la dinámica de invasión de Aristotelia chilensis (Elaeocarpaceae) y Rubus ulmifolius (Rosaceae) en claros de dosel en un bosque de la isla Robinson Crusoe. Tesis Magíster en Áreas Silvestres y Conservación de la Naturaleza, Universidad de Chile (Santiago), Facultad de Ciencias Forestales y de la Conservacion de la Naturaleza.Google Scholar
Armada de Chile. 1965. Mapa Hidrográfico de Isla Mas a Tierra. Valparaíso: Instituto Hidrográfico.Google Scholar
Astudillo, M. 2014. Geomorfología y evolución geológica de la Isla Robinson Crusoe, Archipiélago Juan Fernández. Thesis, Título de Geología, Depto. de Geología, Universidad de Chile, Santiago.Google Scholar
August-Schmidt, E. M., Haro, G., Borager, A., and D’Antonio, C. M. 2015. Preferential associations of invasive Lantana camara (Verbenaceae) in a seasonally dry Hawaiian woodland. Pacific Sci. 69: 385397.Google Scholar
Bacon, C. D., McKenna, M. J., Simmons, M. P., and Wagner, W. H. 2012. Evaluating multiple criteria for species delimitation: an empirical example using Hawaiian palms (Arecaceae: Pritchardia). BMC Evol. Biol. 2012(12): 23.CrossRefGoogle Scholar
Baeza, C., Bastías, C., Stuessy, T., Ruiz, E., Greimler, J., et al. 2013. Cytological investigations on populations of Taraxacum (Asteraceae) from the Juan Fernández Archipelago, Chile. Gayana Bot. 70: 401403.Google Scholar
Baeza, C., Vosyka, E., and Stuessy, T. 2004. Recuentos cromosómicos en plantas que crecen en Chile. Darwiniana 42: 2529.Google Scholar
Baeza, C. M., Marticorena, C., Stuessy, T., Ruiz, E., and Negritto, M. 2007. Poaceae en el archipiélago de Juan Fernández (Robinson Crusoe). Gayana Bot. 64: 125174.Google Scholar
Baeza, C. M., Rodríguez, R., Hoeneisen, F., and Stuessy, T. 1999. Consideraciones anatómicas del leño secundario de Santalum fernandezianum F. Phil. (Santalaceae), especie extinta del Archipiélago de Juan Fernández, Chile. Gayana Bot. 56: 6365.Google Scholar
Baeza, C. M., Stuessy, T. F., and Marticorena, C. 2002. Notes on the Poaceae of the Robinson Crusoe (Juan Fernández) Islands, Chile. Brittonia 54: 154163.Google Scholar
Bahamonde, N. 1966. Capitulo VIII. El mar y sus recursos. In Anonymous (ed.), Geografía Económica de Chile, Primer Apéndice (pp. 8189). Santiago, Chile: Corporación de Fomento de la Producción (CORFO).Google Scholar
Bailey, C. D., Koch, M. A., Mummenhoff, K., Mayer, M., O’Kane, S. L. Jr., et al. 2006. Toward a global phylogeny of the Brassicaceae. Mol. Biol. Evol. 23: 21422160.CrossRefGoogle Scholar
Bailey, I. W., and Smith, A. C. 1942. Degeneriaceae, a new family of flowering plants from Fiji. J. Arnold Arbor. 23: 356365.CrossRefGoogle Scholar
Baker, H. G. 1955. Self-compatibility and establishment after “long-distance” dispersal. Evolution 9: 347348.Google Scholar
Baker, H. G. 1967. Support for Baker’s law – as a rule. Evolution 21: 853856.Google Scholar
Baker, H. G., and Cox, P. A. 1984. Further thoughts on dioecism and islands. Ann. Missouri Bot. Gard. 71: 244253.Google Scholar
Baker, P. E., Gledhill, A., Harvey, P. K., and Hawkesworth, C. J. 1987. Geochemical evolution of the Juan Fernandez Islands, SE Pacific. J. Geol. Soc. 144: 933944.Google Scholar
Baker, P. E., and Keyvan-Scocouhi, F. A. 1982. Petrology and geochemistry of the Juan Fernández Islands, south east Pacific. In Schmincke, H.-U., Baker, P. E., Forjaz, V. H. (eds.), Proceedings Symposium on the Activity of Ocean Volcanoes (pp. 255264). Ponta Delgada: Universidade do Açores.Google Scholar
Baldwin, B. G. 2003. A phylogenetic perspective on the origin and evolution of Madiinae. In Carlquist, S., Baldwin, B. G., and Carr, G. D. (eds.), Tarweeds & Silverswords: Evolution of the Madiinae (Asteraceae) (pp. 193228). St. Louis, MO: Missouri Botanical Garden Press.Google Scholar
Baldwin, B. G., Crawford, D. J., Francisco-Ortega, J., Kim, S.-C., Sang, T., et al. 1998. Molecular phylogenetic insights on the origin and evolution of oceanic island plants. In Soltis, D. E., Soltis, P., and Doyle, J. J. (eds.), Molecular Systematics of Plants, vol. II: DNA Sequencing (pp. 410441). New York: Kluwer Academic.CrossRefGoogle Scholar
Ball, P. W. 1990. Some aspects of the phytogeography of Carex. Canad. J. Bot. 68: 14621472.Google Scholar
Barker, W. R. 1982. Taxonomic studies in Euphrasia L. (Scrophulariaceae): a revised infrageneric classification, and a revision of the genus in Australia. J. Adelaide Bot. Gard. 5: 1304.Google Scholar
Barrera, M. E. 1997. Helechos de Juan Fernández. Mus. Nac. Hist. Nat. Chile Publ. Ocas. 51: 1103.Google Scholar
Barrera, V. H. 1963. Cambio de nombre a las islas de Juan Fernández. Rev. Chil. Hist. Geograf. 131: 425430.Google Scholar
Barrett, S. C. H. 1995. Mating-system evolution in flowering plants: micro- and macroevolutionary approaches. Acta Bot. Neerl. 44: 385402.Google Scholar
Barría, A. 1996. Incendio penetró bosque nativo en Juan Fernández. El Mercurio, February 22.Google Scholar
Barrier, M., Baldwin, B. G., Robichaux, R. H., and Purugganan, M. D. 1999. Interspecific hybrid ancestry of a plant adaptive radiation: allopolyploidy of the Hawaiian silversword alliance (Asteraceae) inferred from floral homeotic gene duplications. Mol. Biol. Evol. 16: 11051113.Google Scholar
Barrington, D. 2014. A world-level phylogenetic and biogeographic perspective on the fern genus Polystichum. Boise, ID: Botany 2014 (Abstr.).Google Scholar
Barros Arana, D. 1904. El Doctor Don Rodolfo Amando Philippi. Su Vida i Sus Obras (1808–1904). Santiago: Imprenta Cervantes.Google Scholar
Bate-Smith, E. C. 1958. Plant phenolics as taxonomic guides. Proc. Linn. Soc. 169: 198211.Google Scholar
Bate-Smith, E. C. 1962. The phenolic constituents of plants and their taxonomic significance: I. Dicotyledons. Bot. J. Linn. Soc. 58: 7395.Google Scholar
Beardsley, P. M., and Olmstead, R. G. 2002. Redefining Phrymaceae: the placement of Mimulus, tribe Mimuleae, and Phryma. Am. J. Bot. 89: 10931102.CrossRefGoogle Scholar
Beardsley, P. M., Schoenig, S. E., Whittall, J. B., and Olmstead, R. G. 2004. Patterns of evolution in western North American Mimulus (Phrymaceae). Am. J. Bot. 91: 474489.Google Scholar
Beheregaray, L. B., Gibbs, J. P., Havell, N., Fritts, T. H., Powell, J. R., et al. 2004. Giant tortoises are not so slow: rapid diversification and biogeographic consensus in the Galápagos. Proc. Natl. Acad. Sci. USA 101: 65146519.Google Scholar
Bellin, J.-N. 1764. Le Petit Atlas Maritime, vol. 2 (map 68). Paris.Google Scholar
Bennett, J. R., and Mathews, S. 2006. Phylogeny of the parasitic plant family Orobanchaceae inferred from phytochrome A. Am. J. Bot. 93: 10391051.CrossRefGoogle ScholarPubMed
Bernardello, G., Aguilar, R., and Anderson, G. J. 2004. The reproductive biology of Sophora fernandeziana (Leguminosae), a vulnerable endemic species from Isla Robinson Crusoe. Am. J. Bot. 91: 198206.Google Scholar
Bernardello, G., Anderson, G., Lopez, P., Cleland, M. A., Stuessy, T. F., et al. 1999. Reproductive biology of Lactoris fernandeziana (Lactoridaceae). Am. J. Bot. 86: 829840.CrossRefGoogle ScholarPubMed
Bernardello, G., Anderson, G. J., Stuessy, T. F., and Crawford, D. J. 2001. A survey of floral traits, breeding systems, floral visitors, and pollination systems of the angiosperms of the Juan Fernández Islands (Chile). Bot. Rev. (Lancaster) 67: 255308.Google Scholar
Bernardello, G., Anderson, G. J., Stuessy, T. F., and Crawford, D. J. 2006. The angiosperm flora of the Archipelago Juan Fernández (Chile): origin and dispersal. Can. J. Bot. 84: 12661281.Google Scholar
Bernardello, G., Galetto, L., and Anderson, G. J. 2000. Floral nectary structure and nectar chemical composition of some species from Robinson Crusoe Island (Chile). Can. J. Bot. 78: 862872.Google Scholar
Berry, E. W. 1911. A revision of the fossil ferns from the Potomac Group which have been referred to the genera Cladophlebis and Thyrsopteris. Proc. US Nat. Mus. 41: 307332.Google Scholar
Bertero, C. G. L. 1830. Notice sur l’histoire naturelle de l’ile Juan Fernandez, extraite d’une lettre de M. Bertero. Ann. Sci. Nat. 21 : 344351.Google Scholar
Beurton, C. 1994. Gynoecium and perianth in Zanthoxylum s.l. (Rutaceae). Plant Syst. Evol. 189: 165191.Google Scholar
Biffin, E., Lucas, E. J., Craven, L. A., Ribeiro da Costa, I., Harrington, M. G., and Crisp, M. D. 2010. Evolution of exceptional species richness among lineages of fleshy-fruited Myrtaceae. Ann. Bot. 106: 7993.Google Scholar
Biodiversa. 2009a. Archipiélago Juan Fernández. Sitio Prioritario para la Conservación de la Biodiversidad Global. Sistematización del Estado Actual del Conocimiento. Santiago: DIPROREN-SAG.Google Scholar
Biodiversa. 2009b. Priorización y Definición de Objectos de Conservación y Líneas de Acción para la Elaboración del Plan de Conservación de la Biodiversidad Terrestre del Archipiélago Juan Fernández. Santiago: DIPROREN-SAG.Google Scholar
Bitter, G. 1910. Die Gattung Acaena. Vorstudien zu einer Monographie. Bibliothec. Bot. 74: 1336.Google Scholar
Bitter, G. 1921. Ein Gattungsbastard zwischen Acaena und Margyricarpus: Margyracaena. Feddes Repert. Spec. Nov. Regni Veg. 17: 239243.CrossRefGoogle Scholar
Bittner, M., Silva, M., Rozas, Z., Jakupovic, J., and Stuessy, T. 1994. Estudio químico del género Gunnera en Chile: II. Metabolitos secundários de dos especies continentales y dos especies de las islas de Juan Fernández. Bol. Soc. Chil. Quim. 39: 7983.Google Scholar
Boeger, M. R. T., Cavichiolo, L. E., Pil, M. W., and Labiak, P. H. 2007. Variabilidade fenotípica de Rumohra adiantiformis (G.Forst) Ching (Dryopteridaceae). Hoehnea 34: 553561.Google Scholar
Böhle, U. R., Hilger, H. H., and Martin, W. F. 1996. Island colonization and evolution of the insular woody habit in Echium L. (Boraginaceae). Proc. Natl. Acad. Sci. USA 93: 1174011745.CrossRefGoogle ScholarPubMed
Bohm, B. A. 1998a. Introduction to Flavonoids. Amsterdam: Harwood Academic.Google Scholar
Bohm, B. A. 1998b. Secondary compounds and evolutionary relationships of island plants. In Stuessy, T. F. and Ono, M. (eds.), Evolution and Speciation of Island Plants (pp. 233295). Cambridge University Press.Google Scholar
Bohm, B. A., and Fong, C. 1990. Non-polar flavonoids of Wilkesia and Argyroxiphium. Phytochemistry 29: 11751177.Google Scholar
Bohm, B. A., and Stuessy, T. F. 2001. Flavonoids of the Sunflower Family (Asteraceae). Wien: Springer-Verlag.Google Scholar
Bohm, B. A., and Yang, J. Y. 2003. Secondary metabolites of Madiinaae and the tarweeds. In Carlquist, S., Baldwin, B. G., and Carr, G. D. (eds.), Tarweeds & Silverswords: Evolution of the Madiinae (Asteraceae) (pp. 145170). St. Louis, MO: Missouri Botanical Garden Press.Google Scholar
Bolnick, D. I., and Fitzpatrick, B. M. 2007. Sympatric speciation: models and empirical evidence. Ann. Rev. Ecol. Evol. Syst. 38: 459487.Google Scholar
Bonin, A., Ehrich, D., and Manel, S. 2007. Statistical anaysis of amplified fragment length polymorphism data: a toolbox for molecular ecologists and evolutionists. Mol. Ecol. 16: 37373758.Google Scholar
Booker, J., Bullard, E. C., and Grasty, R. L. 1967. Palaeomagnetism and age of rocks from Easter Island and Juan Fernandez. Geophys. J. R. Astronom. Soc. 12: 469471.Google Scholar
Borgen, L. 1976. Analysis of a hybrid swarm between Argyranthemum adauctum and A. filifolium in the Canary Islands. Norweg. J. Bot. 23: 121137.Google Scholar
Borregaard, M. K., Matthews, T. J., and Whittaker, R. J. 2016. The general dynamic model: towards a unified theory of island biogeography? Global Ecol. Biogeogr. 25: 805816.Google Scholar
Bottini, M. C. J., De Bustos, A., Jouve, N., and Poggio, L. 2002. AFLP characterization of natural populations of Berberis (Berberidaceae) in Patagonia, Argentina. Plant Syst. Evol. 231: 133142.Google Scholar
Bottini, M. C. J., De Bustos, A., Sanso, A. M., Jouve, N., and Poggio, L. 2007. Relationships in Patagonian species of Berberis (Berberidaceae) based on the characterization of rDNA internal transcribed spacer sequences. Bot. J. Linn. Soc. 153: 321328.Google Scholar
Bourne, W. R. P., Brooke, M. de L., Clark, G. S., and Stone, T. 1992. Wildlife conservation problems in the Juan Fernández Archipelago, Chile. Oryx 26: 4351.Google Scholar
Bouzat, J. L. 2010. Conservation genetics of population bottlenecks: the role of chance, selection, and history. Conserv. Genet. 11: 463478.Google Scholar
Bozek, M., Leitch, A. R., Leitch, I. J., Záveská Drábková, L., and Kuta, E. 2012. Chromosome and genome size variation in Luzula (Juncaceae), a genus with holocentric chromosomes. Bot. J. Linn. Soc. 170: 529541.Google Scholar
Bramwell, D. and Caujapé-Castells, J. (eds.). 2011. The Biology of Island Floras. Cambridge: Cambridge University Press.Google Scholar
Bräuchler, C., Meimberg, H., and Heubl, G. 2010. Molecular phylogeny of Menthinae (Lamiaceae, Nepetoideae, Mentheae): taxonomy, biogeography and conflicts. Mol. Phylogenet. Evol. 55: 501523.Google Scholar
Braun-Blanquet, J. 1928. Pflanzensoziologie, Grundzüge der Vegetationskunde. Berlin: Springer.Google Scholar
Braun-Blanquet, J. 1964. Pflanzensoziologie. Grundzüge der Vegetationskunde (3rd edn.). Wien: Springer.Google Scholar
Brauner, S., Crawford, D. J., and Stuessy, T. F. 1992. Ribosomal and RAPD variation in the rare plant family Lactoridaceae. Am. J. Bot. 79: 14361439.Google Scholar
Bremer, K. 1994. Asteraceae: Cladistics and Classification. Portland, OR: Timber Press.Google Scholar
Briquet, J. 1935. International Rules of Botanical Nomenclature … Revised by the International Botanical Congress of Cambridge 1930. Jena: Gustav Fischer.Google Scholar
Britton, N. L. 1901. Manual of the Flora of the Northern States and Canada. New York: Henry Holt.Google Scholar
Brizicky, G. K. 1962. Taxonomic and nomenclatural notes on Zanthoxylum and Glycosmis (Rutaceae). J. Arnold Arbor. 43: 8093.CrossRefGoogle Scholar
Brochmann, C., Borgen., L., and Stabbetorp, O. E. 2000. Multiple diploid hybrid speciation of the Canary Island endemic Argyranthemum sundingii (Asteraceae). Plant Syst. Evol. 220: 7792.Google Scholar
Brooke, M. de L. 1987a. The Birds of the Juan Fernández Islands, Chile. Cambridge: International Council for Bird Preservation, Fauna and Flora Preservation Society, World Wildlife Fund.Google Scholar
Brooke, M. de L. 1987b. Population estimates and breeding biology of the petrels Pterodroma externa and P. longirostris on Isla Alejandro Selkirk, Juan Fernandez Archipelago. Condor 89: 581586.Google Scholar
Brotherus, V. F. 1924. The musci of the Juan Fernández Islands. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 409448). Uppsala: Almqvist & Wiksells.Google Scholar
Brüggen, J. O. 1950. Fundamentos de la Geología de Chile. Santiago: Instituto Geográfico Militar.Google Scholar
Bruhl, J. J. 1995. Sedge genera of the world: relationships and a new classification of the Cyperaceae. Austral. Syst. Bot. 8: 125305.Google Scholar
Brummitt, R. K. 2014. Taxonomy versus cladonomy in the dicot families. Ann. Missouri Bot. Gard. 100: 8999.Google Scholar
Brys, R., Vanden Broeck, A., Mergeay, J., and Jacquemyn, H. 2013. The contribution of mating system variation to reproductive isolation in two closely related Centaurium species (Gentianaceae) with a generalized flower morphology. Evolution 68: 12811293.Google Scholar
Buckley, R. C., and Knedlhans, S. B. 1986. Beachcomber biogeography: interception of dispersing propagules by islands. J. Biogeogr. 13: 6970.CrossRefGoogle Scholar
Burrows, C. J. 1994. The seeds always know best. New Zealand J. Bot. 32: 349363.Google Scholar
Bush, C. M., Lu, L., Fritsch, P. W., Li, D.-Z., and Kron, K. A. 2009. Phylogeny of Galutherieae (Ericaceae: Vaccinioideae) based on DNA sequence data from matK, ndhF, and nrITS. Int. J. Plant Sci. 170: 355364.Google Scholar
Butlin, R. K., Galindo, J., and Grahame, J. W. 2008. Sympatric, parapatric or allopatric: the most important way to classify speciation? Philos. Trans. R. Soc. Lond. B. 363: 29973007.Google Scholar
Cabrera, A. L. 1949. El género “Senecio” en Chile. Lilloa 15: 27501.Google Scholar
Callander, J. 1768. Terra Australis Cognita, or Voyages to the Terra Australis, or Southern Hemisphere, during the Sixteenth, Seventeenth, and Eighteenth Centuries, vol. 2. Edinburgh (see pp. 528534 for Capt. Cowley, A.).Google Scholar
Calviño, C. I., Martínez, S. G., and Downie, S. R. 2008a. The evolutionary history of Eryngium (Apiaceae, Saniculoideae): rapid radiations, long distance dispersals, and hybridizations. Mol. Phylogenet. Evol. 46: 11291150.Google Scholar
Calviño, C. I., Martínez, S. G., and Downie, S. R. 2008b. Morphology and biogeography of Apiaceae subfamily Saniculoideae as inferred by phylogenetic analysis of molecular data. Am. J. Bot. 95: 196214.Google Scholar
Calviño, C. I., Martínez, S. G., and Downie, S. R. 2010. Unraveling the taxonomic complexity of Eryngium L. (Apiaceae, Saniculoideae): phylogenetic analysis of 11 non-coding cpDNA loci corroborates rapid radiations. Plant Div. Evol. 128: 137149.Google Scholar
Camus, P., Castro, S., and Jaksic, F. 2008. El conejo europeo en Chile: historia de una invasión biológica. Historia 41: 305339.Google Scholar
Candolle, A. P. de. 1833. Genres nouveaux appartenant a la famille des Composées ou Synantherées. Arch. Bot. 2: 330334.Google Scholar
Candolle, A. P. de. 1836. Prodromus Systematis Naturalis Regni Vegetabilis, vol. 5. Paris: Treuttel et Würtz.Google Scholar
Candolle, A. P. 1838. Balbisia and Robinsonia. In Prodromus Systematis Naturalis Regni Vegetabilis, Part 6 (pp. 447448). Paris: Treuttel et Würtz.Google Scholar
Cantley, J. T., Swenson, N. G., Markey, A., and Keeley, S. C. 2014. Biogeographic insights on Pacific Coprosma (Rubiaceae) indicate two colonizations to the Hawaiian Islands. Bot. J. Linn. Soc. 174: 412424.Google Scholar
Carlquist, S. 1964. Morphology and relationships of Lactoridaceae. Aliso 5: 421435.Google Scholar
Carlquist, S. 1965. Island Life. Garden City, NY: Natural History Press.Google Scholar
Carlquist, S. 1966. The biota of long-distance dispersal: IV. Genetic systems in the floras of oceanic islands. Evolution: 20: 433455.Google Scholar
Carlquist, S. 1967. Anatomy and systematics of Dendroseris (sensu lato). Brittonia 19: 99121.Google Scholar
Carlquist, S. 1970. Wood anatomy of insular species of Plantago and the problem of raylessness. Bull. Torrey Bot. Club 97: 353361.Google Scholar
Carlquist, S. 1974. Island Biology. New York: Columbia University Press.Google Scholar
Carlquist, S. 1980. Hawaii: A Natural History (2nd edn.). Honolulu: Pacific Tropical Botanical Garden.Google Scholar
Carlquist, S. 1981. Chance dispersal. Am. Sci. 69: 509516.Google Scholar
Carlquist, S. 1990. Wood anatomy and relationships of Lactoridaceae. Am. J. Bot. 77: 14981505.Google Scholar
Carlquist, S. 1992. Wood anatomy and stem of Chloranthus: summary of wood anatomy of Chloranthaceae, with comments on relationships, vessellessness, and the origin of monocotyledons. IAWA Bull. 13: 316.Google Scholar
Carlquist, S. 2009. Darwin on island plants. Bot. J. Linn. Soc. 161: 2025.Google Scholar
Carlsen, T., Bleeker, W., Hurka, H., Elven, R., and Brochmann, C. 2009. Biogeography and phylogeny of Cardamine (Brassicaceae). Ann. Missouri Bot. Gard. 96: 215236.Google Scholar
Caro, T., Darwin, J., Forrester, T., Ledoux-Bloom, C., and Wells, C. 2011. Conservation in the Anthropocene. Conserv. Biol. 26: 185188.Google Scholar
Carpenter, K. J. 1986. The History of Scurvy and Vitamin C. Cambridge University Press.Google Scholar
Carr, G. D. 1985. Monograph of the Hawaiian Madiinae (Asteraceae): Argyroxiphium, Dubautia, and Wilkesia. Allertonia 4: 1123.Google Scholar
Carr, G. D. 1998. Chromosome evolution and speciation in Hawaiian flowering plants. In Stuessy, T. F. and Ono, M. (eds.), Evolution and Speciation of Island Plants (pp. 547). Cambridge University Press.Google Scholar
Carr, G. D. 2003. Hybridization in Madiinae. In Carlquist, S., Baldwin, B. G., and Carr, G. D. (eds.), Tarweeds & Silverswords: Evolution of the Madiinae (Asteraceae) (pp. 79104). St. Louis: Missouri Botanical Garden Press.Google Scholar
Carr, G. D., and Kyhos, D. W. 1981. Adaptive radiation in the Hawaiian silversword alliance (Compositae-Madiinae): I. Cytogenetics of spontaneous hybrids. Evolution 35: 543556.Google Scholar
Carr, G. D., and Kyhos, D. W. 1986. Adaptive radiation in the Hawaiian silversword alliance (Compositae-Madiinae): II. Cytogenetics of artificial and natural hybrids. Evolution 40: 959976.Google Scholar
Castilla, J. C., and Oliva, D. 1987. Islas oceánicas Chilenas: aspectos descriptivos y potencialidades. In Castilla, J. C. (ed.), Islas Oceánicas Chilenas: Conocimiento Científico y Necesidades de Investigaciones (pp. 1735). Santiago: Ediciones Universidad Católica de Chile.Google Scholar
Castro, M., and Rosselló, J. A. 2007. Karyological observations on plant taxa endemic to the Balearic Islands. Bot. J. Linn. Soc. 153: 463476.Google Scholar
Caujapé-Castells, J. 2010. General Gst and Ø inflation due to biased intra-population sampling, and its consequences for the conservation of the Canarian Flora. Conserv. Genet. 11: 709720.Google Scholar
Caujapé-Castells, J. 2011. Jesters, red queens, boomerangs and surfers: a molecular outlook on the diversity of the Canarian endemic flora. In Bramwell, D. and Caujapé-Castells, J. (eds.), The Biology of Island Floras (pp. 284324). Cambridge University Press.Google Scholar
Caujapé-Castells, J., Marrero-Rodríguez, A., Baccarani-Rosas, M., Carrera-García, N., and Vilches-Navarrete, B. 2008a. Population genetics of the endangered Canarian endemic Atractylis arbuscula (Asteraceae): implications for taxonomy and conservation. Plant Syst. Evol. 273: 99109.Google Scholar
Caujapé-Castells, J., Naranjo-Suárez, J., Santanna, I., Baccarani-Rosas, M., Carrera-García, N., et al. 2008b. Population genetic suggestions to offset the extinction ratchet in the endangered Canarian endemic Atractylis preauxiana (Asteraceae). Plant Syst. Evol. 273: 191199.Google Scholar
Caujapé-Castells, J., Sabbagh, I., Castllano, J. J., Ramos, R., Henriquez, V., et al. 2013. Transformer-4 version 2.0.1, a free multi-platform software to quickly reformat genotype matrices of any marker type, and archive them in the Demiurge information system. Mol. Ecol. Resour. 13: 484493.Google Scholar
Caujapé-Castells, J., Tye, A., Crawford, D. J., Santos-Guerra, A., Sakai, A., et al. 2010. Conservation of oceanic island floras: present and future global challenges. Perspect. Plant Ecol. Evol. Syst. 12: 107129.Google Scholar
Cavieres, A., Mellado, F., Tramón, S., and Rebolledo, S. 2011. Plan de Conservación. Biodiversidad Terrestre. Archipiélago Juan Fernández. Santiago: Fundación Biodiversa.Google Scholar
Cereceda, P., Schemenauer, R. S., and Zunino, H. 1994. Distribución de la precipitación en isla Robinson Crusoe. Rev. Geografía Norte Grande 21: 3336.Google Scholar
Chamorro, S., Ruben, H., Olesen, J., McMullen, C., and Traveset, A. 2012. Pollination patterns and plant breeding systems in the Galápagos: a review. Ann. Bot. 110: 14891501.Google Scholar
Chao, Y.-S., Rouhan, G., Amoroso, V. B., and Chiou, W.-L. 2014. Molecular phylogeny and biogeography of the fern genus Pteris (Pteridaceae). Ann. Bot. 114: 109124.Google Scholar
Chase, M. W., Knapp, S., Cox, A. V., Clarkson, J. J., Butsko, Y., et al. 2003. Molecular systematics, GISH and the origin of hybrid taxa in Nicotiana (Solanaceae). Ann. Bot. 92: 107127.Google Scholar
Chase, M. W., Morton, C. M., and Kallunki, J. A. 1999. Phylogenetic relationships of Rutaceae: a cladistic analysis of the subfamilies using evidence from rbcL and atpB sequence variation. Am. J. Bot. 86: 11911199.Google Scholar
Chase, M. W., Zmarzty, S., Lledo, M. D., Wurdack, K. J., Swensen, S. M., et al. 2002. When in doubt, put it in Flacourtiaceae: a molecular phylogenetic analysis based on plastid rbcL DNA sequences. Kew Bull. 57: 141181.Google Scholar
Chemisquy, M. A. 2009. Novedades nomenclaturales en el género Gavilea (Orchidaceae, Chloraeinae), con especial énfasis en las especies chilenas. Darwiniana 47: 315320.Google Scholar
Chemisquy, M. A., and Morrone, O. 2010. Phylogenetic analysis of the subtribe Chloraeinae (Orchidaceae): a preliminary approach based on three chloroplast markers. Austral. Syst. Bot. 23: 3846.Google Scholar
Chemisquy, M. A., and Morrone, O. 2012. Molecular phylogemy of Gavilea (Chloraeinae: Orchidaceae) using plastid and nuclear markers. Mol. Phylogenet. Evol. 62: 889897.CrossRefGoogle ScholarPubMed
Christenhusz, M. J. M., and Chase, M. W. 2014. Trends and concepts in fern classification. Ann. Bot. 113: 571594.Google Scholar
Christenhusz, M. J. M., Zhang, X.-C., and Schneider, H. 2011. A linear sequence of extant families and genera of lycophytes and ferns. Phytotaxa 19: 754.Google Scholar
Christensen, C. 1910. On some species of ferns collected by Dr. Carl Skottsberg in temperate South America. Arkiv för Bot. 10: 132.Google Scholar
Christensen, C., and Skottsberg, C. 1920. The Pteridophyta of the Juan Fernandez Islands. In Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 2 (pp. 146). Uppsala: Almqvist & Wiksells.Google Scholar
Christophersen, E. 1968. Flowering plants from Tristan da Cunha. Results Norweg. Scient. Exped. Tristan da Cunha 1937–1938, 55: 129.Google Scholar
Cisternas, M. A., Salazar, G. A., Verdugo, G., Novoa, P., Calderon, X., et al. 2012. Phylogenetic analysis of Chloraeinae (Orchidaceae) based on plastid and nuclear DNA sequences. Bot. J. Linn. Soc. 168: 258277.Google Scholar
Clark, L. G. 1989. Systematics of Chusquea section Swallenochloa, section Verticillatae, section Serpentes, and section Longifoliae (Poaceae-Bambusoideae). Syst. Bot. Monogr. 27: 1127.Google Scholar
Clarke, T. 2001. Searching for Crusoe: A Journey Among the Last Real Islands. New York: Ballantine Books.Google Scholar
Clarkson, J. J., Knapp, S., Garcia, V. F., Olmstead, R. G., Leitch, A. R., et al. 2004. Phylogenetic relationships in Nicotiana (Solanaceae) inferred from multiple plastid DNA regions. Mol. Phylogenet. Evol. 33: 7590.Google Scholar
Coates, D. J., and Byrne, M. 2005. Genetic variation in plant populations: assessing cause and pattern. In Henry, R. J. (ed.), Plant Diversity and Evolution: Genotypic and Phenotypic Variation in Higher Plants (pp. 139164). Wallingford: CAB International.Google Scholar
Cody, M. 2006. Plants on Islands: Diversity and Dynamics on a Continental Archipelago. Berkeley: University of California Press.Google Scholar
Cody, M. L., and Overton, J. M. 1996. Short-term evolution of reduced dispersal in island plant populations. J. Ecol. 84: 5361.Google Scholar
Cole, F. R., Medeiros, A. C., Loope, L. L., and Zuehlke, W. 1992. Effects of the Argentine ant on arthropod fauna of Hawaiian high-elevation shrubland. Ecology 73: 13131322.Google Scholar
Colla, A. 1835. Plantae rariores in regionibus Chilensibus a clarissimo M. D. Bertero nuper detectae. Mem. Reale Accad. Sci. Torino 38: 142.Google Scholar
Colla, A. 1836. Herbarium Pedemontanum Justa Methodum Naturalem Dispositum Additis nonnullis Stirpibus Exoticis ad Universos ejusdem Methodi Ordines Exhibendos, vol. 6. Torino (cf. pp. 215216).Google Scholar
Colwell, R. K. 1989. Hummingbirds of the Juan Fernández Islands: natural history, evolution and population status. Ibis 131: 548566.Google Scholar
CONAF. 1976. Plan de Manejo Parque Nacional Juan Fernández. Santiago: Organización de las Naciones Unidas para la Agricultura y la Alimentación.Google Scholar
CONAF. 2009. Plan de Manejo Parque Nacional Archipiélago Juan Fernández. Valparaíso: Gobierno de Chile, Ministerio de Agricultura.Google Scholar
Conner, J. K., and Hartl, D. L. 2004. A Primer of Ecological Genetics. Sunderland, MA: Sinauer Associates.Google Scholar
Constance, L. 1964. Systematic botany: an unending synthesis. Taxon 13: 257273.Google Scholar
Contreras-M., A., and Spooner, D. M. 1999. Revision of Solanum section Etuberosum (subgenus Potatoe). In Nee, M., Symon, D. E., Lester, R. N., and Jessop, J. P. (eds.), Solanaceae IV (pp. 227245). Kew: Royal Botanic Gardens.Google Scholar
Cooke, E. 1712. A Voyage to the South Seas and Round the World Performed in the Years 1708, 1709, 1710 and 1711. London.Google Scholar
Cordingly, D. 2007. Cochrane the Dauntless. The Life and Adventures of Admiral Thomas Cochrane, 1775–1860. London: Bloomsbury.Google Scholar
Correa, M. N. 1956. Las especies argentinas del género Gavilea. Bot. Soc. Argent. Bot. 6: 7386.Google Scholar
Correa, M. N. 1966. Una especie nueva y algunas observaciones críticas del género Gavilea (Orchidaceae). Bol. Soc. Argent. Bot. 11: 6066.Google Scholar
Correa, M. N. 1968. Una nueva especie del género Gavilea (Orchidaceae). Rev. Mus. de la Plata, secc. Bot. 11: 7577.Google Scholar
Cowie, R. H., and Holland, B. S. 2006. Dispersal is fundamental to biogeography and the evolution of biodiversity on oceanic islandsJ. Biogeogr33: 193198.Google Scholar
Cox, C. B., and Moore, P. D. 1993. Biogeography: An Ecological and Evolutionary Approach (5th edn.). Oxford: Blackwell Science.Google Scholar
Coyne, J. A., and Orr, H. A. 2004. Speciation. Sunderland, MA: Sinauer Associates.Google Scholar
Crawford, D. J. 1978. Flavonoid chemistry and angiosperm evolution. Bot. Rev. 44: 431456.Google Scholar
Crawford, D. J. 1983. Phylogenetic and systematic inferences from electrophoretic studies. In Tanksley, S. D. and Orton, T. J. (eds.), Isozymes in Plant Genetics and Breeding, Part A (pp. 257287). Amsterdam: Elsevier.Google Scholar
Crawford, D. J. 1985. Electrophoretic data and plant speciation. Syst. Bot. 10: 405416.Google Scholar
Crawford, D. J. 1989. Enzyme electrophoresis and plant systematics. In Soltis, D. E. and Soltis, P. S. (eds.), Isozymes in Plant Biology (pp. 146164). Portland, OR: Dioscorides.Google Scholar
Crawford, D. J. 1990. Plant Molecular Systematics: Macromolecular Approaches. New York: Wiley.Google Scholar
Crawford, D. J. 2010. Progenitor-derivative species pairs and plant speciation. Taxon 59: 14131423.Google Scholar
Crawford, D. J. 2014. Genetic diversity in plant populations. In Stuessy, T. F., Crawford, D. J., Soltis, D. E., and Soltis, P. S. (eds.), Plant Systematics: The Origin, Interpretation, and Ordering of Plant Biodiversity (pp. 8598). Königstein: Koeltz Scientific Books.Google Scholar
Crawford, D. J., Anderson, G. J., and Bernardello, G. 2011. The reproductive biology of island plants. In Bramwell, D. and Caujapé-Castells, J. (eds.), The Biology of Island Floras (pp. 1136). Cambridge: Cambridge University Press.Google Scholar
Crawford, D. J., Anderson, G. J., Bernardello, G., and Santos-Guerra, A. 2013. The peproductive biology of island plants: the Juan Fernández and Canary Islands. In Caujapé-Castells, J., Niet, G., and Fernández-Palacios, J. M. (eds.), Proceedings of the Amurga International Conferences on Island Biodiversity 2011: Present and Emerging Knowledge on the Evolution, Diversity and Conservation of the Canarian Flora (pp. 166175). Las Palmas de Gran Canaria: Fundación Canaria Amurga-Maspalomas, Jardin Botanico Canario “Viera y Clavijo.”Google Scholar
Crawford, D. J., Anderson, G. J., Silva, L. B., Menezes de Sequeira, M., Moura, M., et al. 2015. Breeding systems in Tolpis (Asteraceae) in the Macaronesian islands: the Azores, Madeira and the Canaries. Plant Syst. Evol. 301: 11911206.Google Scholar
Crawford, D. J., Archibald, J. K., Mort, M. E., and Santos-Guerra, A. 2009. Pollen fertility of synthetic intra- and interspecific F1 hybrids in Canary Island Tolpis (Asteraceae). Plant Syst. Evol. 279: 103113.Google Scholar
Crawford, D. J., Brauner, S., Cosner, M. B., and Stuessy, T. F. 1993a. Use of RAPD markers to document the origin of the intergeneric hybrid ×Margyracaena skottsbergii (Rosaceae) on the Juan Fernandez Islands. Am. J. Bot. 80: 8992.Google Scholar
Crawford, D. J., Lowrey, T. K., Anderson, G. J., Bernardello, G., Santos-Guerra, A., et al. 2009a. Genetic diversity in the colonizing ancestors of Asteraceae endemic to oceanic islands: Baker’s law and polyploidy. In Funk, V. A., Susanna, A., Stuessy, T. F., and Bayer, R. (eds.), Systematics, Evolution, and Biogeography of the Compositae (pp. 137151). Vienna: International Association of Plant TaxonomyGoogle Scholar
Crawford, D. J., Ruiz, E., Stuessy, T. F., Tepe, E., Aqueveque, P., et al. 2001b. Allozyme diversity in the endemic flowering plant species of the Juan Fernández Archipelago, Chile: ecological and historical factors with implications for conservation. Am. J. Bot. 88: 21952203.Google Scholar
Crawford, D. J., Sang, T., Stuessy, T. F., Kim, S.-C., and Silva O., M. 1998. Dendroseris (Asteraceae: Lactuceae) and Robinsonia (Asteraceae: Senecioneae) on the Juan Fernández Islands: similarities and differences in biology and phylogeny. In Stuessy, T. F. and Ono, M. (eds.), Evolution and Speciation of Island Plants (pp. 97119). Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Crawford, D. J., and Stuessy, T. F. 1997. Plant speciation on oceanic islands. In Iwatsuki, K. and Raven, P. H. (eds.), Evolution and Diversification of Land Plants (pp. 249267). Tokyo: Springer-Verlag.Google Scholar
Crawford, D. J., and Stuessy, T. F. 2016. Cryptic variation, molecular data, and the challenge of conserving plant diversity in oceanic archipelagos: the critical role of plant systematics. Korean J. Plant Taxon. 46: 129148.Google Scholar
Crawford, D. J., Stuessy, T. F., Cosner, M. B., Haines, D. W., and Silva O., M. 1993b. Ribosomal and chloroplast DNA restriction site mutations and the radiation of Robinsonia (Asteraceae, Senecioneae) on the Juan Fernández Islands. Plant Syst. Evol. 184: 233239.Google Scholar
Crawford, D. J., Stuessy, T. F., Cosner, M. B., Haines, D. W., Silva, O., , M., et al. 1992a. Evolution of the genus Dendroseris (Asteraceae: Lactuceae) on the Juan Fernández Islands: evidence from chloroplast and ribosomal DNA. Syst. Bot. 17: 676682.Google Scholar
Crawford, D. J., Stuessy, T. F., Cosner, M. B., Haines, D., Wiens, D., et al. 1994. Lactoris fernandeziana (Lactoridaceae) of the Juan Fernández Islands: allozyme uniformity and field observations. Conserv. Biol. 8: 277280.Google Scholar
Crawford, D. J., Stuessy, T. F., Haines, D. W., Cosner, M. B., Silva O., M., et al. 1992b. Allozyme diversity within and divergence among four species of Robinsonia (Asteraceae: Senecioneae), a genus endemic to the Juan Fernandez Islands, Chile. Am. J. Bot. 79: 962966.Google Scholar
Crawford, D. J., Stuessy, T. F., Lammers, T. G., Silva O., M., and Pacheco, P. 1990. Allozyme variation and evolutionary relationships among three species of Wahlenbergia (Campanulaceae) in the Juan Fernández Islands. Bot. Gaz. 151: 119124.Google Scholar
Crawford, D. J., Stuessy, T. F., Rodríguez, R., and Rondinelli, M. 1993c. Genetic diversity in Rhaphithamnus venustus (Verbenaceae), a species endemic to the Juan Fernández Islands. Bull. Torrey Bot. Club. 120: 2328.Google Scholar
Crawford, D. J., Stuessy, T. F., and Silva, M. 1986. Leaf flavonoid chemistry and the relationships of Lactoridaceae. Plant Syst. Evol. 153: 133139.Google Scholar
Crawford, D. J., Stuessy, T. F., and Silva O., M. 1987a. Allozyme divergence and the evolution of Dendroseris (Compositae: Lactuceae) on the Juan Fernández Islands. Syst. Bot. 12: 435443.Google Scholar
Crawford, D. J., Stuessy, T. F., and Silva O., M. 1988. Allozyme variation in Chenopodium sanctae-clarae, an endemic species of the Juan Fernández Islands, Chile. Biochem. Syst. Ecol. 16: 279284.Google Scholar
Crawford, D. J., Tago-Nakazawa, M., Stuessy, T. F., Anderson, G. J., Bernardello, G., et al. 2001a. Intersimple sequence repeat (ISSR) variation in Lactoris fernandziana (Lactoridaceae), a rare endemic of the Juan Fernández archipelago, Chile. Plant Species Biol. 16: 185192.Google Scholar
Crawford, D. J., Whitkus, R., and Stuessy, T. F. 1987b. Plant evolution and speciation on oceanic islands. In Urbanska, K. (ed.), Differentiation Patterns in Higher Plants (pp. 184199). London: Academic Press.Google Scholar
Crisp, M. D., Gilmore, S., and Van Wyk, B.-E. 2000. Molecular phylogeny of the genistoid tribes of papilionoid legumes. In Herendeen, P. S. and Bruneau, A. (eds.), Advances in Legume Systematics, part 9 (pp. 249276). Kew: Royal Botanic Gardens.Google Scholar
Cronk, Q. C. B. 1997. Islands: stability, diversity, conservation. Biodiv. Conserv. 6: 477493.Google Scholar
Cronk, Q. C. B., and Fuller, J. K. 2001. Plant Invaders: The Threat to Natural Ecosystems. London: Earthscan Publications.Google Scholar
Cronquist, A. 1943. The separation of Erigeron from Conyza. Bull. Torrey Bot. Club 70: 629632.Google Scholar
Cronquist, A. 1947. Revision of the North American species of Erigeron, north of Mexico. Brittonia 6: 121302.Google Scholar
Cronquist, A. 1981. An Integrated System of Classification of Flowering Plants. New York: Columbia University Press.Google Scholar
Crutzen, P. J., and Stoermer, E. F. 2000. The “Anthropocene.” Global Change Newsl. 41: 1718.Google Scholar
Cuddihy, L. W., and Stone, C. P. 1990. Alteration of Native Hawaiian Vegetation: Effects of Humans, Their Activities and Introductions. Honolulu: University of Hawaii Cooperative National Park Resources Studies Unit.Google Scholar
Cuevas, J. G. 2002. Informe técnico no. 29. Distribución y abundancia de la flora vascular del Archipiélago de Juan Fernández. Il Parte Final. Unpublished report, Corporación Nacional Forestal V Región, Viña del Mar.Google Scholar
Cuevas, J. G., and Le Quesne, C. 2006. Low vegetation recovery after short-term cattle exclusion on Robinson Crusoe Island. Plant Ecol. 183: 105124.Google Scholar
Cuevas, J. G., Marticorena, A., and Cavieres, L. A. 2004. New additions to the introduced flora of the Juan Fernández islands: origin, distribution, life history traits, and potential of invasion. Rev. Chil. Hist. Nat. 77: 523538.Google Scholar
Cuevas, J. G., and van Leersum, G. 2001. Project “Conservation, restoration, and development of The Juan Fernández islands, Chile.” Rev. Chil. Hist. Nat. 74: 899910.Google Scholar
D’Arcy, W. G. 1976. New names and taxa in the Solanaceae. Ann. Missouri Bot. Gard. 63: 363369.Google Scholar
Daly, H., and Magnacca, K. 2003. Hawaiian Hylaeus (Nesoprosopis) bees (Hymenoptera: Apoidea). In Insects of Hawaii, vol. 17. Honolulu: University of Hawaii Press.Google Scholar
Dampier, W. 1697. A New Voyage Round the World. London.Google Scholar
Dandy, J. E. (ed.). 1958. The Sloane Herbarium. An annotated list of the Horti Sicci composing it; with biographical accounts of the principal contributors. Based on records compiled by the late James Britten. British Museum: London.Google Scholar
Danton, P. 1998. Redécouverte de l’orchidée des îles de Robinson (Chili): Gavilea insularis M. N. Correa. Bull. Mens. Soc. Linn. Lyon 67: 226232.Google Scholar
Danton, P. 2000. Une ascension au Mont Yunque dans l’île de Robinson Crusoe (Chili). Bull. Mens. Soc. Linn. Lyon 69: 205216.Google Scholar
Danton, P. 2004. Plantas Silvestres de la Isla Robinson Crusoe. Viña del Mar: CONAF.Google Scholar
Danton, P. 2006a. Contribution à la flore de l’archipel Juan Fernández (Chili): Description de deux taxons nouveaux: Nicotiana cordifolia subsp. sanctaclarae subsp. nov. (Solanaceae), Robinsonia saxatilis sp. nov. (Asteraceae). Acta Bot. Gallica 153: 249255.Google Scholar
Danton, P. 2006b. La “myrtisylve” de l´archipel Juan Fernandez (Chili), une forêt en voie de disparation rapide. Acta Bot. Gallica 153: 179199.Google Scholar
Danton, P. 2014. Contribution à la flore de l’archipel Juan Fernández (Chili): description de 4 taxons nouveaux: Angiospermae – Erigeron corrales-molinensis sp. nov. (Asteraceae), Euphrasia formossissima Skottsb. subsp. cucharensis subsp. nov. (Orobanchaceae), Haloragis masatierrana var. applanata var. nov. et var. scabrida var. nov. (Haloragidaceae). Acta Bot. Gallica 161: 355371.Google Scholar
Danton, P., Boudrie, M., Bizot, A., and Viane, R. L. L. 2015. Pleopeltis ×cerro-altoensis (Polypodiaceae), a new fern hybrid from Robinson Crusoe Island (Juan Fernández Archipelago, Chile). Fern Gaz. 20: 6578.Google Scholar
Danton, P., Breteau, E., and Baffray, M. 1999. Les Îles de Robinson: Trésor Vivant des Mers du Sud. Entre Légende et Realité. Paris: Nathan.Google Scholar
Danton, P., Buord, S., and Stutzin, M. 2013. Las islas de Robinson Crusoe: riqueza biológica amenazada y propuesta de creación de los Jardin y Conservatorio Botánicos. Chagual 11: 1229.Google Scholar
Danton, P., and Perrier, C. 2005. Notes sur la disparition d’une espèce emblématique: Robinsonia berteroi (DC.) Sanders, Stuessy & Martic. (Asteraceae), dans l’île Robinson Crusoe, archipel Juan Fernández (Chili). J. Bot. Soc. Bot. France 31: 38.Google Scholar
Danton, P., and Perrier, C. 2017. Monografía de las Plantas Vasculares del Archipiélago Juan Fernández, Chile. Ensayo de Valoración para su Preservación. Gignac: Museo Editions.Google Scholar
Danton, P., Perrier, C., and Martínez, R. G. 2006. Nouveau catalogue de la flore vasculaire de l’archipel Juan Fernández (Chili) [Nuevo catálogo de la flora vascular del Archipiélago Juan Fernández (Chile)]. Acta Bot. Gallica 153: 399587.Google Scholar
Darlington, P. J. 1957. Zoogeography. New York: Wiley.Google Scholar
Darwin, C. 1855. Letter to J. D. Hooker, March 7, 1855. In Darwin, F. (ed.), Life and Letters of Charles Darwin. New York: Basic Books.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. London: John Murray.Google Scholar
Davis, M. A., Chew, M. K., Hobbs, R. J., Lugo, A. E., Ewel, J. J., et al. 2011. Don’t judge species on their origins. Nature 474: 153154.Google Scholar
de Joode, D. R., and Wendel, J. F. 1992. Genetic diversity and origin of the Hawaiian Islands cotton, Gossypium tomentosum. Am. J. Bot. 79: 13111319.Google Scholar
de la Sota, E. R. 1972. Notas sobre especies austrosudamericanas del género Blechnum L. (Blechnaceae-Pteridophyta). V. Bol. Soc. Argent. Bot. 14: 190197.Google Scholar
de la Sota, E. R., Salino, A., and Costa Asis, F. 2007. Nuevas combinaciones en Polypodiaceae. Darwiniana 45: 239.Google Scholar
de Queiroz, A. 2005. The resurrection of oceanic dispersal in historical biogeographyTrends Ecol. Evol20: 6873.Google Scholar
de Queiroz, K. 1998. The general lineage concept of species, species criteria, and the process of speciation. In Howard, D. J. and Berlocher, S. H. (eds.), Endless Forms: Species and Speciation (pp. 5775). New York: Oxford University Press.Google Scholar
de Queiroz, K. 2007. Species concepts and species delimitation. Syst. Biol. 56: 879886.Google Scholar
Decaisne, J. 1834. Monographie des genres Balbisia et Robinsonia, de la famille des Composées. Ann. Sc. Natur. Bot. Ser. 2 1: 115.Google Scholar
Defoe, D. 1719. The Life and Strange Surprising Adventures of Robinson Crusoe. London.Google Scholar
Delano, A. 1817. A Narrative of Voyages and Travels in the Northern and Southern Hemispheres: Comprising Three Voyages Round the World; Together with a Voyage of Survey and Discovery, in the Pacific and Oriental Islands. Boston.Google Scholar
Delprete, P. G., Forneris, G., and Pistarino, A. 2002. Carlo Bertero (1789–1831) in the New World. Sida 20: 621644.Google Scholar
Dempster, L. T. 1980. The genus Galium section Lophogalium (Rubiaceae) in South America. Allertonia 2: 247279.Google Scholar
Dempster, L. T. 1981. The genus Galium (Rubiaceae) in South America. II. Allertonia 2: 393426.Google Scholar
Devey, C. W., Hémond, C., and Stoffers, P. 2000. Metasomatic reactions between carbonate plume melts and mantle harzburgite: the evidence from Friday and Domingo seamounts (Juan Fernandez chain, SE Pacific). Contrib. Mineral. Petrol. 139: 6884.Google Scholar
Diaz, H. F., and Markgraf, V. (eds.). 1992. El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation. Cambridge: Cambridge University Press.Google Scholar
Diem, J., and de Lichtenstein, J. S. 1959. Las Himenofiláceas del área argentino-chilena del sud. Darwiniana 11: 611760.Google Scholar
Dirnböck, T., Greimler, J., López S., P., and Stuessy, T. F. 2003. Predicting future threats to the native vegetation of Robinson Crusoe Island, Juan Fernández Archipelago, Chile. Conserv. Biol. 17: 16501659.Google Scholar
Dobzhansky, T. 1937. Genetics and the Origin of Species. New York: Columbia University Press.Google Scholar
Dohzono, I., and Yokoyama, J. 2010. Impacts of alien bees on native plant-pollinator relationships: a review with special emphasis on plant reproduction. Appl. Entomol. Zool. 45: 3747.Google Scholar
Don, D. 1832. Descriptive catalogue of the Compositae contained in the herbarium of Dr. Gillies; with some additions from other sources. Philos. Mag. Ann. Chem. 11: 387392.Google Scholar
Donoso Zegers, C. 1993. Bosques Templados de Chile y Argentina. Santiago: Editorial Universitaria.Google Scholar
Douglas, D. 1914. Journal Kept by David Douglas during His Travels in North America 1823–1817. London: Royal Horticultural Society (reprint 1959. New York: Antiquarian Press).Google Scholar
Drábková, L., Kirschner, J., Seberg, O., Petersen, G., and Vlček, Č. 2003. Phylogeny of the Juncaceae based on rbcL sequences, with special emphasis on Luzula DC. and Juncus L. Plant Syst. Evol. 240: 133147.Google Scholar
Drábková, L., Kirschner, J., and Vlček, Č. 2006. Phylogenetic relationships within Luzula DC. and Juncus L. (Juncaceae): a comparison of phylogenetic signals of trnL-trnF intergenic spacer, trnL intron and rbcL plastome sequence data. Cladistics 22: 132143.Google Scholar
Drábková, L., Kirschner, J., Vlček, Č., and Pačes, V. 2004. TrnL-trnF intergenic spacer and trnL intron define major clades within Luzula and Juncus (Juncaceae): importance of structure mutations. J. Mol. Evol. 59: 110.Google Scholar
Driscoll, H. E., and Barrington, D. S. 2007. Origin of Hawaiian Polystichum (Dryopteridaceae) in the context of a world phylogeny. Am. J. Bot. 94: 14131424.Google Scholar
Dubuisson, J.-Y. 1997a. Systematic relationships within the genus Trichomanes sensu lato (Hymenophyllaceae, Filicopsida): cladistic analysis based on anatomical and morphological data. Bot. J. Linn. Soc. 123: 265296.Google Scholar
Dubuisson, J.-Y. 1997b. rbcL sequences: a promising tool for the molecular systematics of the fern genus Trichomanes (Hymenophyllaceae)? Mol. Phylogenet. Evol. 8: 128138.Google Scholar
Dubuisson, J.-Y., Hennequin, E. J. P., Douzery, R. B., Cranfill, A. R., Smith, A. R., et al. 2003. rbcL phylogeny of the fern genus Trichomanes (Hymenophyllaceae) with special reference to Neotropical taxa. Int. J. Plant Sci. 164: 753761.Google Scholar
Duek, J. J., and Rodríguez, R. 1972. Lista preliminar de las especies de Pteridophyta en Chile continental e insular. Bol. Soc. Bol. Concepción 45: 129174.Google Scholar
Dunbar-Co, S., Wieczorek, A. M., and Morden, C. W. 2008. Molecular phylogeny and adaptive radiation of the endemic Hawaiian Plantago species (Plantaginaceae). Am. J. Bot. 95: 11771188.Google Scholar
Dupont, Y. L., and Skov, C. 2004. Influence of geographical distribution and floral traits on species richness of bees (Hymenoptera: Apoidea) visiting Echium species (Boraginaceae) of the Canary Islands. Int. J. Plant Sci. 165: 377386.Google Scholar
Ebihara, A., Dubuisson, J.-Y., Iwatsuki, K., Hennequin, S., and Ito, M. 2006. A taxonomic revision of Hymenophyllaceae. Blumea 51: 221280.Google Scholar
Ebihara, A., and Iwatsuki, K. 2007. The Hymenophyllaceae of the Pacific Area: 1. Hymenophyllum subgenus Hymenophyllum. Bull. Natl. Mus. Nat. Sci. Ser. B 33: 5568.Google Scholar
Ebihara, A., Iwatsuki, K., Ito, M., Hennequin, S., and Dubuisson, J.-Y. 2007. A global molecular phylogeny of the fern genus Trichomanes (Hymenophyllaceae) with special reference to stem anatomy. Bot. J. Linn. Soc. 155: 127.Google Scholar
Ebihara, A., Nitta, J. H., and Iwatsuki, K. 2010. The Hymenophyllaceae of the Pacific Area: 2. Hymenophyllum (excluding subgen. Hymenophyllum). Bull. Natl. Mus. Nat. Sci. Ser. B 36: 4359.Google Scholar
Edmands, S. 2007. Between a rock and a hard place: evaluating the relative risks of inbreeding and outbreeding for conservation and management. Mol. Ecol. 16: 463475.Google Scholar
Edwards, C. E. 2015. Looking to the future of conservation genetics: the case for using quantitative genetic experiments to estimate the ability of rare plants to withstand climate change. Am. J. Bot. 102: 10111013.Google Scholar
Ehrendorfer, F. 1979. Reproductive biology in island plants. In Bramwell, D. (ed.), Plants and Islands (pp. 293306). London: Academic Press.Google Scholar
Eizaguirre, C., and Baltazar-Soares, M. 2014. Evolutionary conservation: evaluating the adaptive potential of species. Evol. Appl. 7: 963967.Google Scholar
Ekblom, R., and Wolf, J. B. W. 2014. A field guide to whole-genome sequencing, assembly and annotation. Evol. Appl. 7: 10261042.Google Scholar
Eliasson, U. 1995. Patterns of diversity in island plants. In Vitousek, P. M., Loope, L. L., and Adsersen, H. (ed.), Islands: Biological Diversity and Ecosystem Functions (pp. 3550). New York: Springer-Verlag.Google Scholar
Ellstrand, N., and Elam, D. 1993. Population genetic consequences of small population size: Implications for plant conservation. Ann. Rev. Ecol. Syst. 24: 217242.Google Scholar
Emerson, B. C. 2002. Evolution on oceanic islands: molecular phylogenetic approaches to understanding pattern and processMol. Ecol11: 951966.Google Scholar
Emerson, B. C., and Kolm, N. 2005a. Species diversity can drive speciation. Nature 434: 10151017.Google Scholar
Emerson, B. C., and Kolm, N. 2005b. Emerson and Kolm reply. Nature 438: E2.Google Scholar
Engel, M. S. 2000. A new Lasioglossum from the Juan Fernández Islands (Hymenoptera: Halictidae). Rev. Chilena Entomol. 27: 510.Google Scholar
Engler, A. 1882. Versuch einer Entwicklunsgeschichte der Pflanzenwelt, insbesondere der Florengebiete, seit der Tertiärperiode. Leipzig: W. Engelman.Google Scholar
Ernst, B., Chamorro, J., Manríquez, P., Orensanz, J. M. L., Parma, A. M., et al. 2013. Sustainability of the Juan Fernández lobster fishery (Chile) and the perils of generic science-based prescriptions. Global Environ. Change 23: 13811392.Google Scholar
Esquemeling, J. 1685. The Buccaneers of America, the Second Volume Containing the Dangerous Voyage and Bold Attempts of Captain Bartholomew Sharp and Others Performed upon the Coasts of the South Sea for the Space of Two Years from the Original Journal of the Said Voyage Written by Mr. Basil Ringrose, Gent., Who Was All Along Present at Those Transactions (2nd edn., English translation). London.Google Scholar
Esselman, E. J., Crawford, D. J., Brauner, S., Stuessy, T. F., Anderson, G. J., et al. 2000. RAPD marker diversity within and divergence among species of Dendroseris (Asteraceae: Lactuceae). Am. J. Bot. 87: 591596.Google Scholar
Evans, A. 1930. The thallose Hepaticae of the Juan Fernandez Islands. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 5515860). Uppsala: Almqvist & Wiksells.Google Scholar
Evans, T. M., Jabaily, R. S., Faria, A. P. G., Sousa, L. O. F., Wendt, T., et al. 2015. Phylogenetic relationships in Bromeliaceae subfamily Bromelioideae based on chloroplast DNA sequence data. Syst. Bot. 40: 116128.Google Scholar
Faith, D. P. 1994. Phylogenetic diversity: a general framework for the prediction of feature diversity. In Forey, P. L., Humphries, C. J., and Vane-Wright, R. I. (eds.), Systematics and Conservation Evaluation (pp. 251268). Oxford: Clarendon Press.Google Scholar
Fanning, E. 1833. Voyages Round the World; With Selected Sketches of Voyages to the South Seas, North and South Pacific Oceans, China, etc. New York: Collins & Hannay (new edition 1924, under the title of Voyages & Discoveries in the South Seas, 1792–1832, edited and published by the Marine Research Society, Salem, MA).Google Scholar
Farías, M., Vargas, G., Tassara, A., Carretier, S., Baize, S., et al. 2010. Land-level changes produced by the Mw 8.8 2010 Chilean earthquake. Science 329: 916.Google Scholar
Farley, K. A., Basu, A. R., and Craig, H. 1993. He, Sr and Nd isotopic variations in lavas from the Juan Fernandez Archipelago, SE Pacific. Contrib. Mineral. Petrol. 115: 7587.Google Scholar
Felsenstein, J. 2004. Inferring Phylogenies. Sunderland, MA: Sinauer Associates.Google Scholar
Fernández-Palacios, J. M., Kueffer, C., and Drake, D. 2015. A new golden era in island biogeography. Front. Biogeog. 7: 1420.Google Scholar
Ferrara, G., Clarke, W. B., Rama Murthy, V., and Bass, M. N. 1969. K-Ar ages of Juan Fernandez Islands and southeast Pacific dredge hauls. Eos Transact. Am. Geophys. Union 50 (4): 329 (Abstr.).Google Scholar
Ferreyra, M. L., Rius, S. P., and Casati, P. 2012. Flavonoids: biosynthesis, biological functions, and biotechnological applications. Front. Plant Sci. 3(222): 115.Google Scholar
Fior, S., Karis, P. O., Casazza, G., Minuto, L., and Sala, F. 2006. Molecular phylogeny of the Caryophyllaceae (Caryophyllales) inferred from chloroplast matK and nuclear rDNA ITS sequences. Am. J. Bot. 93: 399411.Google Scholar
Fisher, A. E., Clark, L. G., and Kelcher, S. A. 2014. Molecular phylogeny estimation of the bamboo genus Chusquea (Poaceae: Bambusoideae: Bambuseae) and description of two new subgenera. Syst. Bot. 39: 829844.Google Scholar
Fishman, L., Stathos, A., Beardsley, P. M., Williams, C. F., and Hill, J. P. 2013. Chromosomal rearrangements and the genetics of reproductive barriers in Mimulus (monkey flowers). Evolution 67: 25472560.Google Scholar
Fitzpatrick, B. M., Fordyce, J. A., and Gavrilets, S. 2008. What, if anything, is sympatric speciation? J. Evol. Biol. 21: 14521459.Google Scholar
Fitzpatrick, B. M., Fordyce, J. A., and Gavrilets, S. 2009. Pattern, process and geographic modes of speciation. J. Evol. Biol. 22: 23422347.Google Scholar
Fjellheim, S., Jogensen, M. H., Kjos, M., and Borgen, L. 2009. A molecular study of hybridization and homoploid hybrid speciation in Argyranthemum (Asteraceae) on Tenerife, Canary Islands. Bot. J. Linn. Soc. 159: 1931.Google Scholar
Fleischer, R. C., and McIntosh, C. E. 2001. Molecular systematics and biogeography of the Hawaiian avifauna. Stud. Avian Biol. 22: 5160.Google Scholar
Forde, M. G. 1964. Haloragis erecta: a species complex in evolution. New Zealand J. Bot. 2: 425453.Google Scholar
Fosberg, F. R. 1967. Some ecological effects of wild and semi-wild exotic species of vascular plants. In IUCN (ed.), Towards a New Relationship of Man and Nature in Temperate Lands, Part III: Changes due to Introduced Species (pp. 98109). Morges: International Union for Conservation of Nature and Natural Resources.Google Scholar
Fosberg, F. R. 1968. Studies in Pacific Rubiaceae: VI–VIII. Brittonia 20: 287294.Google Scholar
Frakes, L. A. 1979. Climates through Geologic Time. Amsterdam: Elsevier.Google Scholar
Francisco-Ortega, J, Crawford, D. J., Santos-Guerra, A., and Carvalho, J. A. 1996. Isozyme variation in the endemic genus Argyranthemum (Asteraceae: Anthemideae) in the Macaronesian Islands. Plant Syst. Evol. 202: 137152.Google Scholar
Francisco-Ortega, J., Crawford, D. J., Santos-Guerra, A., and Jansen, R. K. 1997. Origin and evolution of Argyranthemum (Asteraceae: Anthemideae) in Macaronesia. In Givnish, T. J. and Sytsma, K. J. (eds.), Molecular Evolution and Adaptive Radiation (pp. 407431). Cambridge: Cambridge University Press.Google Scholar
Francisco-Ortega, J., Santos-Guerra, A., Kim, S.-C., and Crawford, D. J. 2000. Plant genetic diversity in the Canary Islands: a conservation perspective. Am. J. Bot. 87: 909919.Google Scholar
Frankham, R. 1997. Do island populations have less genetic variation than mainland populations? Heredity 78: 311327.Google Scholar
Frankham, R. 2010. Where are we in conservation genetics and where do we need to go? Conserv. Genet. 11: 661663.Google Scholar
Frankham, R. 2015. Genetic rescue of small inbred populations: metaanalysis reveals large and consistent benefits of gene flow. Mol. Ecol. 24: 26102618.Google Scholar
Frankham, R., Ballou, J. D., and Briscoe, D. A. 2010. Introduction to Conservation Genetics (2nd edn.). Cambridge: Cambridge University Press.Google Scholar
Franzke, A., Pollmann, K., Bleeker, W., Kohrt, R., and Hurka, H. 1998. Molecular systematics of Cardamine and allied genera (Brassicaceae): ITS and non-coding chloroplast DNA. Folia Geobot. 33: 225240.Google Scholar
Freire, S. E., Salomón, L., Bayón, N. D., Baeza, C. M., Muñoz-Schick, M., et al. 2016. Taxonomic revision of the genus Gamochaeta (Gnaphalieae, Asteraceae) in Chile. Gayana Bot. 73: 252305.Google Scholar
Frenzke, L., Scheiris, E., Pino, G., Symmank, L., Goetghebeur, P., et al. 2015. A revised infrageneric classification of the genus Peperomia (Piperaceae). Taxon 64: 424444.Google Scholar
Fries, R. E. 1920. Die Myxomyceten der Juan Fernandez-Inseln. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 5558). Uppsala: Almqvist & Wiksells.Google Scholar
Fries, T. C. E. 1921. Die Gasteromyceten der Juan Fernandez- und Osterinseln. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 5960). Uppsala: Almqvist & Wiksells.Google Scholar
Fritsch, P. W., Lu, L., Bush, C. M., Cruz, B. C., Kron, K. A., et al. 2011. Phylogenetic analysis of the wintergreen group (Ericaceae) based on six genic regions. Syst. Bot. 36: 9901003.Google Scholar
Fritts, T. H., and Rodda, G. H. 1998. The role of introduced species in the degradation of island ecosystems. Annu. Rev. Ecol. Syst. 29: 113140.Google Scholar
Fuentes-Bazan, S., Mansion, G., and Borsch, T. 2012a. Towards a species level tree of the globally diverse genus Chenopodium (Chenopodiaceae). Mol. Phylogenet. Evol. 62: 359374.Google Scholar
Fuentes-Bazan, S., Uotila, P., and Borsch, T. 2012b. A novel phylogeny-based generic classification for Chenopodium sensu lato, and a tribal rearrangement of Chenopodioideae (Chenopodiaceae). Willdenowia 42: 524.Google Scholar
Fuenzalida P., H. 1966. Climatología. In Corporación de Fomento de la Producción (CORFO) (ed.), Geografía Económica de Chile (2nd edn., pp. 3144). Santiago de Chile: Editorial Universitaria.Google Scholar
Fuenzalida V., H. 1944. Don Claudio Gay y el Museo Nacional de Historia Natural. Bol. Mus. Nac. Hist. Nat. 22: 516.Google Scholar
Funk, J. L., and Vitousek, P. M. 2007. Resource-use efficiency and plant invasion in low-resource systems. Nature 446: 10791081.Google Scholar
Funk, V. A., and Wagner, W. L. 1995. Biogeographic patterns in the Hawaiian Islands. In Wagner, W. L. and Funk, V. A. (eds.), Hawaiian Biogeography: Evolution on a Hot Spot Archipelago (pp. 379419). Washington, DC: Smithsonian Institution Press.Google Scholar
Funnell, W. 1707. A Voyage Round the World; Containing an Account of Captain Dampier’s expedition into the South-Seas in the Ship St. George, in the Years 1703 and 1704. London.Google Scholar
Fürstenberg, P. 1906. Dr. Rudolph Amandus Philippi. Sein Leben und seine Werke. Verhandl. Deutsch. Wissenschafl. Vereins (Santiago) 5: 139.Google Scholar
Gabel, J. D., Les, D. H., Charlton, W. A., and Coombes, A. J. 2000. Phylogeny of Azara (Flacourtiaceae). Am. J. Bot. 87(6): 128 (Abstr.).Google Scholar
Gabriel y Galán, J. M., Prada, C., Rolleri, C., Ainouche, A., and Vicent, M. 2013. cpDNA supports the identification of the major lineages of American Blechnum (Blechnaceae, Polypodiopsida) established by morphology. Turk. J. Bot. 37: 769777.Google Scholar
Galapagos Conservancy. 2015. Ecosystem Restoration: Project Isabela. Available at www.galapagos.org/conservation/project-isabela/ (accessed March 11, 2016).Google Scholar
Gallagher, R. E. 1964. Byron’s Journal of His Circumnavigation 1764–1766. Cambridge: Cambridge University Press (Hakluyt Society).Google Scholar
Gamerro, J. C., and Barreda, V. 2008. New fossil record of Lactoridaceae in southern South America: a palaeobiogeographical approach. Bot. J. Linn. Soc. 158: 4150.Google Scholar
Ganders, F. R., Berbee, M., and Pirseyedi, M. 2000. ITS base sequence phylogeny in Bidens (Asteraceae): evidence for the continental relatives of Hawaiian and Marquesan Bidens. Syst. Bot. 25: 122133.Google Scholar
Ganders, F. R., Bohm, B. A., and McCormick, S. P. 1990. Flavonoid variation in Hawaiian Bidens. Syst. Bot. 15: 231239.Google Scholar
Ganders, F. R., and Nagata, K. M. 1984. The role of hybridization in the evolution of Bidens on the Hawaiian Islands. In Grant, W. F. (ed.), Plant Biosystematics (pp. 179194). Toronto: Academic Press.Google Scholar
Garcia, J. D. D. 2000. Patterns of insect flower visitation in Lavandula buchii Webb (Lamiaceae), an endemic shrub of Tenerife (Canary Islands). J. Nat. Hist. (Lond.) 34: 21452155.Google Scholar
Garcia-Jacas, N., Susanna, A., Garnatje, T., and Vilatersana, R. 2001. Generic delimitation and phylogeny of the subtribe Centaureinae (Asteraceae): a combined nuclear and chloroplast DNA analysis. Ann. Bot. 87: 503515.Google Scholar
García-Verdugo, C., Friar, E., and Santiago, L. S. 2013. Ecological role of hybridization in adaptive radiations: a case study in the Dubautia arborea-Dubautia ciliolata (Asteraceae) complex. Int. J. Plant Sci. 174: 749759.Google Scholar
García-Verdugo, C., Sajeva, M., La Mantia, T., Harrouni, C., Msanda, F., et al. 2015. Do island plant populations really have lower genetic variation than mainland populations? Effects of selection and distribution range on genetic diversity estimates. Mol. Ecol. 24: 726741.Google Scholar
Garnock-Jones, P. J. 2014. Evidence-based review of the taxonomic status of New Zealand’s endemic seed plant genera. New Zealand J. Bot. 52: 163212.Google Scholar
Gastony, G. J. 1981. Spore morphology in the Dicksoniaceae: I. The genera Cystodium, Thyrsopteris, and Culcita. Am. J. Bot. 68: 808819.Google Scholar
Gavrilets, S. 2014. Models of speciation: where are we now? J. Hered. 105: 743755.Google Scholar
Gay, C. 1854. Atlas de la Historia Física y Política de Chile. Paris: E. Thunot.Google Scholar
Gay, C. 1845–1854. Historia Física y Política de Chile según Documentos Adquiridos en esta República durante Doce Años de Residencia en ella y Publicada bajo los Auspicios del Supremo Gobierno. Botánica. Paris.Google Scholar
Geltman, D. V. 1998. New species and new combinations of Urtica (Urticaceae) in South America. Novon 8: 1517.Google Scholar
Gerlach, D.C., Hart, S. R., Morales, V. W. J., and Palacios, C. 1986. Mantle heterogeneity beneath the Nazca plate: San Felix and Juan Fernandez islands. Nature 322: 165169.Google Scholar
Giannasi, D. E., and Crawford, D. J. 1986. Biochemical systematics II. A reprise. Evol. Biol. 20: 25248.Google Scholar
Gillespie, R. G. 2004. Community assembly through adaptive radiation in Hawaiian spiders. Science 303: 356359.Google Scholar
Gillespie, R. G., and Clague, D. A. (eds.). 2009. Encyclopedia of Islands. Berkeley: University of California Press.Google Scholar
Gillespie, T., Lipkin, B., Sullivan, L., Benowitz, D. R., Pau, S., et al. 2012. The rarest and least protected forests in biodiversity hotspots. Biodivers. Conserv. 21: 35973611.Google Scholar
Gillett, G. W., and Lim, E. K. S. 1970. An experimental study of the genus Bidens in the Hawaiian Islands. Univ. Calif. Publ. Bot. 56: 163.Google Scholar
Gitaí, J., Horres, R., and Benko-Iseppon, A. M. 2005. Chromosome features and evolution of Bromeliaceae. Plant Syst. Evol. 253: 6580.Google Scholar
Gitaí, J., Paule, J., Zizka, G., Schulte, K., and Benko-Iseppon, A. M. 2014. Chromosome numbers and DNA content in Bromeliaceae: additional data and critical review. Bot. J. Linn. Soc. 176: 349368.Google Scholar
Givnish, T. J. 2010. Ecology of plant speciation. Taxon 13261366.Google Scholar
Givnish, T. J. 2015. Adaptive radiation versus ‘radiation’ and ‘explosive diversification’: why conceptual distinctions are fundamental to understanding evolution. New Phytol. 207: 297303.Google Scholar
Givnish, T. J., Barfuss, M. H. J., Van Ee, B., Riina, R., Schulte, K., et al. 2011. Phylogeny, adaptive radiation, and historical biogeography in Bromeliaceae: insights from an eight-locus plastid phylogeny. Am. J. Bot. 98: 872895.Google Scholar
Givnish, T. J., and Montgomery, R. A. 2014. Common-garden studies on adaptive radiation of photosynthetic physiology among Hawaiian lobeliads. Proc. R. Soc. Lond. B 281: 20132944. Available at http://dx.doi.org/10.1098/rspb.2013.2944.Google Scholar
Glenn, T. G., and Schnable, N. A. 2005. Isolating microsatellite DNA loci. Methods Enzymol. 395: 202222.Google Scholar
Global Carex Group. 2015. Making Carex monophyletic (Cyperaceae, tribe Cariceae): a new broader circumscription. Bot. J. Linn. Soc. 179: 142.Google Scholar
Godoy, R. 1989. Espectro biológico de la flora pteridofítica de Chile continental e insular. Anales Jard. Bot. Madrid 46: 599603.Google Scholar
Goldsworthy, S., Francis, J., Boness, D., and Fleischer, R. 2000. Variation in the mitochondrial control region in the Juan Fernández fur seal (Arctocephalus philippii). J. Hered. 91: 371377.Google Scholar
Gonzales R., J., and Kessler, M. 2011. A synopsis of the Neotropical species of Sticherus (Gleicheniaceae), with descriptions of nine new species. Phytotaxa 31: 154.Google Scholar
González, A., and Fuertes, J. 2011. Ecología y evolución de las plantas ornitófilas de la Macaronesia. El Indiferente 21: 6475.Google Scholar
González, F., and Rudall, P. 2001. The questionable affinities of Lactoris: evidence from branching pattern, inflorescence morphology, and stipule development. Am. J. Bot. 88: 21432150.Google Scholar
González, F., Rudall, P. J., and Furness, C. A. 2001. Microsporogenesis and systematics of Aristolochiaceae. Bot. J. Linn. Soc. 137: 221242.Google Scholar
Good, R. 1974. The Geography of Flowering Plants. London: Longman.Google Scholar
Goodspeed, T. H. 1954. The Genus Nicotiana. Origins, Relationships and Evolution of Its Species in the Light of Their Distribution, Morphology and Cytogenetics. Waltham, MA: Chronica Botanica.Google Scholar
Gornall, R. J., and Bohm, B. A. 1979. The distribution of flavonoids in the angiosperms. Bot. Not. 132: 130.Google Scholar
Gotschlich, B. 1904. Biografía del Dr. Rodulfo Amando Philippi (1803–1904). Valdivia: Imprenta Central J. Lampert.Google Scholar
Gottlieb, L. D. 1971. Gel electrophoresis: new approach to the study of evolution. Bioscience 21: 939944.Google Scholar
Gottlieb, L. D. 1973. Genetic differentiation, sympatric speciation, and the origin of a diploid species of Stephanomeria. Am. J. Bot. 60: 545553.Google Scholar
Gottlieb, L. D. 1977. Electrophoretic evidence and plant systematics. Ann. Missouri Bot. Gard. 64: 161180.Google Scholar
Gottlieb, L. D. 1981a. Electrophoretic evidence and plant populations. Prog. Phytochem. 7: 146.Google Scholar
Gottlieb, L. D. 1981b. Gene number in species of Astereae that have different chromosome numbers. Proc. Natl. Acad. Sci. USA 78: 37263729.Google Scholar
Gottlieb, L. D. 1982. Conservation and duplication of isozymes in plants. Science 216: 373380.Google Scholar
Gottlieb, L. D. 2003. Rethinking classic examples of recent speciation in plants. New Phytol. 161: 7182.Google Scholar
Graham, M. 1824. Journal of a Residence in Chile during the Year 1822 and a Voyage from Chile to Brazil in 1823. London.Google Scholar
Grant, P. R. (ed.). 1998. Evolution on Islands. Oxford: Oxford University Press.Google Scholar
Grant, P. R., and Grant, B. R. 2008. How and Why Species Multiply. Princeton, NJ: Princeton University Press.Google Scholar
Grant, P. R., and Grant, B. R. 2010. Sympatric speciation, immigration, and hybridization in island birds. In Losos, J. B. and Ricklefs, R. E. (eds.), The Theory of Island Biogeography Revisited (pp. 326357). Princeton, NJ: Princeton University Press.Google Scholar
Grant, V. 1950. The protection of ovules in flowering plants. Evolution 3: 179201.Google Scholar
Grant, V. 1963. The Origin of Adaptations. New York: Columbia University Press.Google Scholar
Grant, V. 1981. Plant Speciation (2nd edn.). New York: Columbia University Press.Google Scholar
Green, P. T. 1999. Greta’s Garbo: stranded seeds and fruits from Greta Beach, Christmas Island, Indian Ocean. J. Biogeogr. 26: 937946.Google Scholar
Greenberg, A. K., and Donoghue, M. J. 2011. Molecular systematics and character evolution in Caryophyllaceae. Taxon 60: 16371652.Google Scholar
Greimler, J., López, S. P., Stuessy, T. F., and Dirnböck, T. 2002a. The vegetation of Robinson Crusoe Island (Isla Masatierra), Juan Fernández Archipelago, Chile. Pacific Sci. 56: 263284 + 1 map.Google Scholar
Greimler, J., López-Sepúlveda, P., Reiter, K., Baeza, C., Peñailillo, P., et al. 2013. Vegetation of Alejandro Selkirk Island (Isla Masafuera), Juan Fernández Archipelago, Chile. Pacific Sci. 67: 267282.Google Scholar
Greimler, J., Stuessy, T. F., Swenson, U., Baeza, C., and Matthei, O. 2002b. Plant invasions on an oceanic archipelago. Biol. Invasions 4: 7385.Google Scholar
Greuter, W., Barrie, F. R., Burdet, H. M., Chaloner, W. G., Demoulin, V., et al. 1994. International Code of Botanical Nomenclature (Tokyo Code) Adopted by the Fifteenth International Botanical Congress, Yokohama, August–September 1993. Königstein: Koeltz Sci. Books (Regnum Veg. 131).Google Scholar
Grondona, E. 1964. Las especies Argentinas del género Acaena. Darwiniana 13: 209342.Google Scholar
Groppo, M., Pirani, J. R., Salatino, M. L. F., Blanco, S. R., and Kallunki, J. A. 2008. Phylogeny of Rutaceae based on two noncoding regions from cpDNA. Am. J. Bot. 95: 9851005.Google Scholar
Grueber, C. E., Wallis, G. P., and Jamieson, I. G. 2008. Heterozygosity-fitness correlations and their relevance to studies on inbreeding depression in threatened species. Mol. Ecol. 17: 39783984.Google Scholar
Gunckel, L., , H. 1971. Las primeras plantas herborizadas en Chile en 1690. An. Instit. Patagon. 2(1–2): 134141.Google Scholar
Gunckel Luër, H. 1968. Nombres vernaculares de plantas endémicas de Juan Fernández. Rev. Univ. (Univ. Católica de Chile) 53(31): 126.Google Scholar
Gunckel Luër, H. 1984. Helechos de Chile. Santiago: Universidad de Chile.Google Scholar
Gunther, B., and Mahalik, P. 1990. Juania australis in habitat. Principes 34: 7981.Google Scholar
Guppy, H. B. 1890. The dispersal of plants as illustrated by the flora of the Keeling or Cocos Islands. Trans. Proc. Philos. Inst. Victoria 24: 276306.Google Scholar
Guppy, H. B. 1917. Plants, Seeds, and Currents in the West Indies and Azores. London: Williams & Norgate.Google Scholar
Gussarova, G., Popp, M., Vitek, E., and Brochmann, C. 2008. Molecular phylogeny and biogeography of the bipolar Euphrasia (Orobanchaceae): recent radiations in an old genus. Mol. Phylogenet. Evol. 48: 444460.Google Scholar
Haberle, S. G. 2003. Late Quaternary vegetation dynamics and human impact on Alexander Selkirk Island, Chile. J. Biogeogr. 30: 239255.Google Scholar
Habicht, J. K. A. 1979. Paleoclimate, Paleomagnetism and Continental Drift. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Hahn, I. 1996. The birds of the Juan Fernández Islands (Chile) with Special consideration to the Masafuera rayadito (Aphrastura masafuerae). Master’s Thesis, Institute of Landscape Ecology, Westfälische Wilhelms-Universität, Münster, Germany.Google Scholar
Hahn, I. 1998. Untersuchungen zur Ökologie und zum Lebensraum der Landvogelgemeinschaften des Juan Fernandez-Archipels (Chile). Ph.D. dissertation, Westfälische Wilhelms-Universität, Münster, Germany.Google Scholar
Hahn, I., Römer, U., Vergara, P., and Walter, H. 2009. Biogeography, diversity, and conservation of the birds of the Juan Fernández Islands, Chile. Vertebrate Zool. 59: 103114.Google Scholar
Hahn, I. J., Vergara, P. M., Baumeister, J., Soto, G. E., and Römer, U. 2014. Impact of the 2010 tsunami on an endangered insular soil-plant system. Nat. Hazards 71: 987993.Google Scholar
Hajek, E., and Espinoza, G. A. 1987. Meteorología, climatología y bioclimatología de las islas oceánicas Chilenas. In Castilla, J. C. (ed.), Islas Oceánicas Chilenas: Conocimiento Científico y Necesidades de Investigaciones (pp. 5583). Santiago: Ediciones Universidad Católica de Chile.Google Scholar
Hale, R. L. 1923. The Log of a Forty-Niner: Journal of a Voyage from Newbury-Port to San Francisco on the Brig General Worth Commanded by Capt. Samuel Walton, ed. by Russ, C. H.. Boston, MA: B.J. Brimmer.Google Scholar
Hamrick, J. L. 1989. Isozymes and analyes of genetic structure of plant populations. In Soltis, D. E. and Soltis, P. S. (eds.), Isozymes in Plant Biology (pp. 87105). Portland, OR: Dioscorides.Google Scholar
Hamrick, J. L., and Godt, M. J. W. 1989. Allozyme diversity in plant species. In Brown, A. H. D., Clegg, M. T., Kahler, A. L., and Weir, B. S. (eds.), Plant Population Genetics, Breeding and Genetic Resources (pp. 4363). Sunderland, MA: Sinauer Associates.Google Scholar
Hamrick, J. L., and Godt, M. J. W. 1996. Conservation genetics of endemic plant species. In Avise, J. C. and Hamrick, J. L. (eds.), Conservation Genetics: Case Studies from Nature (pp. 281304). New York: Chapman & Hall.Google Scholar
Hamrick, J. L., and Godt, M. J. W. 1997. Effects of life history traits on genetic diversity in plant species. In Silvertown, J., Franco, M., and Harper, J. L. (eds.), Plant Life Histories: Ecology, Phylogeny and Evolution (pp. 102118). Cambridge: Cambridge University Press.Google Scholar
Hamrick, J. L., Godt, M. J. W., Murawski, D. A., and Loveless, M. D. 1991. Correlations between species traits and allozyme diversity: implications for conservation biology. In Falk, D. A. and Holsinger, K. E. (eds.), Genetics and Conservation of Rare Plants (pp. 7586). New York: Oxford University Press.Google Scholar
Hamrick, J. L., Linhart, Y. B., and Mitton, J. B. 1979. Relationships between life history characteristics and electrophoretically detectable genetic variation in plants. Annu. Rev. Ecol. Syst. 10: 173200.Google Scholar
Harbaugh, D., and Baldwin, B. G. 2007. Phylogeny and biogeography of the sandalwoods (Santalum, Santalaceae): repeated dispersals throughout the Pacific. Am. J. Bot. 94: 10281040.Google Scholar
Harborne, J. B. 1967. Comparative Biochemistry of the Flavonoids. London: Academic Press.Google Scholar
Harborne, J. B., and Turner, B. L. 1984. Plant Chemosystematics. London: Academic Press.Google Scholar
Harling, G. 1962. On some Compositae endemic to the Galápagos Islands. Acta Hort. Berg. 20: 63120.Google Scholar
Harley, R. 1986. Cuminia eriantha, Labiatae. Kew Mag. 3: 151156.Google Scholar
Harris, H. 1966. Enzyme polymorphisms in man. Philos. Trans. R. Soc. Lond. B. 164: 298310.Google Scholar
Harrisson, K. A., Pavlova, A., Telonis-Scott, M., and Sunnucks, P. 2014. Using genomics to characterize evolutionary potential for conservation of wild populations. Evol. Appl. 7: 10081025.Google Scholar
Harter, D. E. V., Irl, S. D. H., Seo, B., Steinbauer, M. J., Gillespie, R., et al. 2015. Impacts of global climate change on the floras of oceanic islands: projections, implications and current knowledge. Perspect. Plant Ecol. Evol. Syst. 17: 160183.Google Scholar
Hartley, T. G. 1966. A revision of the Malesian species of Zanthoxylum (Rutaceae). J. Arnold Arbor. 47: 171221.Google Scholar
Haufler, D. H., Soltis, D. E., and Soltis, P. S. 1995. Phylogeny of the Polypodium vulgare complex: insights from chloroplast DNA restriction site data. Syst. Bot. 20: 110119.Google Scholar
Hauk, W. D., Parks, C. R., and Chase, M. W. 2003. Phylogenetic studies of Ophioglossaceae: evidence from rbcL and trnL-F plastid DNA sequences and morphology. Mol. Phylogenet. Evol. 28: 131151.Google Scholar
Hausdorf, B. 2011. Progress toward a general species concept. Evolution 65: 923931.Google Scholar
Heads, M. J. 1996. Biogeography, taxonomy and evolution in the Pacific genus Coprosma (Rubiaceae). Candollea 51: 381405.Google Scholar
Heaney, L. R. 2007. Is a new paradigm emerging for ocean island biogeography? J. Biogeogr. 34: 753757.Google Scholar
Heenan, P. B. 2001. The correct name for Chilean pelu (Fabaceae): the identity of Edwardsia macnabiana and the reinstatement of Sophora cassioides. New Zealand J. Bot. 39: 167170.Google Scholar
Heenan, P. B., Dawson, M. I., and Wagstaff, S. J. 2004. The relationship of Sophora sect. Edwardsia (Fabaceae) to Sophora tomentosa, the type species of the genus Sophora, observed from DNA sequence data and morphological characters. Bot. J. Linn. Soc. 146: 439446.Google Scholar
Helenurm, K., and Ganders, F. R. 1985. Adaptive radiation and genetic differentiation in Hawaiian Bidens. Evolution 39: 753765.Google Scholar
Hemsley, W. B. 1884. Report on the botany of Juan Fernandez, the south-eastern Moluccas, and the Admiralty Islands. In Thomson, C. W. and Murray, J. (superintendents), Report on the Scientific Results of the Voyage of H.M.S. Challenger during the Years 1873–7, vol. 1: Botany (pp. 196). London: Her Majesty’s Stationary Office.Google Scholar
Hendry, A. P., Nosil, P., and Rieseberg, L. H. 2007. The speed of ecological speciation. Funct. Ecol. 21: 455464.Google Scholar
Hennequin, S. 2003. Phylogenetic relationships within the fern genus Hymenophyllum s.l. (Hymenophyllaceae, Filicopsida): contribution of morphology and cytology. C. R. Biol. 326: 599611.Google Scholar
Hennequin, S., Ebihara, A., Dubuisson, J.-Y., and Schneider, H. 2010. Chromosome number evolution in Hymenophyllum (Hymenophyllaceae), with special reference to the subgenus Hymenophyllum. Mol. Phylogenet. Evol. 55: 4759.Google Scholar
Hennequin, S., Ebihara, A., Ito, M., Iwatsuki, K., and Dubuisson, J.-Y. 2003. Molecular systematics of the fern genus Hymenophyllum s.l. (Hymenophyllaceae) based on chloroplastic coding and non-coding regions. Mol. Phylogenet. Evol. 27: 283301.Google Scholar
Hennequin, S., Ebihara, A., Ito, M., Iwatsuki, K., and Dubuisson, J.-Y. 2006. New insights into the phylogeny of the genus Hymenophyllum s.l. (Hymenophyllaceae): revealing the polyphyly of Mecodium. Syst. Bot. 31: 271284.Google Scholar
Hennequin, S., Kessler, M., Lindsay, S., and Schneider, H. 2014. Evolutionary patterns in the assembly of fern diversity on the oceanic Mascarene Islands. J. Biogeogr. 41: 16511663.Google Scholar
Hennig, W. 1966. Phylogenetic Systematics, trans. Davis, D. D. and Zangerl, R.. Urbana: University of Illinois Press.Google Scholar
Henry, R. J. 2013. Evolution of DNA marker technology in plants. In Henry, R. J. (ed.), Molecular Markers in Plants (pp. 419). West Sussex: Wiley-Blackwell.Google Scholar
Hernández, R., and Monleon, J. 1975. La comunidad de pescadores de Juan Fernández. In Orellana R., M. (ed.), Las Islas de Juan Fernández. Historia, Arqueología y Antropología de la Isla Robinson Crusoe (pp. 137153). Santiago: Depto. de Ciéncias Antropológicas y Arqueológicas, Universidad de Chile.Google Scholar
Herzog, T. 1942. Die foliosen Lebermoose der Juan Fernandez.-Inseln und der Osterinsel. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 697752). Uppsala: Almqvist & Wiksells.Google Scholar
Heusser, C. J. 1977. Carlos Muñoz Pizarro (1913–1976). In Prance, G. T. and Elias, T. S. (eds.), Extinction Is Forever (pp. 248250). New York: NewYork Botanical Garden.Google Scholar
Heywood, V. H. 2011. The hazardous future of island floras. In Bramwell, D. and Caujapé-Castells, J. (eds.), The Biology of Island Floras (pp. 488510). Cambridge: Cambridge University Press.Google Scholar
Hill, M. O. 1979. Twinspan: A Fortran Program for Arranging Multivariate Data in an Ordered Two-Way Table by Classification of the Individuals and Attributes. Ithaca, NY: Cornell University Press.Google Scholar
Hind, N., and Johnson, N. 2006. Dendroseris litoralis, Compositae. Curtis’s Bot. Mag. 23: 314324.Google Scholar
Hobdy, R. 1993. Lāna‘i: a case study: the loss of biodiversity on a small Hawaiian island. Pacific Sci. 47: 201210.Google Scholar
Hoeneisen, M., Becerra, J., Bittner, M., Silva, M., Baeza, C., et al. 1998. Sesquiterpenes of the essential oil of Santalum fernandezianum F. Phil (Santalaceae), an extinct species of the Juan Fernández Islands, Chile. Bol. Soc. Chil. Quim. 43: 505508.Google Scholar
Hoggard, R. K., Kores, P. J., Molvray, M., Hoggard, G. D., and Broughton, D. A. 2003. Molecular systematics and biogeography of the amphibious genus Littorella (Plantaginaceae). Am. J. Bot. 90: 429435.Google Scholar
Holttum, R. E. 1966. The genus Arthropteris J. Sm. in Malesia. Blumea 14: 225229.Google Scholar
Holttum, R. E. 1986. Studies in the fern-genera allied to Tectaria Cav: VI. A conspectus of genera in the Old World regarded as related to Tectaria, with descriptions of two genera. Gard. Bull. Straits Settlem. 39: 153167.Google Scholar
Hörandl, E. 2014. Nothing in taxonomy makes sense except in the light of evolution: examples from the classification of Ranunculus. Ann. Missouri Bot. Gard. 100: 1431.Google Scholar
Hörandl, E., and Appelhaus, M. S. (eds.). 2015. Next-Generation Sequencing in Plant Systematics. Königstein: Koeltz Scientific Books.Google Scholar
Hörandl, E., and Emadzade, K. 2012. Evolutionary classification: a case study on the diverse plant genus RanunculusL. (Ranunculaceae). Perspect. Plant Ecol. Evol. Syst. 14: 310324.Google Scholar
Hörandl, E., Paun, O., Johansson, J. T., Lehnebach, C., Armstrong, T., et al. 2005. Phylogenetic relationships and evolutionary traits in Ranunculus s.l. (Ranunculaceae) inferred from ITS sequence analysis. Mol. Phylogenet. Evol. 36: 305327.Google Scholar
Horner, H. T., Wanke, S., and Samain, M.-S. 2009. Evolution and systematic value of leaf crystal macropatterns in the genus Peperomia (Piperaceae). Int. J. Plant Sci. 170: 343354.Google Scholar
Horres, R., Schulte, K., Weising, K., and Zizka, G. 2007. Systematics of Bromelioideae (Bromeliaceae): evidence from molecular and anatomical studies. Aliso 23: 2743.Google Scholar
Hortal, J., Triantis, K. A., Meiri, S., Thébault, E., and Sfenthourakis, S. 2009. Island species richness increases with habitat diversity. Am. Nat. 174(6): E205–17.Google Scholar
Howarth, D. G., and Baum, D. A. 2005. Genealogical evidence of homoploid hybrid speciation in an adaptive radiation of Scaevola (Goodeniaceae) in the Hawaiian Islands. Evolution 59: 948961.Google Scholar
Howell, J. 1829. The Life and Adventures of Alexander Selkirk; Containing the Real Incidents upon which the Romance of Robinson Crusoe Is Founded. Edinburgh: Oliver & Boyd.Google Scholar
Hu, X-S. 2015. Mating system as a barrier to gene flow. Evolution 69: 11581177.Google Scholar
Hubbs, C. L., and Norris, K. S. 1971. Original teeming abundance, supposed extinction, and survival of the Juan Fernández fur seal. Antarctic Res. Ser. 18: 3552.Google Scholar
Hubby, J. L., and Lewontin, R. C. 1966. A molecular approach to the study of genic heterozygosity in natural populations: I. The number of alleles at different loci in Drosophila pseudoobscura. Genetics 54: 577594.Google Scholar
Huene, R., von, Corvalán, J., Flueh, E. R., Hinz, K., Korstgard, J., et al. and the CONDOR Scientists. 1997. Tectonic control of the subducting Juan Fernández Ridge on the Andean margin near Valparáiso, Chile. Tectonica 16: 474488.Google Scholar
Humphries, C. J., and Richardson, P. M. 1980. Hennig’s methods and phytochemistry. In Bisby, F. A., Vaughan, J. G., and Wright, C. A. (eds.), Chemosystematics: Principles and Practice (pp. 3533780). London: Academic Press.Google Scholar
Hunter, M. L. Jr., and Gibbs, J. P. 2007. Fundamentals of Conservation Biology. Malden, MA: Blackwell.Google Scholar
Hurr, K. A., Lockhar, P. J., Heenan, P. B., and Penny, D. 1999. Evidence for the recent dispersal of Sophora (Leguminosae) around the Southern Oceans: molecular data. J. Biogeogr. 26: 565577.Google Scholar
Ingram, K., Bernardello, G., Cover, S., and Wilson, E. 2006. The ants of the Juan Fernández Islands: genesis of an invasive fauna. Biol. Invas. 8: 383387.Google Scholar
IUCN (International Union for the Conservation of Nature). 2001. Categories and Criteria (version 3.1). Available at www.redlist.org/info/categories_criteria2001.html.Google Scholar
Iwashina, T., Kitajima, J., Shiuchi, T., and Itou, Y. 2005. Chalcones and other flavonoids from Asarum sensu lato (Aristolochiaceae). Biochem. Syst. Ecol. 33: 571584.Google Scholar
Iwatsuki, K. 1990. Hymenophyllaceae. In Kramer, K. U. and Green, P. S. (vol. eds.), Pteridophytes and Gymnosperms. In Kubitzki, K. (gen. ed.), The Families and Genera of Vascular Plants, vol. I (pp. 157163). Berlin: Springer-Verlag.Google Scholar
Jaksic, F. M., and Fuentes, E. R. 1991. Ecology of a successful invader: the European rabbit in central Chile. In Groves, R. H. and di Castri, F. (eds.), Biogeography of Mediterranean Invasions (pp. 273283). Cambridge: Cambridge University Press.Google Scholar
Jakubowsky, G., and Stuessy, T. F. 1999. Hybridization and evolution in Eryngium (Apiaceae) of the Juan Fernández Islands. Syst. Assoc. Biennial Conference, Glasgow, p. 23 (Abstr.)Google Scholar
Johannsen, A. 1937. A Descriptive Petrography of the Igneous Rocks III. Chicago: University of Chicago Press.Google Scholar
Johow, F. 1892–93. Los helechos de Juan Fernández. Anal. Univ. Chile 82: 741757, 9771005 (the two parts were also published together in 1893, 46 pp., perhaps privately, and printed by Imprenta Cervantes: Santiago).Google Scholar
Johow, F. 1893. Las plantas de cultivo en Juan Fernández. Anal. Univ. Chile 84: 939970.Google Scholar
Johow, F. 1896. Estudios Sobre la Flora de las Islas de Juan Fernández. Santiago: Gobierno de Chile.Google Scholar
Jones, K. E., Reyes-Betancort, J.A., Hiscock, S. J., and Carine, M. A. 2014. Allopatric diversification, multiple habitat shifts and hybridization in the evolution of Pericallis (Asteraceae), a Macaronesian endemic genus. Am. J. Bot. 101: 637651.Google Scholar
Juan, J., and de Ulloa, A. 1748. Relación Histórica del Viaje a la América Meridional. Madrid [seen in Saumell, A. (ed.). 1990. Antonio de Ulloa, Viaje a la America Meridional, Tomo B: Segunda parte del Viage al Reyno del Peru. Madrid: Historia 16].Google Scholar
Juan, J., and de Ulloa, A. 1826. Noticias Secretas de América, sobre el Estado Naval, Militar, y Político de los Reynos del Perú y Provincias de Quito, Costas de Nueva Granada y Chile. London.Google Scholar
Judd, W. S., Campbell, C. S., Kellogg, E. A., Stevens, P. F., and Donoghue, M. J. 2007. Plant Systematics: A Phylogenetic Approach (3rd edn.). Sunderland, MA: Sinauer Associates.Google Scholar
Judziewicz, E. J., Clark, L. G., Londoño, X., and Stern, M. J. 1999. American Bamboos. Washington, DC: Smithsonian Institution Press.Google Scholar
Kadereit, G., Borsch, T., Weising, K., and Freitag, H. 2003. Phylogeny of Amaranthaceae and Chenopodiaceae and the evolution of C4 photosynthesis. Int. J. Plant Sci. 164: 959986.Google Scholar
Kadereit, G., Mavrodiev, E. V., Zacharias, E. H., and Sukhorukov, A. P. 2010. Molecular phylogeny of Atripliceae (Chenopodioideae, Chenopodiaceae): implications for systematics, biogeography, flower and fruit evolution, and the origin of C4 photosynthesis. Am. J. Bot. 97: 16641687.Google Scholar
Kadereit, J. W., Repplinger, M., Schmaltz, N., Uhink, C. H., and Wörz, A. 2008. The phylogeny and biogeography of Apiaceae subf. Saniculoideae tribe Saniculeae: from south to north and south again. Taxon 57: 365382.Google Scholar
Kang, D.-L., Pan, Q.-H., and Yi, Z.-L. 2008. Phylogenetic study on sect. Duretia (Boehmeria,Urticaceae) based on nrDNA ITS sequences. Plant Sci. J. 26: 450453.Google Scholar
Karrenberg, S., and Widmer, A. 2008. Ecologically relevant genetic variation from a non-Arabidopsis perspective. Curr. Opin. Plant Biol. 11: 156162.Google Scholar
Kausel, E. 1949. Comentario sobre las ericáceas y epacridáceas chilenas. Rev. Universitaria 34: 155178 (Anal. Acad. Chil. Cienc. Nat. 14).Google Scholar
Kawakami, K., and Okochi, I. 2010. Restoring the Oceanic Island Ecosystem: Impact and Management of Invasive Alien Species in the Bonin Islands. Tokyo: Springer.Google Scholar
Keissler, K. 1927. Ascomyceten, Fungi imperfecti und Uredineen von Juan Fernandez. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 473486). Uppsala: Almqvist & Wiksells.Google Scholar
Kelchner, S. A., and Clark, L. G. 1997. Molecular evolution and phylogenetic utility of the chloroplast rpl16 intron in Chusquea and the Bambusoideae (Poaceae). Mol. Phylogenet. Evol. 8: 385397.Google Scholar
Kelchner, S. A., and Bamboo Phylogeny Group. 2013. Higher level phylogenetic relationships within the bamboos (Poaceae: Bambusoideae) based on five plastid markers. Mol. Phylogenet. Evol. 67: 404413.Google Scholar
Kelly, L. M., and González, F. 2003. Phylogenetic relationships in Aristolochiaceae. Syst. Bot. 28: 236249.Google Scholar
Kendrew, W. G. 1961. The Climates of the Continents. Oxford: Oxford University Press.Google Scholar
Kerr, M. S. 2004. A phylogenetic and biogeographic analysis of Sanguisorbeae (Rosaceae), with emphasis on the Pleistocene radiation of the high Andean genus Polylepis. Ph.D. dissertation, University of Maryland, College Park.Google Scholar
Khan, M. Q., and Narayan, R. K. J. 2007. Phylogenetic diversity and relationships among species of genus Nicotiana using RAPDs analysis. Afr. J. Biotech. 6: 148162.Google Scholar
Kiehn, M., Jodl, M., and Jakubowsky, G. 2005. Chromosome numbers of angiosperms from the Juan Fernández Islands, the Tristan da Cunha Archipelago, and from mainland Chile. Pacific Sci. 59: 453460.Google Scholar
Kilian, N., Gemeinholzer, B., and Lack, H. W. 2009. Cichorieae. In Funk, V. A., Susanna, A., Stuessy, T. F. and Bayer, R. J. (eds.), Systematics, Evolution and Biogeography of the Compositae (pp. 341348). Vienna: International Association for Plant Taxonomy.Google Scholar
Kim, C., Deng, T., Chase, M., Zhang, D.-G., Nie, Z.-L., et al. 2015. Generic phylogeny and character evolution in Urticeae (Urticaceae) inferred from nuclear and plastid DNA regions. Taxon 64: 6578.Google Scholar
Kim, S.-C., Crawford, D. J., Francisco-Ortega, J., and Santos-Guerra, A. 1996a. A common origin for woody Sonchus and five related genera in the Macaronesian islands: molecular evidence for extensive radiation. Proc. Natl. Acad. Sci. USA 93: 77437748.Google Scholar
Kim, S.-C., Crawford, D. J., and Jansen, R. K. 1996b. Phylogenetic relationships among the genera of the subtribe Sonchinae (Asteraceae): evidence from ITS sequences. Syst. Bot. 21: 417432.Google Scholar
Kim, S.-C., Crawford, D. J., Jansen, R. K., and Santos-Guerra, A. 1999. The use of a non-coding region of chloroplast DNA in phylogenetic studies of the subtribe Sonchinae (Asteraceae: Lactuceae). Plant Syst. Evol. 215: 8599.Google Scholar
Kim, S.-C., Lee, C., and Mejías, J. A. 2007. Phylogenetic analysis of chloroplast DNA matK gene and ITS of nrDNA sequences reveals polyphyly of the genus Sonchus and new relationships among the subtribe Sonchinae (Asteraceae: Cichorieae). Mol. Phylogenet. Evol. 44: 578597.Google Scholar
Kim, Y.-D., Kim, S.-H., and Landrum, L. R. 2004. Taxonomic and phytogeographic implications from ITS phylogeny in Berberis (Berberidaceae). J. Plant Res. 117: 175182.Google Scholar
Kingston, N., Waldren, S., and Bradley, U. 2003. The phytogeographical affinities of the Pitcairn Islands: a model for south-eastern Polynesia? J. Biogeogr. 30: 13111328.Google Scholar
Kirk, H., and Freeland, J. R. 2011. Applications and implications of neutral versus non-neutral markers in molecular ecology. Int. J. Mol. Sci. 11: 39663988.Google Scholar
Kirschner, J. (ed.). 2002. Rostkovia to Luzula. Species Plantarum: Flora of the World, Part 6: Juncaceae 1. Canberra: Australian Biological Resources Study.Google Scholar
Knapp, S., Chase, M. W., and Clarkson, J. J. 2004. Nomenclatural changes and a new sectional classification in Nicotiana (Solanaceae). Taxon 53: 7382.Google Scholar
Koelling, V. A., Monnahan, P. J., and Kelly, J. K. 2012. A Bayesian method for the joint estimation of outcrossing rate and inbreeding depression. Heredity 109: 393400.Google Scholar
Komarnytskyy, S. I. 2005. Molecular phylogeny of chloroplast DNA of Nicotiana species. Tsitol Genet. 39(4): 1319 (in Russian).Google 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. Mol. Phylogenet. Evol. 39: 830845.Google Scholar
Kover, P. X., and Mott, R. 2012. Mapping the genetic basis of ecologically and evolutionarily relevant traits in Arabidopsis thaliana. Curr. Opin. Plant Biol. 15: 212217.Google Scholar
Koyama, T. 1972. Cyperaceae–Rhynchosporeae and Cladieae. In Maguire, B. et al. (eds.), The Botany of the Guayana Highland, Part IX. Mem. New York Bot. Gard. 23: 2389. (cf. p. 89).Google Scholar
Kramer, K. U. 1990. Dicksoniaceae. In Kramer, K. U. and Green, P. S. (vol. eds.), Pteridophytes and Gymnosperms, in Kubitzki, K. (ed.), The Families and Genera of Vascular Plants, vol. 1 (pp. 9499). Berlin: Springer-Verlag.Google Scholar
Kramer, K. U. 1993. Distribution patterns in major pteridophyte taxa relative to those of angiosperms. J. Biogeogr. 20: 287291.Google Scholar
Kramer, K. U., and Green, P. S. (eds.). 1990. Pteridophytes and gymnosperms. In Kubitzki, K. (ed.), The Families and Genera of Vascular Plants, vol. 1. Berlin: Springer-Verlag.Google Scholar
Kraus, G. 1882. Ueber das Sandelholz von Juan Fernandez. Ber. Sitzungen Naturf. Ges. Halle 1882: 89.Google Scholar
Kueffer, C., Daehler, C. C., Torre-Santana, C. W., Lavergne, C., Meyer, J. Y., et al. 2010. A global comparison of plant invasions on oceanic islands. Perspect. Plant Ecol. Evol. Syst. 12: 145161.Google Scholar
Kunkel, G. 1957. Beobachtungen über die Vegetation auf dem Yunque-Massiv. Bot. Jahrb. 77: 149157.Google Scholar
Kunkel, G. 1965. Catalogue of the pteridophytes of the Juan Fernández Islands (Chile). Nova Hedwigia 9: 245284.Google Scholar
Kuntze, O. 1898. Revisio Generum Plantarum, pars 3. Leipzig: Arthur Felix. (cf. p. 77).Google Scholar
Kunze, G. 1834. Thyrsopteris, eine neue Farngattung. Linnaea 9: 506508.Google Scholar
Kuschel, G. 1952. Los insectos de las islas Juan Fernández. Introducción. Revista Chil. Entomol. 2: 36.Google Scholar
Labiak, P. H., Sundue, M., Rouhan, G., Hanks, J. G., Mickel, J. T., et al. 2014. Phylogeny and historical biogeography of the lastreopsid ferns (Dryopteridaceae). Am. J. Bot. 101: 12071228.Google Scholar
Lack, D. 1947. Darwin’s Finches: An Essay on the General Biological Theory of Evolution. Cambridge: Cambridge University Press.Google Scholar
Lack, H. W., Sack, G., and Bachmann, K. 1978. The genome of Dendroseris litoralis, an arboreal insular endemic of the tribe Lactuceae (Asteraceae). Beitr. Biol. Pflanzen 54: 425441.Google Scholar
Lambelet-Haueter, C. 1990. Mauvaises herbes et flore anthropogène: II. Classifications et catégories. Saussurea 22: 4981.Google Scholar
Lambelet-Haueter, C. 1991. Mauvaises herbes et flore anthropogène: I. Définitions, concepts et caractéristiques écologiques. Saussurea 21: 4773.Google Scholar
Lammers, T. G. 1996. Phylogeny, biogeography, and systematics of the Wahlenbergia fernandeziana complex (Campanulaceae: Campanuloideae). Syst. Bot. 21: 397415.Google Scholar
Lammers, T. G., Stuessy, T. F., and Silva O., M. 1986. Systematic relationships of the Lactoridaceae, an endemic family of the Juan Fernández Islands, Chile. Plant Syst. Evol. 152: 243266.Google Scholar
Lampert, A., Hastings, A., Grosholz, E. D., Jardine, S. L., and Sanchirico, J. N. 2014. Optimal approaches for balancing invasive species eradication and endangered species management. Science 344: 10281031.Google Scholar
Landrum, L. R. 1981a. A monograph of the genus Myrceugenia (Myrtaceae). Fl. Neotrop. Monogr. 29: 1135.Google Scholar
Landrum, L. R. 1981b. The phylogeny and geography of Myrceugenia (Myrtaceae). Brittonia 33: 105129.Google Scholar
Landrum, L. R. 1988. The myrtle family (Myrtaceae) in Chile. Proc. Calif. Acad. Sci. 45: 277317.Google Scholar
Landrum, L. R. 1999. Revision of Berberis (Berberidaceae) in Chile and adjacent southern Argentina. Ann. Missouri Bot. Gard. 86: 793834.Google Scholar
Långström, E., and Chase, M. W. 2002. Tribes of Boraginoideae (Boraginaceae) and placement of Antiphytum, Echiochilon, Ogastemma and Sericostoma: a phylogenetic analysis based on atpB plastid DNA sequence data. Plant Syst. Evol. 234: 137153.Google Scholar
Larsen, C., Ponce, M. M., and Scataglini, M. A. 2013. Revisión de las especies de Hymenophyllum (Hymenophyllaceae) del sur de Argentina y Chile. Gayana Bot. 70: 274329.Google Scholar
Laurance, W. F. 2004. The perils of payoff: corruption as a threat to global biodiversity. Trends Ecol. Evol. 19: 399401.Google Scholar
Lee, J., Park, B.-J., Tsunetsugu, Y., Ohira, T., Kagawa, T., et al. 2011. Effect of forest bathing on physiological and psychological responses in young Japanese male subjects. Publ. Health 125: 93100.Google Scholar
Lee, N., and Hauk, W. 2013. Phylogenetic relationships among species of adders tongue ferns (Ophioglossum s.s., Ophioglossaceae). Botany Conference 2013 (Abstr.).Google Scholar
Lee, W. G., Macmillan, B. H., Partridge, T. R., Lister, R., and Lloyd, K. M. 2001. Fruit features in relation to the ecology and distribution of Acaena (Rosaceae) species in New Zealand. New Zealand J. Ecol. 25: 1727.Google Scholar
Lehnebach, C. A. 2003. Preliminary checklist of the orchids of Chile. Bot. J. Linn. Soc. 143: 449451.Google Scholar
Leinonen, T., O’Hara, R. B., Cano, J. M., and Merilä, J. 2008. Comparative studies of quantitative trait and neutral marker divergence: a meta-analysis. J. Evol. Biol. 21: 117.Google Scholar
Leiva, I., Schiller, R., Chamorro, O., López, B., Araya, G., et al. 2013. Nuevos registros sobre especies En Peligro Crítico de la flora de las islas Robinson Crusoe y Santa Clara, Chile. Gayana Bot. 70: 377382.Google Scholar
Leskinen, E., and Alström-Rapaport, C. 1999. Molecular phylogeny of Salicaceae and closely related Flacourtiaceae: evidence from 5.8S, ITS1 and ITS 2 of the rDNA. Plant. Syst. Evol. 215: 209227.Google Scholar
Levin, D. A. 2002. The Role of Chromosomal Change in Plant Evolution. New York: Oxford University Press.Google Scholar
Levin, D. A. 1978. The origin of isolating mechanisms in flowering plants. Evol. Biol. 11: 185317.Google Scholar
Levin, D. A. 2000. The Origin, Expansion, and Demise of Plant Species. New York: Oxford University Press.Google Scholar
Levin, D. A. 2004. The ecological transition in speciation. New Phytol. 161: 9196.Google Scholar
Levring, T. 1941. Die Meeresalgen der Juan Fernandez-Inseln. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 601670). Uppsala: Almqvist & Wiksells.Google Scholar
Levsen, N. D., Crawford, D. J., Archibald, J. K., Santos-Guerra, A., and Mort, M. E. 2008. Nei’s to Bayes’: comparing methods for estimating patterns of genetic variation in Tolpis (Asteraceae). Am. J. Bot. 95: 14661474.Google Scholar
Lewis, H. 1966. Speciation in flowering plants. Science 152: 167172.Google Scholar
Lewis, O. 1949. Sea Routes to the Gold Fields: The Migration by Water to California in 1849–1852. New York: Knopf.Google Scholar
Li, Q. 2010. Effect of forest bathing trips on human immune function. Environm. Health Prevent. Med. 15: 917.Google Scholar
Lichtenstein, J. 1944. Las Ophioglosáceas de la Argentina, Chile y Uruguay. Darwiniana 6: 380441.Google Scholar
Linklater, E. 1972. The Voyage of the Challenger. London: John Murray.Google Scholar
Linnaeus, C. 1753. Species Plantarum, vol. 2. Stockholm.Google Scholar
Linsley, E. 1966. Pollinating insects of the Galápagos Islands. In Bowman, R. I. (ed.), The Galápagos (pp. 225232). Berkeley: University of California Press.Google Scholar
Liu, H.-M., Jiang, R.-H., Guo, J., Hovenkamp, P., Perrie, L. R., et al. 2013. Towards a phylogenetic classification of the climbing fern genus Arthropteris. Taxon 62: 688700.Google Scholar
Liu, H.-M., Zhang, X.-C., Wang, W., Qiu, Y.-L., and Chen, Z.-D. 2007. Molecular phylogeny of the fern family Dryopteridaceae inferred from chloroplast rbcL and atpB genes. Int. J. Plant Sci. 168: 13111323.Google Scholar
Liu, Z., Lu, Z., Wen, X., Otto-Bliesner, B. L., Timmermann, A., et al. 2014. Evolution and forcing mechanisms of El Niño over the past 21,000 years. Nature 515: 550553.Google Scholar
Lloyd, R., and Klekowski, E. 1970. Spore germination and viability in Pteridophyta: evolutionary significance of chlorophyllous spores. Biotropica 2: 129137.Google Scholar
Lomolino, M. V., Riddle, B. R., and Brown, J. H. 2005. Biogeography (3rd edn.). Sunderland, MA: Sinauer Associates.Google Scholar
Lomolino., M. V., Brown, J. H., and Sax, D. F. 2010. Island biogeography theory: reticulations and reintegration of “a biogeography of the species.” In Losos, J. B. and Ricklefs, R. E. (eds.), The Theory of Island Biogeography Revisited (pp. 1351). Princeton, NJ: Princeton University Press.Google Scholar
Lönnberg, E. 1921. The birds of the Juan Fernandez Islands. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Islands, vol. 3: Zoology (pp. 117). Uppsala: Almqvist & Wiksells.Google Scholar
Looser, G. 1927. La zarzamora (Rubus ulmifolius Schott) en Juan Fernández. Rev. Chil. Hist. Nat. 31: 8485.Google Scholar
Looser, G. 1944. Sinopsis de los Asplenium (Filices) de Chile. Lilloa 10: 233264.Google Scholar
Looser, G. 1945. Notas sobre helechos chilenos–III. Rev. Chil. Hist. Nat. 48: 110111.Google Scholar
Looser, G. 1947. Los Blechnum (Filices) de Chile. Rev. Univ. (Univ. Católica de Chile) 32: 7104.Google Scholar
López-Sepúlveda, P., Baeza, M., Peñailillo, P., Stuessy, T., Greimler, J., et al. 2013a. Introduced plant species in the Biosphere Reserve Alejandro Selkirk Island, Robinson Crusoe archipelago, Chile, and their threat to the native and endemic vegetation. Mem. New York Bot. Gard. 108: 223243.Google Scholar
López-Sepúlveda, P., Takayama, K., Crawford, D. J., Greimler, J., Peñailillo, P., et al. 2016. Genetic consequences and biogeography of anagenetic speciation of Rhaphithamnus venustus (Verbenaceae) in the Juan Fernández archipelago, Chile, based on AFLPs and SSRs. Plant Spec. Biol 32: 223237.Google Scholar
López-Sepúlveda, P., Takayama, K., Greimler, J., Crawford, D. J., Peñailillo, P., et al. 2015a. Speciation and biogeography of Erigeron (Asteraceae) in the Juan Fernández Archipelago, Chile, based on AFLPs and SSRs. Syst. Bot. 40: 888899.Google Scholar
López-Sepúlveda, P., Takayama, K., Greimler, J., Crawford, D. J., Peñailillo, P., et al. 2015b. Progressive migration and anagenesis in Drimys confertifolia of the Juan Fernández Archipelago, Chile. J. Plant Res. 128: 7390.Google Scholar
López-Sepúlveda, P., Takayama, K., Greimler, J., Peñailillo, P., Crawford, D. J., et al. 2013b. Genetic variation (AFLPs and nuclear microsatellites) in two anagenetically derived endemic species of Myrceugenia (Myrtaceae) on the Juan Fernández Islands, Chile. Am. J. Bot. 100: 722734.Google Scholar
Lord, J. 1999. Fleshy-fruitedness in the New Zealand flora. J. Biogeogr. 26: 12491253.Google Scholar
Losos, J. B., and Parent, C. E. 2010. The speciation-area relationship. In Losos, J. B. and Ricklefs, R. E. (eds.), The Theory of Island Biogeography Revisited (pp. 415438). Princeton, NJ: Princeton University Press.Google Scholar
Losos, J. B., and Ricklefs, R. E. (eds.). 2010. The Theory of Island Biogeography Revisted. Princeton, NJ: Princeton University Press.Google Scholar
Lowrey, T. K. 1986. A biosystematic revision of Hawaiian Tetramolopium (Compositae: Astereae). Allertonia 4: 1123.Google Scholar
Lowrey, T. K., and Crawford, D. J. 1985. Allozyme divergence and evolution in Tetramolopium (Compositae: Astereae) on the Hawaiian Islands. Syst. Bot. 10: 6472.Google Scholar
Lowry, D. B., and Gould, B. A. 2016. Speciation continuum. In Kliman, R. (gen. ed.) and Ortiz-Barrientos, D. (section ed.), Encyclopedia of Evolutionary Biology (pp. 159165). New York: Academic Press.Google Scholar
Lowry, D. B., Modliszewski, J. L., Wright, K. M., Wu, C. A., and Willis, J. H. 2008. The strength and genetic basis of reproductive isolating barriers in flowering plants. Philos. Trans. R. Soc. Lond. B 363: 30093021.Google Scholar
Lowry, P. P. II. 2009. Patterns of species richness, endemism, and diversification in oceanic island floras. In Wolanski, E. (ed.), Oceans and Aquatic Ecosystems, vol. 2 (pp. 201220). Paris: Encyclopedia of Life Support Systems (EOLSS).Google Scholar
Lozano, G., and Murillo-A., J. 2001. El género Dysopsis (Euphorbiaceae). Caldasia 23: 419426.Google Scholar
Lu, J.-M., Barrington, D. S., and Li, D.-Z. 2007. Molecular phylogeny of the polystichoid ferns in Asia based on rbcL sequences. Syst. Bot. 32: 2633.Google Scholar
Luteyn, J. L. 1995. Pernettya Gaudichaud. In Luteyn, J. L. (ed.), Ericaceae, Part II: The Superior-Ovaried Genera (Monotropoideae, Pyroloideae, Rhododendroideae, and Vaccinioideae p.p.) (pp. 365383). New York: New York Botanical Garden (Flora Neotropica Monograph 66).Google Scholar
Lynch, J. 2009. San Martín: Argentine Soldier, American Hero. New Haven, CT: Yale University Press.Google Scholar
Mabberley, D. J. 2008. Mabberley’s Plant-Book: A Portable Dictionary of Plants, Their Classification and Uses (3rd edn.). Cambridge: Cambridge University Press.Google Scholar
Mabry, T. J., Markham, K. R., and Thomas, M. B. 1970. The Systematic Identification of Flavonoids. New York: Springer-Verlag.Google Scholar
MacArthur, R. H., and Wilson, E. O. 1967. The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
Macdonald, G. A., Abbott, A. T., and Peterson, F. L. 1983. Volcanoes in the Sea: The Geology of Hawaii (2nd edn.). Honolulu: University of Hawaii Press.Google Scholar
MacPhail, M. K., Partridge, A. D., and Truswell, E. M. 1999. Fossil pollen records of the problematical primitive angiosperm family Lactoridaceae in Australia. Plant Syst. Evol. 214: 199210.Google Scholar
Magee, B. 1795. Observations on the Islands of Juan Fernandez, Massafuero, and St. Ambrose, in the South Pacific Ocean, and the Coast of Chili, in South-America. Extracted from the Journal of Mr. Bernard Magee, First Officer of the Ship Jefferson, in Her Late Voyage Round the Globe. Boston: Massachusetts Historical Society (reprint from Johnson Reprint Corporation, 1968).Google Scholar
Magnacca, K. N., Foote, D., and O’Grady, P. M. 2008. A review of the endemic Hawaiian Drosophilidae and their host plants. Zootaxa 1728: 158.Google Scholar
Malhi, Y., Roberts, J. T., Betts, R. A., Killeen, T. J., Li, W., et al. 2008. Climate change, deforestation, and the fate of the Amazon. Science 319: 169172.Google Scholar
Mallet, J. 2007. Hybrid speciation. Nature 446: 279283.Google Scholar
Mammerickx, J., and Smith, S. M. 1978. Bathymetry of the Southeast Pacific. Geol. Soc. Amer. Map and Chart Series MC–26.Google Scholar
Mandáková, T., Heenan, P. B., and Lysak, M. A. 2010. Island species radiation and karyotypic stasis in Pachycladon allopolyploids. BMC Evol. Biol. 10: 367.Google Scholar
Manen, J.-F. 2000. Relaxation of constraints in promoters of the plastid gene atpB in a particular Rubiaceae lineage. Plant Syst. Evol. 224: 235241.Google Scholar
Manen, J.-F., and Natali, A. 1995. Comparison of the evolution of ribulose-1,5-biphosphate carboxylase (rbcL) and atpB-rbcL noncoding spacer sequences in a recent plant group, the tribe Rubieae (Rubiaceae). J. Mol. Evol. 41: 920927.Google Scholar
Manen, J.-F., Natali, A., and Ehrendorfer, F. 1994. Phylogeny of Rubiaceae-Rubieae inferred from the sequence of a cpDNA intergene region. Plant Syst. Evol. 190: 195211.Google Scholar
Mantese, A., and Medan, D. 1993. Anatomía y arquitectura foliares de Colletia y Adolphia (Rhamnaceae). Darwiniana 32: 9197.Google Scholar
Markham, K. R. 1982. Techniques of Flavonoid Identification. New York: Academic Press.Google Scholar
Marquínez, X., Lohmann, L. G., Faria Salatino, M. L., Salatino, A., and González, F. 2009. Generic relationships and dating of lineages in Winteraceae based on nuclear (ITS) and plastid (rpS16 and psbA-trnH) sequence data. Mol. Phylogenet. Evol. 53: 435449.Google Scholar
Marticorena, A. 1996. Revisión del género Acaena Mutis ex L. (Rosaceae) en Chile. Master’s thesis, Universidad de Concepción, Chile.Google Scholar
Marticorena, A. 2006. Revisión del género Acaena (Rosaceae) en Chile. Ann. Missouri Bot. Gard. 93: 412454.Google Scholar
Marticorena, C. 1990. Contribución a la estadística de la flora chilena. Gayana Bot. 47: 85113.Google Scholar
Marticorena, C., and Baeza, C. 2001. Piperaceae. In Marticorena, C. and Rodríguez, R. (eds.), Flora de Chile, vol. 2: Winteraceae-Ranunculaceae (pp. 2632). Concepción: Universidad de Concepción, Chile.Google Scholar
Marticorena, C., and Quezada, M. 1985. Catálogo de la flora vascular de Chile. Gayana Bot. 42: 1157.Google Scholar
Marticorena, C., Stuessy, T. F., and Baeza, C. M. 1998. Catalogue of the vascular flora of the Robinson Crusoe or Juan Fernández Islands, Chile. Gayana Bot. 55: 187211.Google Scholar
Martin, N. H., and Willis, J. H. 2007. Ecological divergence associated with mating system causes nearly complete reproductive isolation between sympatric Mimulus species. Evolution 61: 6882.Google Scholar
Martin, T. E. 1981. Limitation in small habitat islands: chance or competition? Auk 98: 715734.Google Scholar
Martínez, J. A. 1983. Gualterio Looser Schallemberg. Un naturalista botánico (1898–1982). Rev. Chil. Hist. Nat. 56: 9195.Google Scholar
Marx, H. E., O’Leary, N., Yuan, Y.-W., Lu-Irving, P., Tank, D. C., et al. 2010. A molecular phylogeny and classification of Verbenaceae. Am. J. Bot. 97: 16471663.Google Scholar
Matthei, O. 1986. El género Bromus L. (Poaceae) en Chile. Gayana Bot. 43: 47110.Google Scholar
Matthei, O. 1995. Manual de las Malezas que Crecen en Chile. Santiago de Chile: published by author.Google Scholar
Matthei, O., Marticorena, C., and Stuessy, T. F. 1993. La flora adventicia del archipiélago de Juan Fernández. Gayana Bot. 50: 69102.Google Scholar
Matthei, O. R. 1974. El género Megalachne Steudel (Gramineae). Bol. Soc. Biol. Concepción 48: 165172.Google Scholar
Mattick, F. 1964. Übersicht über die Florenreiche und Florengebiete der Erde. In Melchior, H. (ed.), A. Engler’s Syllabus der Pflanzenfamilien, Bd. 2, 12te Auflage (pp. 626629). Berlin: Gebrueder Borntraeger.Google Scholar
Mayr, E. 1942. Systematics and the Origin of Species from the Viewpoint of a Zoologist. New York: Columbia University Press.Google Scholar
Mayr, E. 1954. Change of genetic environment and evolution. In Huxley, J., Hardy, A. C. and Ford, E. B. (eds.), Evolution as a Process (pp. 157180). London: Allen & Unwin.Google Scholar
Mayr, E. 1963. Animal Species and Evolution. Cambridge, MA: Belknap Press of Harvard University Press.Google Scholar
McAtee, W. L. 1947. Distribution of seeds by birds. Am. Midl. Nat. 38: 214223.Google Scholar
McCall, G. 2009. Easter Island. In Gillespie, R. G. and Clague, D. A. (eds.), Encyclopedia of Islands (pp. 244251). Berkeley: University of California Press.Google Scholar
McGlone, M. S., Duncan, R. P., and Heenan, P. B. 2001. Endemism, species selection and the origin and distribution of the vascular plant floras of New Zealand. J. Biogeogr. 28: 199216.Google Scholar
McGlone, M. S., Kershaw, A. P., and Markgraf, V. 1992. El Niño/Southern Oscillation climatic variability in Australasian and South American paleoenvironmental records. In Diaz, H. F. and Markgraf, V. (eds.), El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation (pp. 435462). Cambridge: Cambridge University Press.Google Scholar
McHenry, M. A., and Barrington, D. S. 2014. Phylogeny and biogeography of exindusiate Andean Polystichum (Dryopteridaceae). Am. J. Bot. 101: 365375.Google Scholar
McKay, J. K., and Latta, R. G. 2002. Adaptive population divergence: markers, QTL and traits. Trends Ecol. Evol. 17: 285291.Google Scholar
McMullen, C. 1987. Breeding systems of selected Galápagos Islands angiosperms. Am. J. Bot. 74: 16941705.Google Scholar
McMullen, C. 1989. The Galápagos carpenter bee, just how important is it? Noticias de Galápagos 48: 1618.Google Scholar
McMullen, C. 1990. Reproductive biology of Galápagos Islands angiosperms. In Lawesson, J. E., Hamann, O., Rogers, G., Reck, G., and Ochoa, H. (eds.), Botanical Research and Management in Galápagos (pp. 3545). St. Louis: Missouri Botanical Garden Press (Monogr. Syst. Bot. Missouri Bot. Gard. 32).Google Scholar
McMullen, C. 1993. Flower-visiting insects of the Galápagos Islands. Pan Pacific Entomol. 69: 95106.Google Scholar
McMullen, C. 2007. Pollination biology of the Galápagos endemic, Tournefortia rufo-sericea (Boraginaceae). Bot. J. Linn. Soc. 153: 2131.Google Scholar
McMullen, C. 2009a. Island flora: more than “wretched-looking little weeds.” In De Roy, T. (ed.), Galápagos: Preserving Darwin’s Legacy (pp. 6268). Ontario: Firefly Press.Google Scholar
McMullen, C. 2009b. Pollination biology of a night-flowering Galápagos endemic, Ipomoea habeliana (Convolvulaceae). Bot. J. Linn. Soc. 160: 1120.Google Scholar
McMullen, C., and Close, D. 1993. Wind pollination in the Galápagos Islands. Noticias de Galápagos 52: 1217.Google Scholar
McNeill, J., Barrie, F. R., Buck, W. R., Demoulin, V., Greuter, W., et al. (eds.). 2012. International Code of Nomenclature for Algae, Fungi, and Plants (Melbourne Code). Königstein: Koeltz (Regnum Veg. 154).Google Scholar
McPhaden, M. J., Zebiak, S. E., and Glantz, M. H. 2006. ENSO as an integrating concept in earth science. Science 314: 17401745.Google Scholar
Medina, J. T. 1923. Don Juan Francisco de Sobrecasas, autor de la relación de la isla de Juan Fernández. Rev. Chil. Hist. Geogr. 45: 461473 (publication of the manuscript by J. F. de Sobrecasas “Relacion orthographica, physico-médico, matemática de la Isla de San Juan Baptista, álias de Juan Fernández,. …” based on his trip to the islands in 1751).Google Scholar
Medina, J. T. 1974. El Piloto Juan Fernández, Descubridor de las Islas que Llevan su Nombre, y Juan Jufre, Armador de la Expedición que Hizo en Busca de Otras en el Mar del Sur (2nd edn.). Santiago: Editora Nacional Gabriela Mistral.Google Scholar
Meisner, C. F. 1839. Plantarum Vascularium Genera. Leipzig: Libraria Weidmannia.Google Scholar
Mejías, J. A., and Kim, S.-C. 2012. Taxonomic treatment of Cichorieae (Asteraceae) endemic to the Juan Fernández and Desventuradas Islands (SE Pacific). Ann. Bot. Fenn. 49: 171178.Google Scholar
Meloni, M., Reid, A., Caujapé-Castells, J., Sotok, M., Fernández-Palacios, J. M., et al. 2015. High genetic diversity and population structure in the endangered Canarian endemic Ruta oreojasme (Rutaceae). Genetica 143: 571580.Google Scholar
Mendoza, A., Pérez-García, B., Jaramillo, I. R., and Ricci, M. 1996. Desarrollo del gametofito de Pteris berteroana (Pteridaceae: Pterideae). Rev. Biol. Trop. 44–45: 5157.Google Scholar
Menéndez, C. A. 1966. La presencia de Thyrsopteris en el Cretácico superior de Cerro Guido, Chile. Ameghiniana 4: 299302.Google Scholar
Merilä, J., and Crnokrak, P. 2001. Comparison of genetic differentiation at marker loci and quantitative traits. J. Evol. Biol. 14: 892903.Google Scholar
Meudt, H. M., and Clark, A. C. 2007. Almost forgotten or latest practice? AFLP applications, analyses, and advances. Trends Plant Sci. 12: 106117.Google Scholar
Meyer, F. G. 1966. Chile-United States botanical expedition to Juan Fernandez Islands, 1965. Antarctic J. 1: 238242.Google Scholar
Meza Torres, E. I., Cerne, B., Ulke, A. G., and Morbelli, M. A. 2015b. Distribution of Ophioglossum reticulatum L. in South America. A case of long-distance jump dispersal? Int. J. Biometeor 59: 137150.Google Scholar
Meza Torres, E. I., Macluf, C. C., Morbelli, M. A., and Ferrucci, M. S. 2015a. The circumscription of problematic species of Ophioglossum (Ophioglossaceae) from Southern South America: a palynological approach. Phytotaxa 205: 145156.Google Scholar
Middleton, D. J. 1990. Pernettya or Gaultheria? Plantsman 12: 167177.Google Scholar
Middleton, D. J. 1991. Infrageneric classification of the genus Gaultheria L. (Ericaceae). Bot. J. Linn. Soc. 106: 229258.Google Scholar
Middleton, D. J., and Wilcock, C. C. 1990. A critical examination of the status of Pernettya as a genus distinct from Gaultheria. Edinb. J. Bot. 47: 291301.Google Scholar
Middleton, R. M. 1909. The first Fuegian collection. J. Bot. 47: 207212.Google Scholar
Midway, S. R., and Hodge, A. M. C. 2012. Carlquist revisited: history, success, and applicability of a natural history modelBiol. Philos27: 497520.Google Scholar
Miller, J. M. 1988. A new species of Degeneria (Degeneriaceae) from the Fiji archipelago. J. Arnold Arbor. 69: 275280.Google Scholar
Miller, J. M. 1989. The archaic flowering plant family Degeneriaceae: its bearing on an old enigma. Natl. Geogr. Res. 512: 218231.Google Scholar
Minster, J. B., and Jordan, T. H. 1978. Present-day plate motions. J. Geophys. Res. 83: 53315354.Google Scholar
Mione, T., and Anderson, G.J. 1992. Pollen ovule ratios and breeding system evolution in Solanum section Basarthrum. Am. J. Bot. 79: 279287.Google Scholar
Mitchell, A. D., and Heenan, P.B. 2002. Sophora sect. Edwardsia (Fabaceae): further evidence from nrDNA sequence data of a recent and rapid radiation around the Southern Oceans. Bot. J. Linn. Soc. 140: 435441.Google Scholar
Mitchell, A. L., and House, S. 1999. David Douglas: Explorer ad Botanist. London: Aurum Press.Google Scholar
Monaghan, J. 1973. Chile, Peru, and the California Gold Rush of 1849. Berkeley: University of California Press.Google Scholar
Monteiro, R. F., Mantovani, A., and Forzza, R. C. 2015. Morphological phylogenetic analysis of two early-diverging genera of Bromelioideae (Bromeliaceae). Rodriguesia 66: 505521.Google Scholar
Montenegro, G., Teillier, S., Arce, P., and Poblete, V. 1991. Introduction of plants into the Mediterranean-type climate area of Chile. In Groves, R. H. and di Castri, F. (eds.), Biogeography of Mediterranean Invasions (pp. 103114). Cambridge: Cambridge University Press.Google Scholar
Moody, M. M., and Les, D. H. 2007. Phylogenetic systematics and character evolution in the angiosperm family Haloragaceae. Am. J. Bot. 94: 20052025.Google Scholar
Moore, D. M. 1983. Flora of Tierra del Fuego. Shropshire: Anthony Nelson and Missouri Botanical Garden.Google Scholar
Moore, H. E. Jr. 1969. The genus Juania (Palmae-Arecoideae). Gentes Herb. 10: 385393.Google Scholar
Moragues, E., and Traveset, A. 2005. Effect of Carpobrotus spp. on the pollination success of native plant species of the Balearic Islands. Biol. Conserv. 122: 611619.Google Scholar
Moran, R. C. 1993. Robinson Crusoe’s ferns. Fiddlehead Forum 20: 4447.Google Scholar
Moreira-Muñoz, A. 2011. Plant Geography of Chile. Dordrecht: Springer.Google Scholar
Moreira-Muñoz, A., Elortegui F., S., Hobohm, C., and Pinto da Silva Menezes de Sequeira, M. 2014. Endemism on islands: case studies. In Hobohm, C. (ed.), Endemism in Vascular Plants (pp. 165204). Dordrecht: Springer.Google Scholar
Morel, Ch. P. 1975. Antecedentes para la restauración de las fortificaciones de la Isla Robinson Crusoe, Archipiélago de Juan Fernández. In Orellana R, M. (ed.), Las Islas de Juan Fernández. Historia, Arqueología y Antropología de la Isla Robinson Crusoe (pp. 8395). Santiago: Depto. de Ciéncias Antropológicas y Arqueológicas, Universidad de Chile.Google Scholar
Mort, M. E., Crawford, D. J., Archibald, J. K., O’Leary, T. R., and Santos-Guerra, A. 2010. Plant DNA barcoding: a test using Macaronesian taxa of Tolpis (Asteraceae). Taxon 59: 581587.Google Scholar
Mort, M. E., Crawford, D. J., Kelly, J. K., Santos-Guerra, A., Menezes de Sequeira, M., et al. 2015. Mulitplexed-shotgun-genotyping data resolve phylogeny within a very recently derived insular lineage. Am. J. Bot. 102: 634641.Google Scholar
Morton, C. V. 1968. The genera, subgenera, and sections of the Hymenophyllaceae. Contrib. US Natl. Mus. 38: 153214.Google Scholar
Moseley, H. N. 1892. Notes by a Naturalist: An Account of Observations Made during the Voyage of H.M.S. “Challenger” Round the World in the Years 1872–1876. London: John Murray.Google Scholar
Muasya, A. M., Vrijdaghs, A., Simpson, D. A., Chase, M. W., Goetghebeur, P., et al. 2009. What is a genus in Cypereae: phylogeny, character homology assessment and generic circumscription in Cypereae. Bot. Rev. 75: 5266.Google Scholar
Mueller-Dombois, D., and Fosberg, F. R. 1998. Vegetation of the Tropical Pacific Islands. New York: Springer.Google Scholar
Müller, K., and Borsch, T. 2005. Phylogenetics of Amaranthaceae based on matK/trnK sequence data: evidence from parsimony, likelihood, and Bayesian analyses. 2005. Ann. Missouri Bot. Gard. 92: 66102.Google Scholar
Muñoz Pizarro, C. 1944. El itinerario de Don Claudio Gay. Bol. Mus. Nac. Hist. Nat. 22: 2744.Google Scholar
Muñoz Pizarro, C. 1960. Las Especies de Plantas Descritas por R. A. Philippi en el Siglo XIX. Santiago: Universidad de Chile.Google Scholar
Muñoz Pizarro, C. 1974. El Archipiélago de Juan Fernández y la conservación de sus recursos naturales renovables. Santiago: Museo Nacional de Historia Natural (Serie Educativa No. 9).Google Scholar
Munroe, E. G. 1948. The geographical distribution of butterflies in the West Indies. Ph.D. dissertation, Cornell University, Ithaca, NY.Google Scholar
Munroe, E. G. 1963. Perspectives in biogeography. Can. Entomol. 95: 299308.Google Scholar
Murakami, N., and Schaal, B. A. 1994. Chloroplast DNA variation and the phylogeny of Asplenium sect. Hymenasplenium (Aspleniaceae) in the New World Tropics. J. Plant Res. 107: 245251.Google Scholar
Murillo-A., J., Ruiz, E., Landrum, L. R., Stuessy, T. F., and Barfuss, M. H. J. 2012. Phylogenetic relationships of Myrceugenia (Myrtaceae) based on plastid and nuclear DNA sequences. Mol. Phylogenet. Evol. 62: 764776.Google Scholar
Murillo-A., J., Stuessy, T. F., and Ruiz, E. 2013. Phylogenetic relationships among Myrceugenia, Blepharocalyx, and Luma (Myrtaceae) based on paired-sites models and the secondary structures of ITS and ETS sequences. Plant Syst. Evol. 299: 713729.Google Scholar
Murillo-A., J. Stuessy, T. F., and Ruiz, E. 2016. Explaining disjunct distributions in the flora of southern South America: evolutionary history and biogeography of Myrceugenia (Myrtaceae). J. Biogeogr. 43: 979990.Google Scholar
Murillo-Aldana, J., and Ruiz, E. 2011. Revalidación de Nothomyrcia (Myrtaceae), un género endémico del Archipiélago de Juan Fernández. Gayana Bot. 68: 129134.Google Scholar
Murphy, B. P., Johnson, J. P. L., Gasparini, N. M., and Sklar, L. S. 2016. Chemical weathering as a mechanism for the climatic control of bedrock river incision. Nature 532: 223227.Google Scholar
Murray, D. R. 1986. Seed dispersal by water. In Murray, D. R. (ed.), Seed Dispersal (pp. 4985). Sydney: Academic Press.Google Scholar
Nakai, T. 1950. A new classification of the Gleicheniales. Bull. Nat. Sci. Mus. Tokyo 29: 171.Google Scholar
Narayan, R. 1987. Nuclear DNA changes, genome differentiation and evolution in Nicotiana (Solanaceae). Plant Syst. Evol. 157: 161180.Google Scholar
Natali, A., and Manen, J.-F. 1995. The phylogenetic reconstruction inferred from the sequence of a cpDNA intergene region as a framework for the taxonomic revision of a systematic group: the example of the tribe Rubieae (Rubiaceae). Giorn. Bot. Ital. 129: 5766.Google Scholar
Natland, J. H. 2003. Capture of helium and other volatiles during the growth of olivine phenocrysts in picritic basalts from the Juan Fernandez Islands. J. Petrol. 44: 421456.Google Scholar
Naumann, J., Salomo, K., Der, J. P., Wafula, E. K., Bolin, J. F., et al. 2013. Single-copy nuclear genes place haustorial Hydnoraceae within Piperales and reveal a Cretaceous origin of multiple parasitic angiosperm lineages. PLoS One 8(11):112.Google Scholar
Naumann, J., Symmank, L., Samain, M.-S., Müller, K. F., Neinhuis, C., et al. 2011. Chasing the hare: evaluating the phylogenetic utility of a nuclear single copy gene region at and below species level within the species rich group Peperomia (Piperaceae). BMC Evol. Biol. 11: 357. Available at www.biomedcentral.com/1471–2148/11/357.Google Scholar
Navas B., E. 1961. El género Urtica en Chile. Bol. Soc. Argent. Bot. 9: 395413.Google Scholar
Nazaire, M., and Hufford, L. 2012. A broad phylogenetic analysis of Boraginaceae: implications for the relationships of Mertensia. Syst. Bot. 37: 758783.Google Scholar
Neal, P. R., and Anderson, G. J. 2005. Are ‘mating systems’ ‘breeding systems’ of inconsistent and confusing terminology in plant reproductive biology? Or is it the other way around? Plant Syst. Evol. 250: 173185.Google Scholar
Nee, M. 1996. A new species of Acanthosyris (Santalaceae) from Bolivia and a key to the woody South American Santalaceae. Brittonia 48: 574579.Google Scholar
Nei, M. 1973. Analysis of gene diversity in subdivided populations. Proc. Natl. Acad. Sci. USA 70: 33213323.Google Scholar
Nei, M. 1978. Estimation of average heterozygosity and genetic distance from a small number of individuals. Genetics 89: 583590.Google Scholar
Nei, M., Maruyama, T., and Chakraborty, R. 1975. The bottleneck effect and genetic variability in populations. Evolution 29: 110.Google Scholar
Neshyba, S., and Silva S., N. 1985. Algunos aspectos climatológicos de las islas Robinson Crusoe y Santa Clara (Pacífico Sud Oriental). In Arana, P. (ed.), Investigaciones Marinas en el Archipiélago de Juan Fernández (pp. 4354). Santiago: Editorial Universitaria.Google Scholar
Nesom, G. L. 2008. Classification of subtribe Conyzinae (Asteraceae: Astereae). Lundellia 11: 838.Google Scholar
Newell, R. E., Kidson, J. W., Vincent, D. G., and Boer, G. J. 1972. The General Circulation of the Tropical Atmosphere, vol. 1. Cambridge, MA: MIT Press.Google Scholar
Newman, D., and Pilson, D. 1997. Increased probability of extinction due to decreased genetic effective population size: experimental populations of Clarkia pulchella. Evolution 51: 354362.Google Scholar
Nielsen, L. R., Cowan, R. S., Siegismund, H. R., Adsersen, H., Philipp, M., et al. 2003. Morphometric, AFLP and plastid microsatellite variation in populations of Scalesia divisa and S. incisa (Asteraceae) from the Galápagos Islands. Bot. J. Linn. Soc. 143: 243254.Google Scholar
Nielsen, L. R., Philipp, M., Adsersen, H., and Siegismund, H. R. 2000. Breeding system of Scalesia divisa Andersson – an endemic Asteraceae from the Galápagos Islands. Norske Vidensk.-Akad., Mat.-Naturvidensk. Kl., Skrift. N. Ser. 39: 127138.Google Scholar
Nishida, H., and Nishida, M. 1979. Thyrsopterorachis, gen. nov., a tree fern rachis from the Upper Cretaceous of Hokkaido, Japan. Bot. Mag. Tokyo 92: 187195.Google Scholar
Nishida, H., and Nishida, M. 1981. The vegetation of Mas a Tierra (Robinson Crusoe) Island, Juan Fernandez. Fac. Sci. Chiba Univ. 4148.Google Scholar
Noben, S., and Lehnert, M. 2013. The genus Dicksonia (Dicksoniaceae) in the western Pacific. Phytotaxa 155: 2334.Google Scholar
North Dakota Agricultural Experiment Station (NDAES). 1975. Recommended Chemical Soil Test Procedures for the North Central Region. Fargo: NDAES (Bulletin No. 499).Google Scholar
Nosil, P., Harmon, L. J., and Seehausen, O. 2009. Ecological explanations for (incomplete) speciation. Trends Ecol. Evol. 24: 145156.Google Scholar
Novoa, P. 2015. Expedición Botánica a la Isla Alejandro Selkirk. Viña del Mar: Jardin Botánico Nacional.Google Scholar
Novoa, R., and Villaseca, C. (eds.). 1989. Mapa Agroclimático de Chile. Santiago de Chile: Instituto de Investigaciones Agropecuarias (INIA), Ministerio de Agricultura.Google Scholar
Noyes, R. D. 2000. Biogeographical and evolutionary insights on Erigeron and allies (Asteraceae) from ITS sequence data. Plant Syst. Evol. 220: 93114.Google Scholar
Oakley, C. G. 2015. The influence of natural variation in population size on the ecological and quantitative genetics of the endangered endemic plant Hypericum cumulicola. Int. J. Plant Sci. 176: 1119.Google Scholar
Oehrens, E., and Garrido, N. 1986. Posibilidad del control biológico de la zarzamora en el archipiélago de Juan Fernández. Bol. Soc. Biol. Concepción Chile 57: 205206.Google Scholar
Oldfield, S. 2011. Botanic gardens and the conservation of island floras. In Bramwell, D. and Caujapé-Castells, J. (eds.), The Biology of Island Floras (pp. 474487). Cambridge: Cambridge University Press.Google Scholar
Olesen, J. M., Valido, A., and Dupont, Y. L. 2002. Polinización de las plantas Canarias. El Indiferente 13: 1829.Google Scholar
Oliver, W. R. B. 1935. The genus Coprosma. B.P. Bishop Mus. Bull. 132: 1207.Google Scholar
Ollerton, J., Cranmer, L., Stelzer, R., Sullivan, S., and Chittka, L. 2009. Bird pollination of Canary Island endemic plants. Naturwissenschaften 96: 221232.Google Scholar
Olmstead, R. G., dePamphilis, C. W., Wolfe, A. D., Young, N. D., Elisens, W. J., et al. 2001. Disintegration of the Scrophulariaceae. Am. J. Bot. 88: 348361.Google Scholar
Olmstead, R. G., and Reeves, P. A. 1995. Evidence for the polyphyly of the Scrophulariaceae based on chloroplast rbcL and ndhF sequences. Ann. Missouri Bot. Gard. 82: 176193.Google Scholar
Ono, M. 1975. Chromosome numbers of some endemic species of the Bonin Islands. I. Bot. Mag. (Tokyo) 88: 323328.Google Scholar
Ono, M. 1991. The flora of the Bonin (Ogasawara) Islands: endemism and dispersal modes. Aliso 13: 95105.Google Scholar
Ono, M. 1998. Conservation of the endemic vascular plant species of the Bonin (Ogasawara) Islands. In Stuessy, T. F. and Ono, M. (eds.), Evolution and Speciation in Island Plants (pp. 169180). Cambridge: Cambridge University Press.Google Scholar
Ono, M., and Kobayashi, S. 1985. Flowering plants endemic to the Bonin Islands. In Ono, M. and Okutomi, K. (eds.), Endemic Plant Species and Vegetation of the Bonin (Ogasawara) Islands (pp. 196). Kamakura: Aboc-sha.Google Scholar
Ono, M., and Masuda, Y. 1981. Chromosome numbers of some endemic species of the Bonin Islands, part II. Ogasawara Res. 4: 124.Google Scholar
Orchard, A. E. 1975. Taxonomic revisions in the family Haloragaceae: I. The genera Haloragis, Haloragodendron, Glischrocaryon, Meziella and Gonocarpus. Bull. Auckland Instit. & Mus. 10: 1299.Google Scholar
Orellana R., M. 1975. Historia de los primeros poblamientos de la Isla de Robinson Crusoe. In Orellana R., M. (ed.), Las Islas de Juan Fernandez. Historia, Arqueología y Antropología de la Isla Robinson Crusoe (pp. 1122). Santiago: Dpto. Ciéncias Antropológicas y Arqueológicas, Universidad de Chile.Google Scholar
Ortiz R., A. (ed.). 1982. Estudio de los Recursos Físicos: Archipiélago Juan Fernández. Santiago: Instituto Nacional de Investigación de Recursos Naturales – CORFO.Google Scholar
Otto, E. M., Janßen, T., Kreier, H.-P., and Schneider, H. 2009. New insights into the phylogeny of Pleopeltis and related Neotropical genera (Polypodiaceae, Polypodiopsida). Mol. Phylogenet. Evol. 53: 190201.Google Scholar
Ovalle, A. de. 1649. An Historical Relation of the Kingdom of Chile, of the Islands of the Kingdom of Chile. Rome (translated from Spanish and included in Pinkerton, J. 1813. A General Collection of the Best and Most Interesting Voyages and Travels in All Parts of the World, London, pp. 9091).Google Scholar
Oxelman, B., Kornhall, P., Olmstead, R. G., and Bremer, B. 2005. Further disintegration of Scrophulariaceae. Taxon 54: 411425.Google Scholar
Pacheco, P., Crawford, D. J., Stuessy, T. F., and Silva O., M. 1985. Flavonoid evolution in Robinsonia (Compositae) of the Juan Fernandez Islands. Am. J. Bot. 72: 989998.Google Scholar
Pacheco, P., Crawford, D. J., Stuessy, T. F., and Silva O., M. 1991b. Flavonoid evolution in Dendroseris (Compositae) in the Juan Fernández Islands, Chile. Am. J. Bot. 78: 534543.Google Scholar
Pacheco, P., Crawford, D. J., Stuessy, T. F., and Silva O., M. 1993. Flavonoid chemistry and evolution of Gunnera (Guneraceae) in the Juan Fernández Islands, Chile. Gayana Bot. 50: 1730.Google Scholar
Pacheco, P., Stuessy, T. F., and Crawford, D. J. 1991a. Natural interspecific hybridization in Gunnera (Gunneraceae) of the Juan Fernández Islands, Chile. Pacific Sci. 45: 389399.Google Scholar
Pannell, J. R. 2015. Evolution of the mating system in colonizing plants. Mol. Ecol. 24: 20182037.Google Scholar
Paradis, E. 2006. Analysis of Phylogenetics and Evolution with R. New York: Springer.Google Scholar
Parker de Bassi, M. T. 1987. Tras la Estela del Dresden. Santiago: Ediciones Tusitala.Google Scholar
Parodi, L. R. 1945. Sinopsis de las gramíneas chilenas del género Chusquea. Rev. Univ. (Univ. Católica de Chile) 30(1): 6171.Google Scholar
Parra, M. J., Rodríguez, R., Cavieres, L., Muñoz-Tapia, L., and Atala, C. 2012. Latitudinal patterns in Pteridophyte distribution of continental Chile. Gayana Bot. 72: 5869.Google Scholar
Patiño, J, Carine, M., Fernández-Palacios, J. M., Otto, R., Schaefer, H., et al. 2014. The anagenetic world of spore-producing land plants. New Phytol. 201: 305311.Google Scholar
Peck, S. B. 1990. Eyeless arthropods of the Galapagos Islands, Ecuador: composition and origin of the cryptozoic fauna of a young, tropical, oceanic archipelago. Biotropica 22: 366381.Google Scholar
Peck, S. B., and Kukalova-Peck, J. 1990. Origin and biogeography of the beetles (Coleoptera) of the Galápagos Archipelago, Ecuador. Can. J. Zool. 68: 16171638.Google Scholar
Pedley, L. 1986. Derivation and dispersal of Acacia (Leguminosae), with particular reference to Australia, and the recognition of Senegalia and Racosperma. Bot. J. Linn. Soc. 92: 219284.Google Scholar
Pedley, L. 1987. Racosperma Martius (Leguminosae: Mimosoideae) in Queensland: a checklist. Austrobaileya 2: 344357.Google Scholar
Pelser, P. B., Kennedy, A. H., Tepe, E. J., Shidler, J. B., Nordenstam, B., et al. 2010a. Patterns and causes of incongruence between plastid and nuclear Senecioneae (Asteraceae) phylogenies. Am. J. Bot. 97: 856873.Google Scholar
Pelser, P. B., Nordenstam, B., Kadereit, J. W., and Watson, L. E. 2007. An ITS phylogeny of tribe Senecioneae (Asteraceae) and a new delimitation of Senecio L. Taxon 56: 10771104.Google Scholar
Pelser, P. B., Tepe, E. J., Kennedy, A. H., and Watson, L. E. 2010b. The fate of Robinsonia (Asteraceae): sunk in Senecio, but still monophyletic? Phytotaxa 5: 3146.Google Scholar
Peña, C., Negritto, M. A., Ruiz, E., Baeza, C. M., and Finot, V. L. 2017. Revisión de Megalachne Steud. (Poaceae: Pooideae: Poeae), género endémico del Archipiélago de Juan Fernández, Chile. Gayana Bot. 74: 189199.Google Scholar
Peña, R. C., and Cassels, B. K. 1996. Phylogenetic relationships among Chilean Sophora species. Biochem. Syst. Evol. 24: 725733.Google Scholar
Peña, R. C., Iturriaga, L., Montenegro, G., and Cassels, B. K. 2000. Phylogenetic and biogeographic aspects of Sophora sect. Edwardsia (Papilionaceae). Pacific Sci. 54: 159167.Google Scholar
Peña, R. C., Iturriaga, L., Mujica, A. M., and Montenegro, G. 1993. Análisis micromorfológico de polen de Sophora (Papilionaceae). Hipótesis filogenética sobre el origin de la sección Edwardsia. Gayana Bot. 50: 5765.Google Scholar
Pereira S., E. 1971. Los Primeros Contactos entre Chile y los Estados Unidos 1778–1809. Santiago: Andres Bello.Google Scholar
Pérez, M. E. 2010. Niña de 12 años alertó de la ola gigante en Juan Fernández y evitó una tragedia mayor. La Tercera 1: 1819.Google Scholar
Pérez de Paz, J., and Caujapé-Castells, J. 2013. A review of the allozyme data set for the Canarian endemic flora: causes of the high genetic diversity levels and implications for conservation. Ann. Bot. 111: 10591073.Google Scholar
Perrie, L. R., Bayly, M. J., Lehnebach, C. A., and Brownsey, P. J. 2007. Molecular phylogenetics and molecular dating of the New Zealand Gleicheniaceae. Brittonia 59: 129141.Google Scholar
Perrie, L. R., and Brownsey, P. J. 2005. Insights into the biogeography and polyploid evolution of New Zealand Asplenium from chloroplast DNA sequence data. Am. Fern J. 95: 121.Google Scholar
Perrie, L. R., Brownsey, P. J., Lockhart, P. J., Brown, E. A., and Large, M. F. 2003. Biogeography of temperate Australasian Polystichum ferns as inferred from chloroplast sequence and AFLP. J. Biogeogr. 30: 17291736.Google Scholar
Perrie, L. R., Wilson, R. K., Shepherd, L. D., Ohlsen, D. J., Batty, E. L., et al. 2014. Molecular phylogenetics and generic taxonomy of Blechnaceae ferns. Taxon 63: 745758.Google Scholar
Peterson, B. 1964. Carl Skottsberg 1880–1963. Taxon 13: 17.Google Scholar
Pfosser, M. F., Guzy-Wrobelska, J., Sun, B. Y., Stuessy, T. F., Sugawara, T., et al. 2002. The origin of species of Acer (Sapindaceae) endemic to Ullung Island, Korea. Syst. Bot. 27: 351367.Google Scholar
Pfosser, M., Jakubowsky, G., Schluter, P. M., Fer, T., Kato, H., et al. 2005. Evolution of Dystaenia takesimana (Apiaceae), endemic to Ullung Island, Korea. Plant Syst. Evol. 256: 159170.Google Scholar
Pfrender, M. E., Spitze, K., Hicks, J., Morgan, K., Latta, L., et al. 2000. Lack of concordance between genetic diversity estimates at the molecular and quantitative-trait levels. Conserv. Genet. 1: 263269.Google Scholar
Philbrick, C. T., and Anderson, G. J. 1987. Implications of pollen/ovule ratios and pollen size for the reproductive biology of Potamogeton and autogamy in aquatic angiosperms. Syst. Bot. 12: 98105.Google Scholar
Philippi, F. 1892. El sándalo de Juan Fernández. Anal. Mus. Nac. Chile, Segunda Sección, Bot. 57.Google Scholar
Philippi, R. 1856a. Observaciones sobre la flora de Juan Fernández. Anales Univ. Chile 1856: 157169.Google Scholar
Philippi, R. A. 1856b. Bemerkungen über die Flora der Insel Juan Fernandez. Bot. Zeitung (Berlin) 14: 641650.Google Scholar
Pickard, J. 1983. Vegetation of Lord Howe Island. Cunninghamia 1(2): 133266.Google Scholar
Pilger, R. 1920. Ueber einige Gramineae der Skottsbergschen Sammlung von Juan Fernández. Repert. Spec. Nov. Regni. Veg. 16: 385388.Google Scholar
Porter, D. M. 1976. Geography and dispersal of Galápagos islands vascular plants. Nature 264: 745746.Google Scholar
Porter, D. M. 1983. Vascular plants of the Galapagos: origins and dispersal. In Bowman, R. I., Berson, M., and Leviton, A. E. (eds.), Patterns of Evolution in Galápagos Organisms (pp. 3396). San Francisco: Pacific Division, AAAS.Google Scholar
Potter, D., Eriksson, T., Evans, R. C., Oh, S., Smedmark, J. E. E., et al. 2007. Phylogeny and classification of Rosaceae. Plant Syst. Evol. 266: 543.Google Scholar
Powell, E. A., and Kron, K. A. 2001. An analysis of the phylogenetic relationships in the Wintergreen Group (Gaultheria, Diplycosia, Pernettya, Tepuia; Ericaceae). Syst. Bot. 26: 808817.Google Scholar
Prado, J., Del Nero Rodrigues, C., Salatino, A., and Salatino, M. L. F. 2007. Phylogenetic relationships among Pteridaceae, including Brazilian species, inferred from rbcL sequences. Taxon 56: 355364.Google Scholar
Prince, R. A., Schweller, W. J., Ness, G. E., Coulbourn, W. T., Shepherd, G. L., et al. 1980. Bathymetry of the Peru-Chile Trench and Continental Margin (Geol. Soc. Am. Map and Chart Series MC-26), Pt. I.Google Scholar
Prohens, J., Anderson, G. J., Herraiz, J., Bernardello, G., Santos Guerra, A., et al. 2007. Genetic diversity and conservation of two endangered eggplant relatives (Solanum vespertilio Aiton and S. lidii Sunding) endemic to the Canary Islands. Genet. Res. Crop Evol. 54: 451464.Google Scholar
Pryer, K. M., Smith, A. R., Hunt, J. S., and Dubuisson, J.-Y. 2001. rbcL data reveal two monophyletic groups of filmy ferns (Filicopsida: Hymenophyllaceae). Am. J. Bot. 88: 11181130.Google Scholar
Pugsley, H. W. 1936. Enumeration of the species of Euphrasia L. Sect. Semicalcaratae Benth. J. Bot. (Lond.) 74: 273288.Google Scholar
Puppo, P., Curto, M., Velo-Antón, G., Pérez de Paz, P. L., and Meimberg, H. 2014. The influence of geological history on diversification in insular species: genetic and morphological patterns of Micromeria Benth. (Lamiaceae) in Tenerife (Canary archipelago). J. Biogeogr. 41: 18711882.Google Scholar
Purvis, A., Gittleman, J. L., and Brooks, T. (eds.). 2005. Phylogeny and Conservation. Cambridge: Cambridge University Press.Google Scholar
Pyšek, P., Hulme, P. E., Meyerson, L. A., Smth, G. F., Boatwright, J. S., et al. 2013. Hitting the right target: taxonomic challenges for, and of, plant invasions. AoB Plants 5: plate 042; doi:10.1093/aobpla/plt042Google Scholar
Pyšek, P., Richardson, D. M., Rejmánek, M., Webster, G., Williamson, M., et al. 2004. Alien plants in checklists and floras: towards better communication between taxonomists and ecologists. Taxon 53: 131143.Google Scholar
Qiu, Y.-L., Chase, M. W., Les, D. H., and Parks, C. R. 1993. Molecular phylogenetics of the Magnoliidae: cladistic analyses of nucleotide sequences of the plastid gene rbcL. Ann. Missouri Bot. Gard. 80: 587606.Google Scholar
Quensel, P. 1954. Additional comments on the geology of the Juan Fernandez Islands. In Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 1 (pp. 3787). Uppsala: Almqvist & Wiksells.Google Scholar
Rahmstorf, S. 2002. Ocean circulation and climate during the past 120,000 years. Nature 419: 207214.Google Scholar
Rahn, K. 1984. Plantago sect. Oliganthos in southern South America, a taxonomic revision. Nord. J. Bot. 4: 601627.Google Scholar
Rahn, K. 1996. A phylogenetic study of the Plantaginaceae. Bot. J. Linn. Soc. 120: 145198.Google Scholar
Ramalho, R. S., Quartau, R., Trenhaile, A. S., Mitchell, N. C., Woodroffe, C. D., et al. 2013. Coastal evolution on volcanic oceanic islands: a complex interplay between volcanism, erosion, sedimentation, sea-level change and biogenic production. Earth Sci. Rev. 127: 140170.Google Scholar
Ramsey, J., Bradshaw, H. D., and Schemske, D. W. 2003. Components of reproductive isolation between the monkeyflowers Mimulus lewisii and M. cardinalis (Phrymaceae). Evolution 57: 15201534.Google Scholar
Raven, P. H., Kyhos, D. W., and Cave, M. S. 1971. Chromosome numbers and relationships in Annoniflorae. Taxon 20: 479483.Google Scholar
Reed, D. H., and Frankham, R. 2001. How closely correlated are molecular and quantitative measures of genetic variation? A meta-analysis. Evolution 55: 10951103.Google Scholar
Reiche, C. 1896–1911. Flora de Chile. Santiago: Imprenta Cervantes.Google Scholar
Renner, S. 2014. The relative and absolute frequencies of angiosperm sexual systems: dioecy, monoecy, gynodioecy, and an updated online database. Am. J. Bot. 101: 15881596.Google Scholar
Renner, S., and Ricklefs, R. 1995. Dioecy and its correlates in the flowering plants. Am. J. Bot. 82: 596606.Google Scholar
Retamales, H. A., and Scharaschkin, T. 2015. Comparative leaf anatomy and micromorphology of the Chilean Myrtaceae: taxonomic and ecological implications. Flora 217: 138154.Google Scholar
Reynolds, S. C. P. 2002. A Catalogue of Alien Plants in Ireland. Glasnevin: National Botanic Gardens.Google Scholar
Reznicek, A. A. 1990. Evolution in sedges (Carex, Cyperaceae). Can. J. Bot. 68: 14091432.Google Scholar
Ricci, M. 1996. Variation in distribution and abundance of the endemic flora of Juan Fernández Islands, Chile: Pteridophyta. Biodivers. Conserv. 5: 15211532.Google Scholar
Ricci, M. 2001. Evaluation of conservation status of Lactoris fernandeziana Philippi (Lactoridaceae) in Chile. Biodivers. Conserv. 10: 21292138.Google Scholar
Ricci, M. 2006. Conservation status and ex situ cultivation efforts of endemic flora of the Juan Fernández Archipelago. Biodivers. Conserv. 15: 31113130.Google Scholar
Ricci, M. and Eaton, L. 1994. The rescue of Wahlenbergia larrainii in Robinson Crusoe Island, Chile. Biol. Conserv. 68: 8993.Google Scholar
Richardson, J. E., Fay, M. F., Cronk, Q. C. B., Bowman, D., and Chase, M. W. 2000. A phylogenetic analysis of Rhamnaceae using rbcL and trnL-F plastid DNA sequences. Am. J. Bot. 87: 13091324.Google Scholar
Rick, C. 1966. Some plant-animal relations on the Galápagos Islands. In Bowman, R. I. (ed.), The Galápagos (pp. 215224). Berkeley: University of California Press.Google Scholar
Ridley, H. N. 1930. The Dispersal of Plants Throughout the World. Ashford, Kent: L. Reeve.Google Scholar
Rieseberg, L. H. 1997. Hybrid origins of plant species. Annu. Rev. Ecol. Syst. 28: 359389.Google Scholar
Rieseberg, L. H., and Brouillet, L. 1994. Are many plant species paraphyletic? Taxon 43: 2132.Google Scholar
Rieseberg, L. H., and Willis, J. H. 2007. Plant speciation. Science 317: 910914.Google Scholar
Ritland, K. 2002. Extensions of models for the estimation of mating systems using n independent loci. Heredity 88: 221228.Google Scholar
Roalson, E. H., Columbus, J. T., and Friar, E. A. 2001. Phylogenetic relationships in Cariceae (Cyperaceae) based on ITS (nrDNA) and trnT-L-F (cpDNA) region sequences: assessment of subgeneric and sectional relationships in Carex with emphasis on section Acrocystis. Syst. Bot. 26: 318341.Google Scholar
Rodríguez, R. 1995. Pteridophyta. In Marticorena, C. and Rodríguez, R. (eds.), Flora de Chile, vol. I: Pteridophyta: Gymnospermae (pp. 119309). Concepción, Chile: Universidad de Concepción.Google Scholar
Rodríguez, R. 1990. Comentarios fitogeográficos y taxonómicos de Pteridophyta chilenos. Gayana Bot. 46: 199208.Google Scholar
Rodríguez, R. 2011. Rutaceae. In Marticorena, C. and Rodríguez, R. (eds.), Flora de Chile, vol. 3(1): Misodendraceae-Zygophyllaceae (pp. 109114). Concepción, Chile: Universidad de Concepción.Google Scholar
Rodríguez, R. 2015. Notas taxonómicas sobre Pteridófitos chilenos. Gayana Bot. 72: 94100.Google Scholar
Rodríguez, R., and Ponce, M. 2007. Nuevas combinaciones en Gleicheniaceae. Darwiniana 45: 236241.Google Scholar
Rodríguez-Ríos, R. 1990. Gleichenia lepidota n. sp. y la familia Gleicheniaceae del archipiélago de Juan Fernández, Chile. Gayana Bot. 47: 3745.Google Scholar
Roff, D. A. 1990. The evolution of flightlessness in insects. Ecol. Monogr. 60: 389421.Google Scholar
Rogers, W. 1712. A Cruising Voyage Round the World: First to the South-Seas, Thence to the East-Indies, and Homewards by the Cape of Good Hope. Begun in 1708, and Finish’d in 1711. London.Google Scholar
Roggeveen, J. 1838. Daagverhaal der Ontdekkings-reis von Mr. Jacob Roggeveen in de Jaren 1721 en 1722. Middelburg (in the English translation by Sharp, A. (ed.), The Journal of Jacob Roggeveen. Oxford University Press, 1970).Google Scholar
Romell, L. 1926. Basidiomycetes from Juan Fernandez, collected by C. Skottsberg. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 465471). Uppsala: Almqvist & Wiksells.Google Scholar
Romero A., H. 1985. Geografía de los Climas. In López S., E. (dir.), Geografia de Chile, vol. 11. Santiago: Instituto Geográfico Militar.Google Scholar
Romero García, A. T., Blanca López, G., and Morales Torres, C. 1988. Revisión del género Agrostis L. (Poaceae) en la Peninsula Ibérica. Ruizia 7: 1160.Google Scholar
Rønsted, N., Chase, M. W., Albach, D. C., and Bello, M. A. 2002. Phylogenetic relationships within Plantago (Plantaginaceae): evidence from nuclear ribosomal ITS and plastid trnL-F sequence data. Bot. J. Linn. Soc. 139: 323338.Google Scholar
Rossbach, R. P. 1940. Spergularia in North and South America. Rhodora 42: 5783, 105143, 158193, 203213.Google Scholar
Rossbach, R. P. 1943. El género Spergularia (Caryophyllaceae) en Chile. Darwiniana 6: 211256.Google Scholar
Rothfels, C. J., Sundue, M. A., Kuo, L.-Y., Larsson, A., Kato, M., et al. 2012. A revised family-level classification for eupolypod II ferns (Polypodiidae: Polypodiales). Taxon 61: 515533.Google Scholar
Rothfels, C. J., Windham, M. D., Gruszi, A. L., Gastony, G. J., and Pryer, K. M. 2008. Toward a monophyletic Notholaena (Pteridaceae): resolving patterns of evolutionary convergence in xeric-adapted ferns. Taxon 57: 712724.Google Scholar
Rothfels, C. J., Windham, M. D., and Pryer, K. M. 2013. A plastid phylogeny of the cosmopolitan fern family Cystopteridaceae (Polypodiopsida). Syst. Bot. 38: 295306.Google Scholar
Roy, M. S., Torres-Mura, J. C., and Hertel, F. 1998. Evolution and history of hummingbirds (Aves: Trochilidae) from the Juan Fernandez Islands, Chile. Ibis 140: 265273.Google Scholar
Rúgolo de Agrasar, Z., and De Paula, M. E. 1978. Agrostis L. In Nicora, E. G. (ed.), Flora Patagonica, Parte III: Gramineae (pp. 369394). Buenos Aires: INTA.Google Scholar
Rúgolo de Agrasar, Z. E., and Molina, A. M. 1997. Las especies del género Agrostis L. (Gramineae: Agrostideae) de Chile. Gayana Bot. 54: 91156.Google Scholar
Ruh, M. 1975. Alfredo de Rodt, subdelegado en Juan Fernández 1877–1905. In Orellana R., M. (ed.), Las Islas de Juan Fernández. Historia, Arqueología y Antropología de la Isla Robinson Crusoe (pp. 97136). Santiago: Depto. Ciéncias Antropológicas y Arqueológicas, Universidad de Chile.Google Scholar
Ruiz, E. 2001. Ranunculaceae. In Marticorena, C. and Rodríguez, R. (eds.), Flora de Chile, vol. 2: Winteraceae-Ranunculaceae (pp. 4084). Concepción, Chile: Universidad de Concepción.Google Scholar
Ruiz, E., Becerra, J., Silva O., M., Crawford, D. J., and Stuessy, T. F. 1994. Flavonoid chemistry of the endemic species of Myrceugenia (Myrtaceae) of the Juan Fernández Islands and relatives in continental South America. Brittonia 46: 187193.Google Scholar
Ruiz, E., Crawford, D. J., González, F., Samuel, R., Becerra, J., et al. 2004. Phylogenetic relationships and genetic divergence among endemic species of Berberis, Gunnera, Myrceugenia and Sophora of the Juan Fernández Islands (Chile) and their continental progenitors based on isozymes and nrITS sequences. Taxon 53: 321332.Google Scholar
Ruiz, E., Donoso, C., González, F., Becerra, J., Marticorena, C., et al. 1999. Phenetic relationships between Juan Fernández and continental Chilean species of Sophora (Fabaceae) based on flavonoid patterns. Bol. Chil. Quim. 44: 351356.Google Scholar
Ruiz, E., Marticorena, C., Crawford, D. J., Stuessy, T. F., González, F., et al. 2000. Morphological and ITS sequence divergence between taxa of Cuminia (Lamiaceae), an endemic genus of the Juan Fernández Islands, Chile. Brittonia 52: 341350.Google Scholar
Ruiz, E., Toro, O., Crawford, D. J., Stuessy, T. F., Negritto, M. A., et al. 2008. Phylogenetic relationships among Chilean species of Drimys (Winteraceae) based on ITS sequences and insertion/deletion events. Gayana Bot. 65: 220228.Google Scholar
Rundle, H. D., and Nosil, P. 2005. Ecological speciation. Ecol. Lett. 8: 336352.Google Scholar
Russ, C. H. 1923. The Log of a Forty-Niner [Richard L. Hale]: Being the Record of Adventures by Sea and Shore to the California Gold-Fields and the Pacific Northwest, 1849–1854. Boston, MA: B. J. Brimmer.Google Scholar
Ryves, T. B., Clement, E. J., and Foster, M. C. 1996. Alien Grasses of the British Isles. London: Botanical Society of the British Isles.Google Scholar
Saarela, J. M., Liu, Q., Peterson, P. M., Soreng, R. J., and Paszko, B. 2010. Phylogenetics of the grass “Aveneae-type plastid DNA clade” (Poaceae: Pooideae, Poeae) based on plastid and nuclear ribosomal DNA sequence data. In Seberg, O., Petersen, G., Barfod, A. S., and Davis, J. I. (eds.), Diversity, Phylogeny, and Evolution in the Monocotyledons (pp. 557586). Aarhus, Denmark: Aarhus University Press.Google Scholar
Saarela, J. M., Peterson, P. M., Keane, R. M., Cayouette, J., and Graham, S. W. 2007. Molecular phylogenetics of Bromus (Poaceae: Pooideae) based on chloroplast and nuclear DNA sequence data. Aliso 23: 450467.Google Scholar
Saiz, F., and Ojeda, P. 1988. Oryctolagus cuniculus L. en Juan Fernández. Problema y control. Ann. Mus. Hist. Nat. Valparaíso 19: 9198.Google Scholar
Saiz, G. F., De la Hoz, E., Toro, H., Zuniga, L., Vasquez, E., et al. 1982. Proposición de un Método de Control Integrado del Conejo en el Archipiélago de Juan Fernández. Valparaíso: Universidad Católica de Valparaíso.Google Scholar
Sakai, A., Wagner, W., Ferguson, D., and Herbst, D. 1995. Origins of dioecy in the Hawaiian flora. Ecology 76: 25172529.Google Scholar
Salisbury, E. J. 1964. Carl Johan Fredrik Skottsberg 1880–1963. Biograph. Mem. Fellows R. Soc. 10: 244256.Google Scholar
Samain, M.-S., Mathieu, G., Pino, G., Symmank, L., Cieza, N., et al. 2011. The geophytic Peperomia subgenus Tildenia (Piperaceae) in the Andes with the description of new species in a phylogenetic framework. Plant Ecol. Evol. 144: 148176.Google Scholar
Samain, M.-S., Vanderschaeve, L., Chaerle, P., Goetghebeur, P., Neinhuis, C., et al. 2009. Is morphology telling the truth about the evolution of the species rich genus Peperomia (Piperaceae)? Plant Syst. Evol. 278: 121.Google Scholar
Sampson, F. B. 1995. Pollen morphology of Lactoridaceae: a re-examination. Grana 34: 100107.Google Scholar
Sanders, R., Stuessy, T. F., and Marticorena, C. 1982. Recent changes in the flora of the Juan Fernández Islands, Chile. Taxon 31: 284289.Google Scholar
Sanders, R. W., Stuessy, T. F., Marticorena, C., and Silva O., M. 1987. Phytogeography and evolution of Dendroseris and Robinsonia, tree-Compositae of the Juan Fernández Islands. Opera Bot. 92: 195215.Google Scholar
Sanders, R. T., Stuessy, T. F., and Rodríguez, R. 1983. Chromosome numbers from the flora of the Juan Fernandez Islands. Am. J. Bot. 70: 799810.Google Scholar
Sang, T., Crawford, D. J., Kim, S.-C., and Stuessy, T. F. 1994. Radiation of the endemic genus Dendroseris (Asteraceae) on the Juan Fernández Islands: evidence from sequences of the ITS regions of nuclear ribosomal DNA. Am. J. Bot. 81: 14941501.Google Scholar
Sang, T., Crawford, D. J., Stuessy, T. F., and Silva O., M. 1995. ITS sequences and the phylogeny of the genus Robinsonia (Asteraceae). Syst. Bot. 20: 5564.Google Scholar
Santibañez E., J. 1945. La corriente de Humboldt en las costas de Chile. Liga Marítima de Chile Publ. 28: 137.Google Scholar
Saunders, A., Glen, A., Campbell, K., Atkinson, R., Sawyer, J., et al. 2011. Estudio Sobre la Factibilidad del Manejo de Especies Invasoras en el Archipiélago de Juan Fernández, Chile. Lincoln: Landcare Research.Google Scholar
Saville, D. B. O., and Calder, J. A. 1953. Phylogeny of Carex in the light of parasitism by the smut fungi. Can. J. Bot. 31: 164174.Google Scholar
Sax, D. F., and Gaines, S. D. 2008. Species invasions and extinction: the future of native biodiversity on islands. Proc. Natl. Acad. Sci. USA 105: 1149011497.Google Scholar
Schneider, H., Russell, S. J., Cox, C. J., Bakker, F., Henderson, S., et al. 2004. Chloroplast phylogeny of asplenioid ferns based on rbcL and trnL-F spacer sequences (Polypodiidae, Aspleniaceae) and its implications for biogeography. Syst. Bot. 29: 260274.Google Scholar
Schneider, J., Winterfeld, G., Hoffmann, M. H., and Röser, M. 2011. Duthieeae, a new tribe of grasses (Poaceae) identified among the early diverging lineages of subfamily Pooideae: molecular phylogenetics, morphological delineation, cytogenetics and biogeography. Syst. Biodivers. 9: 2744.Google Scholar
Schoener, T. W. 2010. The MacArthur-Wilson equilibrium model: a chronicle of what it said and how it was tested. In Losos, J. B. and Ricklefs, (eds.), The Theory of Island Biogeography Revisited (pp. 5287). Princeton, NJ: Princeton University Press.Google Scholar
Schönswetter, P., and Tribsch, A. 2005. Vicariance and dispersal in the alpine perennial Bupleurum stellatum L. (Apiaceae). Taxon 54: 725732.Google Scholar
Schopf, T. J. M. 1980. Paleoceanography. Cambridge, MA: Harvard University Press.Google Scholar
Schouten, W. C. 1619. The Relation of a Wonderfull Voiage. Amsterdam.Google Scholar
Schuettpelz, E., and Pryer, K. M. 2007. Fern phylogeny inferred from 400 leptosporangiate species and three plastid genes. Taxon 56: 10371050.Google Scholar
Schulte, K., Barfuss, M. H. J., and Zizka, G. 2009. Phylogeny of Bromelioideae (Bromeliaceae) inferred from nuclear and plastid DNA loci reveals the evolution of the tank habit within the subfamily. Mol. Phylogenet. Evol. 51: 327339.Google Scholar
Schulte, K., Horres, R., and Zizka, G. 2005. Molecular phylogeny of Bromelioideae and its implications on biogeography and the evolution of CAM in the family (Poales, Bromeliaceae). Senckenberg. Biol. 85: 113.Google Scholar
Schulz, O. E. 1903. Monographie der Gattung Cardamine. Engler Bot. Jahrb. 32: 280623.Google Scholar
Schumer, M., Rosenthal, G. G., and Andolfatto, P. 2014. How common is homoploid hybrid speciation? Evolution 68: 15531560.Google Scholar
Schwaar, J. 1979. Feuchtwälder auf Juan Fernandez. Phytocoenologia 6: 514523.Google Scholar
Scott, A. J. 1978. A review of the classification of Chenopodium L. and related genera (Chenopodiaceae). Bot. Jahrb. Syst. 100: 205220.Google Scholar
Seaman, F. C., and Funk, V. A. 1983. Cladistic analysis of complex natural products: developing transformation series from sesquiterpene lactone data. Taxon 32: 127.Google Scholar
Sede, S. M., Dürnhöfer, M. S., and Zapata, F. 2013. Phylogenetics of Escallonia (Escalloniaceae) based on plastid DNA sequence data. Bot. J. Linn. Soc. 173: 442451.Google Scholar
Seigler, D. S. 1998. Plant Secondary Metabolism. Boston, MA: Kluwer.Google Scholar
Seltzer, G. O., Rodbell, D. T., and Wright, H. E. 2003. Late-quaternary paleoclimates of the southern tropical Andes and adjacent regions. Palaeogeogr. Palaeoclim. Palaeoecol. 194: 13 (introduction to symposium, pp. 5338).Google Scholar
Setchell, W. A. 1937. The Codiums of the Juan Fernandez Islands. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 587600). Uppsala: Almqvist & Wiksells.Google Scholar
Seward, A. C. 1910. Fossil Plants: A Text-Book for Students of Botany and Geology. Cambridge: Cambridge University Press.Google Scholar
Shapiro, B. J., Leducq, J.-P., and Mallet, J. 2016. What is speciation? PLoS Genet. 12: e1005860 doi: 10.1371/journal.pgen.1005860.Google Scholar
Sharp, A. (ed.). 1970. The Journal of Jacob Roggeveen. Oxford: Clarendon Press.Google Scholar
Shelvocke, G. 1726. A Voyage Round the World. London (in the edited version by Perrin, W. G. (ed.), London: Cassell, 1928).Google Scholar
Shepherd, L. D., Perrie, L. R., Parris, B. S., and Brownsey, P. J. 2007. A molecular phylogeny for the New Zealand Blechnaceae ferns from analyses of chloroplast trnL-trnF DNA sequences. New Zealand J. Bot. 45: 6780.Google Scholar
Shinohara, W., Nakato, N., Yatabe-Kakugawa, Y., Oka, T., Kim, J. K., et al. 2013. The use of matK in Ophioglossaceae phylogeny and determination of Mankyua chromosome number shed light on chromosome number evolution in Ophioglossaceae. Syst. Bot. 38: 564570.Google Scholar
Sigel, E. M., Windham, M. D., Haufler, C. H., and Pryer, K. M. 2014. Phylogeny, divergence time estimates, and phylogeography of the diploid species of the Polypodium vulgare complex (Polypodiaceae). Syst. Bot. 39: 10421055.Google Scholar
Silva, M. 2016. Guardaparques hallan plantas extintas en archipiélago Juan Fernández. Economía y Negocios. Available online at www.economiaynegocios.cl/noticias (accessed April 3, 2016).Google Scholar
Simberloff, D. 2001. Eradication of island invasives: practical actions and results achieved. Trends Ecol. Evol. 16: 273274.Google Scholar
Simberloff, D., Martin, J.-L., Genovesi, P., Maris, V., Wardle, D. A., et al. 2013. Impacts of biological invasions: what’s what and the way forward. Trends Ecol. Evol. 28: 5866.Google Scholar
Simberloff, D. S., and Wilson, E. O. 1970. Experimental zoogeography of islands. A two-year record of colonization. Ecology 51: 934937.Google Scholar
Simpson, B. B. 1975. Pleistocene changes in the flora of the high tropical Andes. Paleobiology 1: 273294.Google Scholar
Simpson, G. G. 1953. The Major Features of Evolution. New York: Columbia University Press.Google Scholar
Sjöstedt, B. 1975. Revision of the genus Cardamine L. (Cruciferae) in South and Central America. Bot. Notiser 128: 819.Google Scholar
Skottsberg, C. 1910. Juan Fernandez-öarnas sandeltrad. Svensk Bot. Tidskr. 4: 167173.Google Scholar
Skottsberg, C. 1914. Botanische Ergebnisse der Schwedischen Expedition nach Patagonien und dem Feuerlande 1907–1909. IV. Studien über die Vegetation der Juan Fernandez-Inseln. Kungl. Svenska Vetenskapsakad. Handl. 51: 173.Google Scholar
Skottsberg, C. 1918. Til Robinson-ön och Världens ände. Stockholm: Albert Bonniers.Google Scholar
Skottsberg, C. (ed.). 1920–1956. The Natural History of Juan Fernandez and Easter Island, vol. 1: Geography, Geology, Origin of Island Life; vol. 2: Botany. Uppsala: Almqvist & Wiksells.Google Scholar
Skottsberg, C. (ed.). 1921–1940. The Natural History of Juan Fernandez and Easter Island, vol. 3: Zoology. Uppsala: Almqvist & Wiksells.Google Scholar
Skottsberg, C. 1921. The phanerogams of the Juan Fernandez Islands. In Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 2 (pp. 95240). Uppsala: Almqvist & Wiksells.Google Scholar
Skottsberg, C. 1925. Juan Fernandez and Hawaii. A phytogeographical discussion. Bernice P. Bishop Mus. Bull. 16: 147.Google Scholar
Skottsberg, C. 1928. Pollinationsbiologie und Samenverbreitung auf den Juan Fernandez-Inseln. In Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 2: Botany (pp. 503547). Uppsala: Almqvist & Wiksells.Google Scholar
Skottsberg, C. 1934. Le peuplement des îles pacifiques du Chili. In Berland, L. et al. (eds.), Contribution à l’étude du Peuplement Zoologique et Botanique des îles du Pacifique (pp. 271280). Paris: Paul Lechevalier et fils.Google Scholar
Skottsberg, C. 1936. Greigia berteroi Skottsb. and its systematic position. Meddel. Göteb. Botan. Trädg. 11: 209222.Google Scholar
Skottsberg, C. 1941. Marine algal communities of the Juan Fernandez Islands, with remarks on the composition of the flora. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 671696). Uppsala: Almqvist & Wiksells.Google Scholar
Skottsberg, C. 1946. Peperomia berteroana Miq. and P. tristanensis Christoph., an interesting case of disjunction. Acta Horti Gothob. 16: 251288.Google Scholar
Skottsberg, C. 1953a. The vegetation of Juan Fernandez and Easter Islands. In Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 2: Botany, Part 6 (pp. 793960). Uppsala: Almqvist and Wiksells.Google Scholar
Skottsberg, C. 1953b. A supplement to the phanerogams of the Juan Fernandez Islands. In Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 2: Botany, Part 6 (pp. 763792). Uppsala: Almqvist & Wiksells.Google Scholar
Skottsberg, C. 1954. A geographical sketch of the Juan Fernandez Islands. In Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 1 (pp. 89192). Uppsala: Almqvist & Wiksells.Google Scholar
Skottsberg, C. 1956. Derivation of the flora and fauna of Juan Fernandez and Easter Island. In Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 1: Geography, Geology, Origin of Island Life (pp. 193438). Uppsala: Almqvist & Wiksells.Google Scholar
Skottsberg, C. 1958. Über Yunquea Tenzii Skottsb. Ber. Deutsch. Bot. Ges. 71: 4552.Google Scholar
Sleumer, H. 1936. Revision der Gattung Pernettya Gaud. Notiz. Bot. Gard. Mus. Berlin-Dahlem 12: 626655.Google Scholar
Sleumer, H. 1977. Revision der Gattung Azara R. et P. (Flacourtiaceae). Bot. Jahrb. Syst. 98: 151175.Google Scholar
Sleumer, H. 1985. Taxonomy of the genus Pernettya Gaud. (Ericaceae). Bot. Jahrb. Syst. 105: 449480.Google Scholar
Smith, A. R. 1993. Phytogeographic principles and their use in understanding fern relationships. J. Biogeogr. 20: 255264.Google Scholar
Smith, A. R., and Moran, R. C. 1987. New combinations in Megalastrum (Dryopteridaceae). Am. Fern J. 77: 124130.Google Scholar
Smith, A. R., Pryer, K. M., Schuettpelz, E., Korall, P., Schneider, H., et al. 2006. A classification for extant ferns. Taxon 55: 705731.Google Scholar
Smith, A. R., and Tejero-Díez, J. D. 2014. Pleopeltis (Polypodiaceae), a redefinition of the genus and nomenclatural novelties. Bot. Sci. 92: 4358.Google Scholar
Smith, C. W. 1989. Non-native plants. In Stone, C. P. and Stone, D. B. (eds.), Conservation Biology in Hawaii (pp. 6069). Honolulu: University of Hawaii Cooperative National Park Resources Studies Unit.Google Scholar
Smith, L. B., and Downs, R. J. 1979. Bromelioideae (Bromeliaceae). Fl. Neotrop. Monogr. 14: 14932142.Google Scholar
Smith, R. J., Muir, R. D. J., Walpole, M. J., Balmford, A., and Leader-Williams, N. 2003. Governance and the loss of biodiversity. Nature 426: 6770.Google Scholar
Smith-Ramírez, C., and Arellano-Cataldo, G. 2013. Necesidad de que las Fuerzas Armadas realicen un control de las plantas invasoras en el Archipiélago de Juan Fernández, Chile. Bosque (Valdivia) 34 (1): 36.Google Scholar
Smith-Ramírez, C., Arellano, G., Hagen, E., Vargas, R., Castillo, J., et al. 2013. El rol de Turdus falcklandii (Aves: Passeriforme) como dispersor de plantas invasoras en el archipiélago de Juan FernándezRev. Chil. Hist. Nat86: 3348.Google Scholar
Sobel, J. M. 2014. Ecogeographic isolation and speciation in the genus Mimulus. Am. Nat. 184: 565579.Google Scholar
Sobel, J. M., and Chen, G. F. 2014. Unification of methods for estimating the strength of reproductive isolation. Evolution 68: 15111523.Google Scholar
Sobel, J. M., Chen, G. F., Watt, L. R., and Schemske, D. W. 2010. The biology of speciation. Evolution 64: 295315.Google Scholar
Sobel, J. M., and Streisfeld, M. A. 2015. Strong premating reproductive isolation drives incipient speciation in Mimulus aurantiacus. Evolution 69: 447461.Google Scholar
Solbrig, O. T. 1962. The South American species of Erigeron. Contrib. Gray Herb. 191: 379.Google Scholar
Solbrig, O. T. 1971. The population biology of dandelions. Am. Sci. 59: 686694.Google Scholar
Solbrig, O. T., Anderson, L. C., Kyhos, D. W., and Raven, P. H. 1969. Chromosome numbers in Compositae VII: Astereae III. Am. J. Bot. 56: 348353.Google Scholar
Soltis, D. E., Albert, V. A., Leebens-Mack, J., Bell, C. D., Paterson, A., et al. 2009. Polyploidy and angiosperm diversification. Am. J. Bot. 96: 336348.Google Scholar
Soltis, D. E., Gitzendanner, M. A., Stull, G., Chester, M., Chanderbali, A., et al. 2013. The potential of genomics in plant systematics. Taxon 62: 886898.Google Scholar
Soltis, D. E., Smith, S. A., Cellinese, N., Wurdack, K. J., Tank, D. C., et al. 2011. Angiosperm phylogeny: 17 genes, 640 taxa. Am. J. Bot. 98: 704730.Google Scholar
Soltis, D. E., Soltis, P. S., Endress, P. K., and Chase, M. W. 2005. Phylogeny and Evolution of Angiosperms. Sunderland, MA: Sinauer Associates.Google Scholar
Soltis, D. E., Soltis, P. S., Nikrent, D. L., Johnson, L. A., Hahn, W. J., et al. 1997. Angiosperm phylogeny inferred from 18S ribosomal DNA sequences. Ann. Missouri Bot. Gard. 84: 149.Google Scholar
Soltis, D. E., Visger, C. J., and Soltis, P. S. 2014. The polyploidy revolution then … and now: Stebbins revisited. Am. J. Bot. 101: 10571078.Google Scholar
Soraru, S. B. 1972. Revisión de las Urticaceae argentinas. Darwiniana 17: 246325.Google Scholar
Soreng, R. J., Davidse, G., Peterson, P. M., Zuloaga, F. O., Judziewicz, E. J., et al. 2000. Catalogue of New World grasses (Poaceae). Available at www.tropicos.org/project/cnwg (accessed November 29, 2015).Google Scholar
Soza, V. L., and Olmstead, R. G. 2010. Evolution of breeding systems and fruits in New World Galium and relatives (Rubiaceae). Am. J. Bot. 97: 16301646.Google Scholar
Spalik, K., Piwczyński, M., Danderson, C. A., Kurzyna-Młynik, R., Bone, T. S., et al. 2010. Amphitropic amphiantarctic disjunctions in Apiaceae subfamily Apioideae. J. Biogeogr. 37: 19771994.Google Scholar
Spooner, D., Stuessy, T. F., Crawford, D. J., and Silva O., M. 1987. Chromosome numbers from the flora of the Juan Fernandez Islands. II. Rhodora 89: 351356.Google Scholar
Spooner, D. M., Anderson, G. J., and Jansen, R. K. 1993. Chloroplast DNA evidence for the interrelationships of tomatoes, potatoes, and pepinos (Solanaceae). Am. J. Bot. 80: 676688.Google Scholar
Spooner, D. M., Douches, D. S., and Contreras M., A. 1992. Allozyme variation within Solanum sect. Petota, ser. Etuberosa (Solanaceae). Am. J. Bot. 79: 467471.Google Scholar
Spooner, D. M., Tivang, J., Nienhuis, J., Miller, J. T., Douches, D. S., et al. 1996. Comparison of four molecular markers in measuring relationships among the wild potato relatives Solanum section Etuberosum (subgenus Potatoe). Theor. Appl. Genet. 92: 532540.Google Scholar
Spry, W. J. J. 1877. The Cruise of Her Majesty’s Ship “Challenger”: Voyages over Many Seas, Scenes in Many Lands. New York: Harper & Brothers.Google Scholar
St. John, H., and Takeuchi, W. 1988. Enlargement of O‛ahu Cyrtandra (Gesneriaceae). Hawaiian plant studies 48. Phytologia 65: 207215.Google Scholar
Stafleu, F. A., and Cowan, R. S. 1976. Taxonomic Literature, vol. I: A-G. Utrecht: Bohn, Scheltema & Holkema.Google Scholar
Stafleu, F. A., Demoulin, V., Greuter, W., Hiepko, P., Linczevski, I. A., et al. 1978. International Code of Botanical Nomenclature, Adopted by the Twelfth International Botanical Congress, Leningrad, July 1975. Utrecht: Bohn, Scheltema & Holkema (Regnum Veg. 97).Google Scholar
Staples, G. W., and Cowie, R. H. (eds.). 2001. Hawaii’s Invasive Species: A Guide to the Invasive Plants and Animals in the Hawaiian Islands. Honolulu: Mutual Publishing and Bishop Museum Press.Google Scholar
Starr, J. R., and Ford, B. A. 2009. Phylogeny and evolution in Cariceae (Cyperaceae): current knowledge and future directions. Bot. Rev. 75: 110137.Google Scholar
Starr, J. R., Harris, S. A., and Simpson, D. A. 2004. Phylogeny of the unispicate taxa in Cyperaceae tribe Cariceae I: generic relationships and evolutionary scenarios. Syst. Bot. 29: 528544.Google Scholar
Starr, J. R., Harris, S. A., and Simpson, D. A. 2008. Phylogeny of the unispicate taxa in Cyperaceae tribe Cariceae II: the limits of Uncinia. In Ford, B. A. and Naczi, R. C. F. (eds.), Sedges: Uses, Diversity, and Systematics of the Cyperaceae (pp. 243262). St. Louis, MO: Missouri Botanical Garden Press.Google Scholar
Stathos, A., and Fishman, L. 2014. Chromosomal rearrangements directly cause underdominant F1 pollen sterility in Mimulus lewisii-Mimulus cardinalis hybrids. Evolution 68: 31093119.Google Scholar
Stearns, H. T. 1966. Geology of the State of Hawaii. Palo Alto, CA: Pacific Books.Google Scholar
Stebbins, G. L. Jr. 1950. Variation and Evolution in Plants. New York: Columbia University Press.Google Scholar
Stebbins, G. L. 1957. Self-fertilization and population variability in the higher plants. Am. Nat. 41: 337354.Google Scholar
Stebbins, G. L. 1958. The inviability, weakness, and sterility of intespecific hybrids. Adv. Genet. 9: 147215.Google Scholar
Stebbins, G. L. 1971. Chromosomal Evolution in Higher Plants. London: Edward Arnold.Google Scholar
Stebbins, G. L., Jenkins, J. A., and Walters, M. S. 1953. Chromosomes and phylogeny in the Compositae, tribe Cichorieae. Univ. Calif. Publ. Bot. 26: 401430.Google Scholar
Steele, P. R., and Pires, J. C. 2011. Biodiversity assessment: state-of-the-art techniques in phylogenomics and species identification. Am. J. Bot. 98: 415425.Google Scholar
Steinbauer, M. J., et al. 2016. Topography-driven isolation, speciation and a global increase of endemism with elevation. Global Ecol. Biogeogr. 25: 10971107.Google Scholar
Stevens, P. F. 1971. A classification of the Ericaceae: subfamilies and tribes. Bot. J. Linn. Soc. 64: 153.Google Scholar
Stiefkens, L., Bernardello, G., and Anderson, G. J. 2001. The somatic chromosomes of Sophora fernandeziana (Fabaceae), an endemic tree from Robinson Crusoe Island. Pacific Sci. 55: 7175.Google Scholar
Stone, C. P., Smith, C. W., and Tunison, J. T. (eds.). 1992. Alien Plant Invasions in Native Ecosystems of Hawaii: Management and Research. Honolulu: University of Hawaii Cooperative National Park Resource Studies Unit.Google Scholar
Ström, K. M. 1921. Freshwater algae from Juan Fernandez and Easter Island. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany (pp. 8593). Uppsala: Almqvist & Wiksells.Google Scholar
Strong, M. T. 1997. Machaerina (Cyperaceae) in South America. Novon 7: 308319.Google Scholar
Strother, J. L. 1976. Chromosome studies in Compositae. Am. J. Bot. 63: 247250.Google Scholar
Stuessy, T. F. 1995. Fernandezian region: CPD site PO8. Juan Fernandez Islands, Chile. In Davis, S. D., Heywood, V. H., and Hamilton, A. C. (eds.), Centres of Plant Diversity, vol. 2 (pp. 565568). Cambridge: WWF & IUCN.Google Scholar
Stuessy, T. F. 2007. Evolution of specific and genetic diversity during ontogeny of island floras: the importance of understanding process for interpreting island biogeographic patterns. In Ebach, M. C. and Tangney, R. S. (eds.), Biogeography in a Changing World (pp. 117133). Boca Raton, FL: CRC Press.Google Scholar
Stuessy, T. F. 2009. Plant Taxonomy: The Systematic Evaluation of Comparative Data (2nd edn.). New York: Columbia University Press.Google Scholar
Stuessy, T. F. 2010. Paraphyly and the origin and classification of angiosperms. Taxon 59: 689693.Google Scholar
Stuessy, T. F., and Crawford, D. J. 1983. Flavonoids and phylogenetic reconstruction. Plant Syst. Evol. 143: 83107.Google Scholar
Stuessy, T. F., and Crawford, D. J. 1998. Chromosomal stasis during speciation in angiosperms of oceanic islands. In Stuessy, T. F. and Ono, M. (eds.), Evolution and Speciation of Island Plants (pp. 307324). Cambridge: Cambridge University Press.Google Scholar
Stuessy, T. F., Crawford, D. J., Anderson, G. J., and Jensen, R. J. 1998c. Systematics, biogeography, and conservation of Lactoridaceae. Perspect. Plant Ecol. Evol. Syst. 1: 267290.Google Scholar
Stuessy, T. F., Crawford, D. J., and Marticorena, C. 1990. Patterns of phylogeny in the endemic vascular flora of the Juan Fernández Islands, Chile. Syst. Bot. 15: 338346.Google Scholar
Stuessy, T. F., Crawford, D. J., Marticorena, C., and Rodríguez, R. 1998a. Island biogeography of angiosperms of the Juan Fernández archipelago. In Stuessy, T. F. and Ono, M. (eds.), Evolution and Speciation of Island Plants (pp. 121138). Cambridge: Cambridge University Press.Google Scholar
Stuessy, T. F., Crawford, D. J., Marticorena, C., and Silva O., M. 1998e. Isolating mechanisms and modes of speciation in endemic angiosperms of Juan Fernández Islands. In Stuessy, T. F. and Ono, M. (eds.) Evolution and Speciation of Island Plants (pp. 7996). Cambridge: Cambridge University Press.Google Scholar
Stuessy, T. F., Crawford, D. J., Soltis, D. E., and Soltis, P. S. 2014a. Plant Systematics: The Origin, Interpretation, and Ordering of Plant Biodiversity. Königstein: Koeltz Scientific Books.Google Scholar
Stuessy, T. F., Foland, K. A., Sutter, J. F., Sanders, R. W., and Silva O., M. 1984. Botanical and geological significance of potassium-argon dates from the Juan Fernandez Islands. Science 255: 4951.Google Scholar
Stuessy, T. F., Greimler, J., and Dirnböck, T. 2005a. Landscape modification and impact on specific and genetic diversity in oceanic islands. In Balslev, H. and Friis, I. (eds.), Plant Diversity and Complexity Patterns: Local, Regional and Global Dimensions (pp. 89101). Copenhagen: Danish Academy of Sciences (Biol. Skr. 55).Google Scholar
Stuessy, T. F., Jakubowsky, G., Gómez, R. S., Pfosser, M., Schluter, P. M., et al. 2006. Anagenetic evolution in island plants. J. Biogeogr. 33: 12591265.Google Scholar
Stuessy, T. F., König, C., and López Sepúlveda, P. 2014b. Paraphyly and endemic genera of oceanic islands: implications for conservation. Ann. Missouri Bot. Gard. 100: 5078.Google Scholar
Stuessy, T. F., Marticorena, C., Rodríguez, R., Crawford, D. J., and Silva O., M. 1992. Endemism in the vascular flora of the Juan Fernández Islands. Aliso 13: 297307.Google Scholar
Stuessy, T. F., and Ono, M. (eds.). 1998. Evolution and Speciation of Island Plants. Cambridge: Cambridge University Press.Google Scholar
Stuessy, T. F., Ruiz, E., Crawford, D. J., and Tremetsberger, K. 2005b. Testing degrees of genetic divergence and populational variation in oceanic island archipelagos: Juan Fernández as a model system. Nova Acta Leopoldina 92: 147165.Google Scholar
Stuessy, T. F., Sanders, R. W., and Matthei, O. R. 1983. Juania australis revisited in the Juan Fernández Islands, Chile. Principes 27: 7174.Google Scholar
Stuessy, T. F., Sanders, R. W., and Silva O., M. 1984. Phytogeography and evolution of the flora of the Juan Fernández Islands: a progress report. In Radovsky, F. J., Raven, P. H., and Sohmer, S. H. (eds.), Biogeography of the Tropical Pacific (pp. 5569). Lawrence, KS: Association of Systematics Collections and the Bishop Museum.Google Scholar
Stuessy, T. F., Sang, T., and DeVore, M. 1996. Phylogeny and biogeography of the subfamily Barnadesioideae with implications for early evolution of the Compositae. In Hind, D. J. and Beentje, H. J. (eds.), Proceeding of the International Compositae Conference (1994), vol. 1: Compositae: Systematics (pp. 463490). Kew: Royal Botanic Gardens.Google Scholar
Stuessy, T. F., Swenson, U., Crawford, D. J., Anderson, G., and Silva O., M. 1998d. Plant conservation in the Juan Fernández archipelago, Chile. Aliso 16: 89101.Google Scholar
Stuessy, T. F., Swenson, U., Marticorena, C., Matthei, O., and Crawford, D. J. 1998b. Loss of plant diversity and extinction on Robinson Crusoe Islands, Chile. In Peng, C.-I. and Lowry, P. P. II (eds.), Rare, Threatened, and Endangered Floras of Asia and the Pacific Rim (pp. 243257). Taipei: Institute of Botany, Academica Sinica (Monograph Series No. 16).Google Scholar
Stuessy, T. F., Takayama, K., and López-Sepúlveda, P. 2012. Founder effects are invisible in endemic species of oceanic islands. J. Biogeog. 39: 15651566.Google Scholar
Stuessy, T. F., Takayama, K., López-Sepúlveda, P., and Crawford, D. J. 2014c. Interpretation of patterns of genetic variation in endemic plant species of oceanic islands. Bot. J. Linn. Soc. 174: 276288.Google Scholar
Sugiura, S. 2008. Male territorial behaviour of the endemic large carpenter bee, Xylocopa (Koptortosoma) ogasawarensis (Hymenoptera: Apidae), on the oceanic Ogasawara Islands. Eur. J. Entomol. 105: 153157.Google Scholar
Sukhorukov, A. P., and Zhang, M. 2013. Fruit and seed anatomy of Chenopodium and related genera (Chenopodioideae, Chenopodiaceae/Amarathaceae): implications for evolution and taxonomy. PLoS One 8(4): e61906.Google Scholar
Sun, B.-Y., and Stuessy, T. F. 1998. Preliminary observations on the evolution of endemic angiosperms of Ullung Island, Korea. In Stuessy, T. F. and Ono, M. (eds.), Evolution and Speciation of Island Plants (pp. 181202). Cambridge: Cambridge University Press.Google Scholar
Sun, B.-Y., Stuessy, T., and Crawford, D. 1990. Chromosome counts from the flora of the Juan Fernández Islands, Chile. III. Pacific Sci. 44: 258264.Google Scholar
Sun, B.-Y., Stuessy, T. F., Humaña, A. M., Riveros, M., and Crawford, D. J. 1996. Evolution of Rhaphithamnus venustus (Verbenaceae), a gynodioecious hummingbird-pollinated endemic of the Juan Fernández Islands, Chile. Pacific Sci. 50: 5565.Google Scholar
Sun, M., and Ganders, F. R. 1988. Mixed mating sytems in Hawaiian Bidens. Evolution 42: 516527.Google Scholar
Sundue, M., Rouhan, G., and Moran, R. 2010. Megalastrum (Dryopteridaceae) of the Circumaustral region: Chile, Argentina, and Southern Islands of the Atlantic, Pacific, and Indian Oceans. Syst. Bot. 35: 461475.Google Scholar
Sungkaew, S., Stapleton, C. M. A., Salamin, N., and Hodkinson, T. R. 2009. Non-monophyly of the woody bamboos (Bambuseae; Poaceae): a multi-gene region phylogenetic analysis of Bambusoideae s.s. J. Plant Res. 122: 95108.Google Scholar
Susanna, A., Galbany-Casals, M., Romaschenko, K., Barres, L., Martín, J., et al. 2011. Lessons from Plectocephalus (Compositae, Cardueae-Centaureinae): ITS disorientation in annuals and Beringian dispersal as revealed by molecular analyses. Ann. Bot. 108: 263277.Google Scholar
Susanna, A., and Garcia-Jacas, N. 2001. Tribe Cardueae. In Anderberg, A. et al. (eds.), Compositae. In Kadereit, J. W. and Jeffrey, C. (eds.), Flowering Plants, vol. VIII: Eudicots, Asterales. Compositae (pp. 123147). In Kubitzki, K. (series ed.), The Families and Genera of Vascular Plants. Berlin: Springer.Google Scholar
Sutcliffe, T. [“T.S.”]. 1839. The Earthquake of Juan Fernandez, as It Occurred in the Year 1835, Authenticated by the Retired Governor of that Island, to Which is Added a Refutation of Several Misstatements that Have Been Published in the “Nautical Magazine” of 1837, and the Public Papers. Manchester.Google Scholar
Sutcliffe, T. 1841. Sixteen Years in Chile and Peru from 1822 to 1839. London: Fisher (on p. x, end of the Preface, June 1,1841, is given, but the book may have appeared later).Google Scholar
Swain, T. (ed.). 1963. Chemical Plant Taxonomy. New York: Academic Press.Google Scholar
Swenson, U., Stuessy, T. F., Baeza, M., and Crawford, D. J. 1997. New and historical plant introductions, and potential pests, in the Juan Fernandez Islands, Chile. Pacific Sci. 51: 233253.Google Scholar
Swenson, U., Ulfsson, V., and Havran, C. 2016. Hawaiiöarna: en flora på fallrepet. Svensk Bot. Tidskr. 110: 80107.Google Scholar
Symmank, L., Samain, M.-S., Smith, J. F., Pino, G., Stoll, A., et al. 2011. The extraordinary journey of Peperomia subgenus Tildenia (Piperaceae): insights into diversification and colonization patterns from its cradle in Peru to the Trans-Mexican Volcanic Belt. J. Biogeogr. 38: 23372349.Google Scholar
Takayama, K., Crawford, D. J., López-Sepúlveda, P., Greimler, J., and Stuessy, T. F. Submitted. Factors driving adaptive radiation in plants of oceanic islands: a case study from the Juan Fernández Archipelago. J. Plant Res.Google Scholar
Takayama, K., López, P., König, C., Kohl, G., Novak, J., et al. 2011. A simple and cost-effective approach for microsatellite isolation in non-model plant species using small-scale 454 pyrosequencing. Taxon 60: 14421449.Google Scholar
Takayama, K., López-Sepúlveda, P., Greimler, J., Crawford, D. J., Peñailillo, P., et al. 2015a. Relationships and genetic consequences of contrasting modes of speciation among endemic species of Robinsonia (Asteraceae, Senecioneae) of the Juan Fernández Archipelago, Chile, based on AFLPs and SSRs. New Phytol. 205: 415428.Google Scholar
Takayama, K., López-Sepúlveda, P., Greimler, J., Crawford, D. J., Peñailillo, P., et al. 2015b. Genetic consequences of cladogenetic vs. anagenetic speciation in endemic plants of oceanic islands. AoB Plants 7: plv102. doi: 10.1093/aobpla/plv102Google Scholar
Takayama, K., López-Sepúlveda, P., Kohl, G., Novak, J., and Stuessy, T. F. 2012a. Development of microsatellite markers in species of Erigeron (Asteraceae) endemic to the Juan Fernández Archipelago, Chile. Am. J. Bot. Primer Notes e1e3. Doi: 10.3732/ajb.1200218.Google Scholar
Takayama, K., Sun, B.-Y., and Stuessy, T. F. 2012b. Genetic consequences of anagenetic speciation in Acer okamotoanum (Sapindaceae) on Ullung Island, Korea. Ann. Bot. 109: 321330.Google Scholar
Takayama, K., Sun, B.-Y., and Stuessy, T. 2013. Anagenetic speciation in Ullung Island, Korea: genetic diversity and structure in the island endemic species, Acer takesimense (Sapindaceae). J. Plant Res. 126: 323333.Google Scholar
Takhtajan, A. 1969. Flowering Plants: Origin and Dispersal. Edinburgh: Oliver & Boyd.Google Scholar
Takhtajan, A. 1986. Floristic Regions of the World. Berkeley: University of California Press.Google Scholar
Tank, D. C., Beardsley, P. M., Kelchner, S. A., and Olmstead, R. G. 2006. Review of the systematics of Scrophulariaceae s.l. and their current disposition. Austral. Syst. Bot. 19: 289307.Google Scholar
Tatemichi, Y. (ed.). 1990. Illustrated Book of the Genus Nicotiana. Toyoda: Japan Tobacco.Google Scholar
Tay, M. L., Meudt, H. M., Garnock-Jones, P. J., and Ritchie, P. A. 2010. DNA sequences from three genomes reveal multiple long-distance dispersals and non-monophyly of sections in Australasian Plantago (Plantaginaceae). Austral. Syst. Bot. 23: 4768.Google Scholar
Taylor, C. 2003. Urticaceae. In Marticorena, C. and Rodríguez, R. (eds.), Flora de Chile, vol. 3(1): Misodendraceae-Zygophyllaceae (pp. 4461). Concepción, Chile: Universidad de Concepción.Google Scholar
Taylor, C. M., and Muñoz-Schick, M. 1994. The botanical works of Philippi, father and son, in Chile. Ann. Missouri Bot. Gard. 81: 743748.Google Scholar
Taylor, R. L., and Mulligan, G. A. 1968. Flora of the Queen Charlotte Islands, part 1: Systematics of the Vascular Plants; part 2: Cytological Aspects of the Vascular Plants. Ottawa: Canada Department of Agriculture (Research Branch Monograph no. 4).Google Scholar
Taylor, S., and Kumar, L. 2014. Climate change and weed impacts on small island ecosystems: Lantana camara L. (Magnoliopsida: Verbenaceae) distribution in Fiji. Pacific Sci. 68: 117133.Google Scholar
Teillier, S., and Escobar, F. 2013. Revisión del género Gaultheria L. (Ericaceae) en Chile. Gayana Bot. 70: 136153.Google Scholar
Templeton, A. R. 1981. Mechanisms of speciation: a population genetic approach. Annu. Rev. Ecol. Evol. 12: 2348.Google Scholar
Templeton, A. R. 1982. Genetic architecture of speciation. In Barigozzi, C. (ed.), Mechanisms of Speciation (pp. 105121). New York: Alan R. Liss.Google Scholar
Templeton, A. R. 2008. The reality and importance of founder speciation in evolution. BioEssays 30: 470479.Google Scholar
Těšitel, J., Řiha, P., Svobodová, Š., Malinová, T., and Štech, M. 2010. Phylogeny, life history evolution and biogeography of the rhinanthoid Orobanchaceae. Folia Geobot. 45: 347367.Google Scholar
The Marie Currie SPECIATION Network. 2012. What do we need to know about speciation? Trends Ecol. Evol. 27: 2739.Google Scholar
Thompson, P. M., and Ollason, J. C. 2001. Lagged effects of ocean climate change on fulmar population dynamics. Nature 413: 417420.Google Scholar
Thorne, R. F. 1963. Biotic distribution patterns in the tropical Pacific. In Gressitt, J. L. (ed.), Pacific Basin Biogeography (pp. 311354). Honolulu: Bishop Museum Press.Google Scholar
Thornton, I. 2007. Island Colonization: The Origin and Development of Island Communities, ed. New, T.. Cambridge: Cambridge University Press.Google Scholar
Tichy, L. 2002. JUICE, software for vegetation classification. J. Veg. Sci. 13: 451453.Google Scholar
Tiffin, P., and Ross-Ibarra, J. 2014. Advances and limits of using population genetics to understand local adaptation. Trends Ecol. Evol. 29: 673680.Google Scholar
Tobe, H., Stuessy, T. F., Raven, P. H., and Oginuma, K. 1993. Embryology and karyomorphology of Lactoridaceae. Am. J. Bot. 80: 933946.Google Scholar
Toft, C. A., and Schoener, T. W. 1983. Abundance and diversity of orb spiders on 106 Bahamian Islands: biogeography at an intermediate trophic level. Oikos 41: 411426.Google Scholar
Tomb, A., Chambers, K., Kyhos, D., Powell, M., and Raven, P. 1978. Chromosome numbers in the Compositae: XIV. Lactuceae. Am. J. Bot. 65: 717721.Google Scholar
Tomlinson, P. B. 1969. The anatomy of the vegetative organs of Juania australis (Palmae). Gentes Herb. 10: 412424.Google Scholar
Torres, D. 1987. Antecedentes sobre el lobo fino de Juan Fernández Arctocephalus philippii y proyecciones para su estudio. In Castilla, J. C. (ed.), Islas Oceánicas Chilenas: Conocimiento Científico y Necesidades de Investigaciones (pp. 289317). Santiago: Universidad Católica de Chile.Google Scholar
Tortosa, R. D. 1989. El género Colletia (Rhamnaceae). Parodiana 5: 279332.Google Scholar
Traveset, A., and Richardson, D. 2006. Biological invasions as disruptors of plant reproductive mutualisms. Trends Ecol. Evol. 21: 208216.Google Scholar
Trénel, P., Gustafsson, M. H. G., Baker, W. J., Asmussen-Lange, C. B., Dransfield, J., et al. 2007. Mid-Tertiary dispersal, not Gondwanan vicariance explains distribution patterns in the wax palm subfamily (Ceroxyloideae: Arecaceae). Mol. Phylogenet. Evol. 45: 272288.Google Scholar
Trusty, J. L., Kesler, H. C., Rodríguez, J., and Francisco-Ortega, J. 2011. Conservation status of endemic plants on Isla del Coco, Costa Rica: applying IUCN Red List criteria on a small island. In Bramwell, D. and Caujapé-Castells, J. (eds.), The Biology of Island Floras (pp. 452473). Cambridge: Cambridge University Press.Google Scholar
Trusty, J. L., Olmstead, R. G., Bogler, D. J., Santos-Guerra, A., and Francisco-Ortega, J. 2004. Using molecular data to test a biogeographic connection of the Macaronesian genus Bystropogon (Lamiaceae) to the New World: A case of conflicting phylogenies. Syst. Bot. 29: 702715.Google Scholar
Tryon, R. 1971. Development and evolution of fern floras of oceanic islands. Biotropica 2: 7684 (also published as pp. 5462 in Stern, W. L. (ed.), Adaptive Aspects of Insular Evolution. Pullman, WA: Washington State University Press, 1971).Google Scholar
Tryon, R. M. 1956. A revision of the American species of Notholaena. Contr. Gray Herb. 179: 1106.Google Scholar
Tryon, R. M., and Tryon, A. F. 1982. Ferns and Allied Plants, with Special Reference to Tropical America. New York: Springer-Verlag.Google Scholar
Tryon, R. M., Tryon, A. F., and Kramer, K. U. 1990. Pteridaceae. In Kramer, K. U. and Green, P. S. (vol. ed.), Pteridophytes and Gymnosperms. In Kubitzki, K. (gen. ed.), The Families and Genera of Vascular Plants, vol. 1 (pp. 230256). Berlin: Springer-Verlag.Google Scholar
Tucker, S. C., and Douglas, A. W. 1996. Floral structure, development, and relationships of paleoherbs: Saruma, Cabomba, Lactoris, and selected Piperales. In Taylor, D. W. and Hickey, L. J. (eds.), Flowering Plant Origin, Evolution and Phylogeny (pp. 141175). New York: Chapman & Hall.Google Scholar
Turner, M. G., Baker, W. L., Peterson, C. J., and Peet, R. K. 1998. Factors influencing succession: lessons from large, infrequent natural disturbances. Ecosystems 1: 511523.Google Scholar
Turrill, W. B. 1959. Plant geography. In Rollins, R. C. and Taylor, G. (eds.), Vistas in Botany, vol. 2 (pp. 172229). London: Pergamon.Google Scholar
Uhl, N. W. 1969. Floral anatomy of Juania, Ravenea, and Ceroxylon (Palmae-Arecoideae). Gentes Herb. 10: 394411.Google Scholar
UNEP. 2003. Pilot assessmants: ecological and socio-economic impacts of invasive alien species on islands. CBD/SBSTTA/9/ INF/33.Google Scholar
Valdebenito, H., Stuessy, T. F., and Crawford, D. J. 1990a. A new biogeographic connection between islands in the Atlantic and Pacific Oceans. Nature 347: 549550.Google Scholar
Valdebenito, H., Stuessy, T. F., and Crawford, D. J. 1990b. Synonymy in Peperomia berteroana (Piperaceae) results in biological disjunction between Pacific and Atlantic Oceans. Brittonia 42: 121124.Google Scholar
Valdebenito, H., Stuessy, T. F., Crawford, D. J., and Silva O., M. 1992a. Evolution of Erigeron (Compositae) in the Juan Fernández Islands, Chile. Syst. Bot. 17: 470480.Google Scholar
Valdebenito, H., Stuessy, T. F., Crawford, D. J., and Silva O., M. 1992b. Evolution of Peperomia (Piperaceae) in the Juan Fernández Islands, Chile. Plant Syst. Evol. 182: 107119.Google Scholar
Valdebenito, H. A. 1989. Evolution of Erigeron (Compositae) and Peperomia (Piperaceae) in the Juan Fernández Islands, Chile. Ph.D. dissertation, The Ohio State University, Columbus.Google Scholar
Valenzuela F., R. 1978. La Protección Jurídica del Patrimonio Ambiental de las Islas Oceánicas Chilenas. Valparaíso: Ediciones Universitarias.Google Scholar
van Balgooy, M. M. J. 1960. Preliminary plant-geographical analysis of the Pacific. Blumea 10: 385430.Google Scholar
van Balgooy, M. M. J. 1971. Plant-geography of the Pacific. Blumea Suppl. 6: 1222.Google Scholar
Van den Heede, C. J., Viane, R. L. L., and Chase, M. W. 2003. Phylogenetic analysis of Asplenium subgenus Ceterach (Pteridophyta: Aspleniaceae) based on plastid and nuclear ribosomal ITS DNA sequences. Am. J. Bot. 90: 481495.Google Scholar
Van der Aart, P. J. M., and Vulto, J. C. 1992. Evolutionary status, biogeography and human effects, and general ecology. In Kuiper, P. J. C. and Bos, M. (eds.), Plantago: A Multidisciplinary Study (pp. 46). Berlin: Springer-Verlag.Google Scholar
van der Pijl, L. 1982. Principles of Dispersal in Higher Plants (3rd edn.). New York: Springer-Verlag.Google Scholar
van Hengstum, T., Lachmuth, S., Oostermeijer, J. G. B., den Nijs, H. C. M., Meirmans, P. G., et al. 2012. Human-induced hybridization among congeneric endemic plants on Tenerife, Canary Islands. Plant Syst. Evol. 298 : 11191131.Google Scholar
Van Loon, H. 1972. Pressure in the southern hemisphere. In Newton, C. W. (ed.), Meteorology of the Southern Hemisphere (pp. 5986). Boston: American Meteorological Society.Google Scholar
Vanderpoorten, A., Devos, N., Goffinet, B., Hardy, O. J., and Shaw, A. J. 2008. The barriers to oceanic island radiation in bryophytes: insights from the phylogeography of the moss Grimmia montana. J. Biogeogr. 35: 654663.Google Scholar
Vargas, R., Cuevas, J. G., Le Quesne, C., Reif, A., and Bannister, J. 2010. Spatial distribution and regeneration strategies of the main forest trees on Robinson Crusoe Island. Rev. Chil. Hist. Nat. 83: 349363.Google Scholar
Vargas, R., Reif, A., and Faúndez, M. J. 2011. The forests of Robinson Crusoe Island, Chile: an endemism hotspot in danger. Bosque 32: 155164.Google Scholar
Vaux, F., Trewick, S. A., and Morgan-Richards, M. 2016. Lineages, splits and divergence challenge whether the terms anagenesis and cladogenesis are necessary. Biol. J. Linn. Soc. 117: 165176.Google Scholar
Vazačová, K., and Münzbergová, Z. 2014. Dispersal ability of island endemic plants: what can we learn using multiple dispersal traits? Flora 209: 530539.Google Scholar
Verboom, G. A. 2006. A phylogeny of the schoenoid sedges (Cyperaceae: Schoeneae) based on plastid DNA sequences, with special reference to the genera found in Africa. Mol. Phylogenet. Evol. 38: 7989.Google Scholar
Vergara C., H., and Morales G., E. 1985. Morfología submarina del segmento central del Cordón Asísmico Juan Fernández, Pacífico Suroriental. In Arana E., P. (ed.), Investigaciones Marinas en el Archipiélago de Juan Fernández (pp. 2534). Santiago: Editorial Universitaria.Google Scholar
Vergara C., H., and Valenzuela A., E. 1982. Morfología submarina del guyot O’Higgins, estremo oriental del cordón asísmico Juan Fernández. Actas III Congr. Geol. Chil. 1: 132145.Google Scholar
Vicuña Mackenna, B. 1883. Juan Fernández. Historia Verdadera de la Isla de Robinson Crusoe. Santiago: Rafael Jover (facsimile ed., 1974, Ediciones Universitarias de Valparaíso, Chile).Google Scholar
Vignolo-Lutati, F. 1955. L’opera botanica del dottore Carlo Bertero di S. Vittoria d’Alba (1789–1831) nelle Antille e Sud-America (1816–1821 e 1827–1831), quale risulta dalle collezioni dell’Instituto del Orto Botanico della Universita di Torino. Mem. Acad. Sci. Torino, ser. 3a, t. II (1): 1266.Google Scholar
Viljoen, J.-A., Muasya, A. M., Barrett, R. L., Bruhl, J. J., Gibbs, A. K., et al. 2013. Radiation and repeated transoceanic dispersal of Schoeneae (Cyperaceae) through the southern hemisphere. Am. J. Bot. 100: 24942508.Google Scholar
Vives Moreno, A. 2008. In memoriam Doctor Eugene Gordon Munroe (1919–2008). SHILAP Rev. Lepidopterol. 36: 449455.Google Scholar
von Bohlen V., C. 1995. El género Mimulus L. (Scrophulariaceae) en Chile. Gayana Bot. 52: 728.Google Scholar
von Rodt, C. 1907. Aus Central- und Südamerica. Bern: W. Wälchli (see pp. 120 and 219 for comments on her cousin, von Rodt, K. A.).Google Scholar
von Wettstein, R. 1896. Monographie der Gattung Euphrasia. Leipzig: Wilhelm Engelmann.Google Scholar
Vuilleumier, B. S. 1971. Pleistocene changes in the fauna and flora of South America. Science 173: 771780.Google Scholar
Wace, N. M. 1960. The botany of southern oceanic islands. Proc. R. Soc. Lond. Series B Biol. Sci. 152: 475490.Google Scholar
Wagner, S. T., Hesse, L., Isnard, S., Samain, M.-S., Bolin, J., et al. 2014. Major trends in stem anatomy and growth forms in the perianth-bearing Piperales, with species focus on Aristolochia. Ann. Bot. 113: 11391154.Google Scholar
Wagner, W. L., and Funk, V. A. (eds.). 1995. Hawaiian Biogeography: Evolution on a Hot Spot Archipelago. Washington, DC: Smithsonian Institution Press.Google Scholar
Wagner, W. L., Bruegmann, M. M., Herbst, D. M., and Lau, J. Q. C. 1999. Hawaiian vascular plants at risk: 1999. Bishop Mus. Occas. Pap. 60: 158.Google Scholar
Wagner, W., Weller, S., and Sakai, A. 2005. Monograph of Schiedea. Syst. Bot. Monogr. 72: 1169.Google Scholar
Wagner, W. L., Herbst, D. R., and Sohmer, S. H. 1990. Manual of the Flowering Plants of Hawaii, vols. 1 and 2. Honolulu: University of Hawai‛i Press and Bishop Museum Press.Google Scholar
Walker, L. R., and Bellingham, P. 2011. Island Environments in a Changing World. Cambridge: Cambridge University Press.Google Scholar
Wallace, A. R. 1880. Island Life. London: Macmillan.Google Scholar
Wallace, A. R. 1895. Natural Selection and Tropical Nature: Essays on Descriptive and Theoretical Biology. London: Macmillan.Google Scholar
Walter, R., and Robins, B. 1748. A Voyage Round the World in the Years MDCCXL, I, II, III, IV by George Anson. London (edited with an introduction by Glyndwr Williams, Oxford University Press, 1974).Google Scholar
Walter, H. 1973. Vegetation of the Earth, in Relation to Climate and the Eco-physiological Conditions. London: English Universities Press.Google Scholar
Wang, J. 2015. Does Gst underestimate differentiation from marker data? Mol. Ecol. 24: 35463558.Google Scholar
Wanke, S., Jaramillo, M. A., Borsch, T., Samain, M.-S., Quandt, D., et al. 2007. Evolution of Piperales–matK gene and trnK intron sequence data reveal lineage specific resolution contrast. Mol. Phylogenet. Evol. 42: 477497.Google Scholar
Wanke, S., Samain, M.-S., Vanderschaeve, L., Mathieu, G., Goetghebeur, P., et al. 2006. Phylogeny of the genus Peperomia (Piperaceae) inferred from the trnK/matK region (cpDNA). Plant Biol. 8: 93102.Google Scholar
Wanntorp, L., and Wanntorp, H.-E. 2003. The biogeography of Gunnera L.: vicariance and dispersal. J. Biogeogr. 30: 979987.Google Scholar
Wanntorp, L., Wanntorp, H.-E., and Källersjö, M. 2002. Phylogenetic relationships of Gunnera based on nuclear ribosomal DNA ITS region, rbcL and rps16 intron sequences. Syst. Bot. 27: 512521.Google Scholar
Wanntorp, L., Wanntorp, H.-E., Källersjö, M., and Oxelman, B. 2001. Phylogeny of Gunnera. Plant Syst. Evol. 226: 85107.Google Scholar
Wanntorp, L., Wanntorp, H.-E., and Rutishauser, R. 2003. On the homology of the scales in Gunnera (Gunneraceae). Bot. J. Linn. Soc. 142: 301308.Google Scholar
Warren, B. H., et al. 2014. Islands as model systems in ecology and evolution: prospects fifty years after MacArthur-Wilson. Ecol. Lett. 18: 200217.Google Scholar
Warwick, S. I., Mummenhoff, K., Sauder, C. A., Koch, M. A., and Al-Shehbaz, I. A. 2010. Closing the gaps: phylogenetic relationships in the Brassicaceae based on DNA sequence data of nuclear ribosomal ITS region. Plant Syst. Evol. 285: 209232.Google Scholar
Waterman, P. G. 1975. New combinations in Zanthoxylum L. (1753). Taxon 24: 361366.Google Scholar
Waterway, M. J., and Starr, J. R. 2007. Phylogenetic relationships in tribe Cariceae (Cyperaceae) based on nested analyses of four molecular data sets. Aliso 23: 165192.Google Scholar
Watson, L., and Dallwitz, M. J. 1992. The Grass Genera of the World. Cambridge: Cambridge University Press.Google Scholar
Weeks, A. R., Sgro, C. M., Young, A. G., Frankham, R., Mitchell, N. J., et al. 2011. Assessing the benefits and risks of translocations in changing environments: a genetic perspective. Evol. Appl. 4: 709725Google Scholar
Weigelt, P., Steinbauer, M. J., Cabral, J. S., and Kreft, H. 2016. Late quaternary climate change shapes island biodiversity. Nature 532: 99102.Google Scholar
Weigend, M., and Luebert, F. 2009. Weeding the nettles: I. Clarifying species limits in perennial, rhizomatous Urtica (Urticaceae) from southern and central Chile and Argentina. Phytotaxa 2: 112.Google Scholar
Wessinger, C. A., Freeman, C. C., Mort, M. E., Rausher, M. D., and Hileman, L. C. 2016. Multiplexed shotgun genotyping resolves species relationships within the North American genus Penstemon. Am. J. Bot. 103: 9121–922.Google Scholar
Wester, L. 1991. Invasions and extinctions on Masatierra (Juan Fernández Islands): a review of early historical evidence. J. Hist. Geogr. 17: 1834.Google Scholar
Wettstein, R. 1922. Euphrasia formosissima. Addendum in Skottsberg, C. (ed.), The Natural History of Juan Fernandez and Easter Island, vol. 2 (pp. 209210). Uppsala: Almqvist & Wiksells.Google Scholar
Wettstein, R. von. 1896. Monographie der Gattung Euphrasia. Leipzig: Wilhelm Engelmann.Google Scholar
Wheeler, G. A. 1988. Taxonomic notes on Carex (Cyperaceae) of austral South America. Aliso 12: 97102.Google Scholar
Wheeler, G. A. 2007. Carex and Uncinia (Cyperaceae, Cariceae) from the Juan Fernández Archipelago, Chile. Darwiniana 45: 120141.Google Scholar
Wheeler, G. A., and Goetghebeur, P. 1995. Four new species of Uncinia (Cyperaceae) from northern South America. Aliso 14: 141146.Google Scholar
Whelan, C. J., Wenny, D. G., and Marquis, R. J. 2008. Ecosystem services provided by birdsAnn. NY Acad. Sci1134: 2560.Google Scholar
Whittaker, R. J. 1998. Island Biogeography: Ecology, Evolution, and Conservation. Oxford University Press.Google Scholar
Whittaker, R. J., and Fernández-Palacios, J. M. 2007. Island Biogeography: Ecology, Evolution, and Conservation (2nd edn.). New York: Oxford University Press.Google Scholar
Whittaker, R. J., Ladle, R. J., Araújo, M. B., Fernández-Palacios, J. M., Delgado, J. D., et al. 2007. The island immaturity-speciation pulse model of island evolution: an alternative to the “diversity begets diversity” model. Ecography 30: 321327.Google Scholar
Whittaker, R. J., Triantis, K. A., and Ladle, R. J. 2008. A general dynamic theory of oceanic island biogeography. J. Biogeogr. 35: 977994.Google Scholar
Whittaker, R. J., Triantis, K. A., and Ladle, R. J. 2010. A general dynamic theory of oceanic island biogeography: extending the MacArthur-Wilson theory to accommodate the rise and fall of volcanic islands. In Losos, J. B. and Ricklefs, R. E. (eds.), The Theory of Island Biogeography Revisited (pp. 88115). Princeton, NJ: Princeton University Press.Google Scholar
Wichman, S. R., Wright, S. D., Cameron, E. K., Keeling, D. J., and Gardner, R. C. 2002. Elevated genetic heterogeneity and Pleistocene climatic instability: inferences from nrDNA in New Zealand Coprosma (Rubiaceae). J. Biogeogr. 29: 943954.Google Scholar
Wiggins, I., and Porter, D. M. 1971. Flora of the Galapagos Islands. Stanford, CA: Stanford University Press.Google Scholar
Will, B., and Zizka, G. 1999. A review of the genus Greigia Regel (Bromeliaceae) in Chile. Harvard Papers Bot. 4: 205239.Google Scholar
Williamson, M. 1996. Biological Invasions. London: Chapman & Hall.Google Scholar
Willmer, P. 2011. Pollination and Floral Ecology. Princeton, NJ: Princeton University Press.Google Scholar
Willson, M. F., Irvine, A. K., and Walsh, N. G. 1989. Vertebrate dispersal syndromes in some Australian and New Zealand plant communities with geographic comparisons. Biotropica 21: 133147.Google Scholar
Wilmot-Dear, C. M., and Friis, I. 1996. The New World species of Boehmeria and Pouzolzia (Urticaceae, tribus Boehmerieae). A taxonomic revision. Opera Bot. 129: 1103.Google Scholar
Wilson, E. O. 1973. The ants of Easter Island and Juan Fernández. Pacific Insects 15: 285287.Google Scholar
Wilson, K. A. 1996. Alien ferns in Hawaii. Pacific Sci. 50: 127141.Google Scholar
Windham, M. D. 1986. Reassessment of the phylogenetic relationships of Notholaena. Am. J. Bot. 73: 742 (Abstr.).Google Scholar
Windham, M. D. 1987. Argyrochosma, a new genus of cheilanthoid ferns. Am. Fern J. 77: 3741.Google Scholar
Windham, M. D., Wolf, P. G., and Ranker, T. A. 1986. Factors affecting prolonged spore viability in herbarium collections of three species of Pellaea. Am. Fern J. 76: 141148.Google Scholar
Wolf, P. G., Sipes, S. D., White, M. R., Martines, M. L., Pryer, K. M., et al. 1999. Phylogenetic relationships of the enigmatic fern families Hymenophyllopsidaceae and Lophosoriaceae: evidence from rbcL nucleotide sequences. Plant Syst. Evol. 219: 263270.Google Scholar
Woodward, R. L. Jr. 1969. Robinson Crusoe’s Island. Chapel Hill: University of North Carolina Press.Google Scholar
Wörz, A. 2005. A new subgeneric classification of the genus Eryngium L. (Apiaceae, Saniculoideae). Bot. Jahrb. Syst. 126: 253259.Google Scholar
Wörz, A. 2007. The “Botanische Reiseverein” – a 19th-century joint stock company for the collecting of herbarium specimens. Huntia (Pittsburgh) 13: 121141.Google Scholar
Wörz, A. 2011. Revision of Eryngium L. (Apiaceae-Saniculoideae): general part and Palaearctic species. Biblioth. Bot. 159: 1498.Google Scholar
Wright, S. 1951. The genetic structure of populations. Ann. Eugenics 15: 323354.Google Scholar
Wu, M.-J., Huang, T.-C., and Huang, S.-F. 2009. Phylogenetic biogeography of Euphrasia section Malesianae (Orobanchaceae) in Taiwan and Malesia. Blumea 54: 242247.Google Scholar
Wysocki, W. P., Clark, L. G., Attigala, L., Ruiz-Sánchez, E., and Duvall, M. R. 2015. Evolution of the bamboos (Bambusoideae; Poaceae): a full plastome phylogenomic analysis. BMC Evol. Biol. 15: 50. Doi 10.1186/s12862-015–0321–5.Google Scholar
Yakimowski, S. B., and Rieseberg, L. H. 2014. The role of homoploid hybridization in evolution: a century of studies synthesizing genetics and ecology. Am. J. Bot. 101: 112.Google Scholar
Yañez, E., Silva, C., Vega, R., Espíndola, F., Álvarez, L., et al. 2009. Seamounts in the southeastern Pacific Ocean and biodiversity on Juan Fernández seamounts, Chile. Latin Am. J. Aquat. Res. 37: 555570.Google Scholar
Yañez R., E., Pizarro G., L., Barbieri B., M. A., and Barra A., O. 1985. Dinámica del stock de langosta (Jasus frontalis H. Milne Edwards, 1837) explotado en el archipiélago de Juan Fernández (33°40’S – 80°W). In Arana, P. (ed.), Investigaciones Marinas en el Archipiélago de Juan Fernández (pp. 251271). Valparaíso: Escuela de Ciencias del Mar, Universidad Católica de Valparaíso.Google Scholar
Yatskievych, G., and Arbeláez A., A. L. 2008. A new species and three generic transfers in the fern genus Notholaena (Pteridaceae). Novon 18: 120124.Google Scholar
Yelles-Chaouche, A., Francheteau, J., and Patriat, P. 1987. Evolution of the Juan Fernandez microplate during the last three million years. Earth Planet. Sci. Lett. 86: 269286.Google Scholar
Yen, A. C., and Olmstead, R. G. 2000a. Molecular systematics of Cyperaceae tribe Cariceae based on two chloroplast DNA regions: ndhF and trnL intron-intergenic spacer. Syst. Bot. 25: 479494.Google Scholar
Yen, A. C., and Olmstead, R. G. 2000b. Phylogenetic analysis of Carex (Cyperaceae): generic and subgeneric relationships based on chloroplast DNA. In Wilson, K. L. and Morrison, D. A. (eds.), Monocots: Systematics and Evolution (pp. 602609). Collingwood, Victoria, Austrialia: CSIRO Publishing.Google Scholar
Yu, R., Baloch, S. U., Liu, L., Pan, Q., Gong, S., et al. 2015. The phylogenetic relationships among germplasm resources of wild ramie (Boehmeria nivea L. Gaud) in China based on trnL-F and ITS sequences. Pakistan J. Bot. 47: 14511457.Google Scholar
Yuncker, T. 1953. The Piperaceae of Argentina, Bolivia and Chile. Lilloa 27: 97303.Google Scholar
Zahlbruckner, A. 1924. Die Flechten der Juan Fernandez-Inseln. In Skottsberg, C. (ed.), The Natural History of the Juan Fernandez and Easter Island, vol. 2: Botany. Uppsala: Almqvist & Wiksells.Google Scholar
Zapata, F. 2013. A multilocus phylogenetic analysis of Escallonia (Escalloniaceae): diversification in montane South America. Am. J. Bot. 100: 526545.Google Scholar
Zavada, M. S., and Benson, J. M. 1987. First fossil evidence for the primitive angiosperm family Lactoridaceae. Am. J. Bot. 74: 15901594.Google Scholar
Zavada, M. S., and Taylor, T. N. 1986. Pollen morphology of Lactoridaceae. Plant Syst. Evol. 154: 3139.Google Scholar
Záveská Drábková, L., and Vlček, Č. 2010. Molecular phylogeny of the genus Luzula DC. (Juncaceae, Monocotyledones) based on plastome and nuclear ribosomal regions: a case of incongruence, incomplete lineage sorting and hybridization. Mol. Phylogenet. Evol. 57: 536551.Google Scholar
Zhang, L., Rothfels, C. J., Ebihara, A., Schuettpelz, E., Le Péchon, T., et al. 2015. A global plastid phylogeny of the brake fern genus Pteris (Pteridaceae) and related genera in the Pteridoideae. Cladistics 31: 406423.Google Scholar
Zhang, L., Schuettpelz, E., Rothfels, C. J., Zhou, X.-M., Gao, X.-F., and Zhang, L.-B. 2016. Circumscription and phylogeny of the fern family Tectariaceae based on plastid and nuclear markers, with the description of two new genera: Draconopteris and Malaifilix (Tectariaceae). Taxon 65: 723738.Google Scholar
Zhang, L.-B., Zhang, L., Dong, S.-Y., Sessa, E. B., Gao, X.-F., et al. 2012. Molecular circumscription and major evolutionary lineages of the fern genus Dryopteris (Dryopteridaceae). BMC Evol. Biol. 12: 180; www.biomedcentral.com/1471–2148/12/180.Google Scholar
Ziegler, A. C. 2002. Hawaiian Natural History, Ecology, and Evolution. Honolulu: University of Hawaii Press.Google Scholar
Zimmerman, E. C. 1948. Insects of Hawaii. Honolulu: University of Hawaii Press.Google Scholar
Zizka, G. 1991. Die Chonta-Palme (Juania australis) der Juan Fernández Inseln (Chile). Der Palmengarten 2: 2124.Google Scholar
Zizka, G., Trumpler, K., and Zöllner, O. 2002. Revision of the genus Ochagavia (Bromeliaceae, Bromelioideae). Willdenowia 32: 331350.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
×