Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-v5vhk Total loading time: 0 Render date: 2024-06-23T06:01:31.614Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 June 2016

David Briggs
Affiliation:
University of Cambridge
S. Max Walters
Affiliation:
University of Cambridge Botanic Garden
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2016

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

Ab-Shukor, N. A., Kay, Q. O. N., Stevens, D. P. & Skibinski, D. O. F. (1988). Salt tolerance in natural populations of Trifolium repens. New Phytologist, 109, 483–9.CrossRefGoogle Scholar
Abbott, R. J. & Brochmann, C. (2003). History and evolution of the arctic flora: in the footsteps of Eric Hultén. Molecular Ecology, 12, 299–313.CrossRefGoogle ScholarPubMed
Abbott, R. J. & Forbes, D. G. (2002). Extinction of the Edinburgh lineage of the allopolyploid neospecies, Senecio cambrensis Rosser (Asteraceae). Heredity, 88, 267–9.CrossRefGoogle Scholar
Abbott, R. J., Ingram, R. & Noltie, H. J. (1983a). Discovery of Senecio cambrensis Rosser in Edinburgh. Watsonia, 14, 407–8.Google Scholar
Abbott, R. J., Noltie, H. J. & Ingram, R. (1983b). The origin and distribution of Senecio cambrensis Rosser in Edinburgh. Transactions of the Botanical Society of Edinburgh, 44, 103–6.CrossRefGoogle Scholar
Abbott, R. J., Ireland, H. E. & Rogers, H. J. (2007). Population decline despite high genetic diversity in the new allopolyploid species Senecio cambrensis (Asteraceae). Molecular Ecology, 16, 1023–33.CrossRefGoogle Scholar
Abbott, R. J.et al. (2010). Homoploid hybrid speciation in action. Taxon, 59, 1375–86.Google Scholar
Abbott, R. J.et al. (2012). Hybridization and speciation. Journal of Evolutionary Biology, 26, 229–46.Google Scholar
Abramovitz, J. N. (1996). Imperiled waters, impoverished future: the decline of freshwater ecosystems. Washington DC: Worldwatch.Google Scholar
Adams, J. U. (2008). Transcriptome: connecting the genome to gene function. Nature Education, 1, 195.Google Scholar
Adams, W. M. (2004). Against extinction: the story of conservation. London & Sterling, VA: Earthspan.Google Scholar
Agnarsson, I. & Kuntner, M. (2007). Taxonomy in a changing world: seeking solutions for a science in crisis. Systematic Biology, 56, 531–9.CrossRefGoogle Scholar
Agrawal, A. A., Laforsch, C. & Tollrian, R. (1999). Transgenerational induction of defences in animals and plants. Nature, 401, 60–3.CrossRefGoogle Scholar
Ågren, J. & Schemske, D. W. (2012). Reciprocal transplants demonstrate strong adaptive differentiation of the model organism Arabidopsis thaliana in its native range. New Phytologist, 194, 1112–22.CrossRefGoogle ScholarPubMed
Aguilar, R.et al. (2006). Plant reproductive susceptibility to habitat fragmentation: review and synthesis through a meta-analysis. Ecological Letters, 9, 968–80.CrossRefGoogle ScholarPubMed
Ahmed, S.et al. (2009). Wind-borne insects mediate directional pollen transfer between desert fig trees 160 kilometers apart. Proceedings of the National Academy of Sciences, USA, 106, 20342–7.CrossRefGoogle ScholarPubMed
Ainouche, M. L.et al. (2009). Hybridization, polyploidy and invasion: lessons from Spartina (Poaceae). Biological Invasions, 11, 1159–73.CrossRefGoogle Scholar
Akçakaya, H. R. (2002). RAMAS Metapop: Viability Analysis for Stage-structured Metapopulations (version 4.0). Setauket, NY: Applied Biomathematics.Google Scholar
Akeroyd, J. R. (1994). Some problems with introduced plants in the wild. In The common ground of wild and cultivated plants, ed. Perry, A. R. & Ellis, R. G., pp. 31–40. Cardiff: National Museum of Wales.Google Scholar
Akeroyd, J. R. & Wyse Jackson, P. (1995). A handbook for botanic gardens on the reintroduction of plants to the wild. London: Botanic Gardens Conservation International.Google Scholar
Al-Hiyaly, S. E. K., McNeilly, T. & Bradshaw, A. D. (1988). The effect of zinc contamination from electricity pylons – evolution in a replicated situation. New Phytologist, 110, 571–80.CrossRefGoogle Scholar
Al-Hiyaly, S. E. K., McNeilly, T. & Bradshaw, A. D. (1990). The effect of zinc contamination from electricity pylons – contrasting patterns of evolution in five grass species. New Phytologist, 114, 183–90.CrossRefGoogle Scholar
Al-Hiyaly, S. E. K., McNeilly, T., Bradshaw, A. D. & Mortimer, A. M. (1993). The effect of zinc contamination from electricity pylons. Genetic constraints on selection for zinc tolerance. Heredity, 70, 22–32.CrossRefGoogle Scholar
Al-Kaff, N.et al. (2008). Detailed dissection of the chromosomal region containing the Ph1 Locus in Wheat Triticum aestivum: with deletion mutants and expression profiling. Annals of Botany, 101, 863–72.CrossRefGoogle ScholarPubMed
Albert, V. A., Williams, S. E. & Chase, M. W. (1992). Carnivorous plants: phylogeny and structural evolution. Science, 257, 1491–5.CrossRefGoogle ScholarPubMed
Albert, V. A., Oppenheimer, D. & Lindqvist, C. (2002). Pleiotropy, redundancy and the evolution of flowers. Trends in Plant Science, 7, 297–301.CrossRefGoogle ScholarPubMed
Alberto, F. J.et al. (2013a). Imprints of natural selection along environmental gradients in phenology-related genes of Quercus petraea. Genetics, 195, 495–512.CrossRefGoogle ScholarPubMed
Alberto, F. J.et al. (2013b). Potential for evolutionary responses to climate change – evidence from tree populations. Global Climate Change, 19, 1645–61.Google ScholarPubMed
Alcázar, R.et al. (2012). Signals of speciation within Arabidopsis thaliana in comparison with its relatives. Current Opinion in Plant Biology, 15, 205–11.CrossRefGoogle ScholarPubMed
Aldridge, S. (1996). The thread of life. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Alexander, D. (2011). The language of genetics. London: Longman.Google Scholar
Ali, N. S. & Trivedi, C. (2010). Plant diversity and climate change – a review of Kew's science activities relevant to climate change. Richmond, UK: Royal Botanic Gardens, Kew.Google Scholar
Ali, N. S. & Trivedi, C. (2011). Botanic Gardens and climate change – an audit of scientific activities at the Royal Botanic Gardens, Kew. Biodiversity Conservation, 20, 295–307.CrossRefGoogle Scholar
Allan, E. & Pannell, J. R. (2009). Rapid divergence in physiological and life-history traits between northern and southern populations of the British introduced neo-species, Senecio squalidus. Oikos, 118, 1053–61.CrossRefGoogle Scholar
Allen, A. M.et al. (2011). Pollen–pistil interactions and self-incompatibility in the Asteraceae: new insights from studies of Senecio squalidus (Oxford Ragwort). Annals of Botany, 108, 687–98.CrossRefGoogle Scholar
Allendorf, F. W, Luikart, G. & Aitken, S. N. (2013) Conservation and the genetics of populations, edn. Oxford: Blackwell Publishing.Google Scholar
Allison, L. A. (2007). Fundamental molecular biology. Malden, MA, Oxford & Carlton, Australia: Blackwell.Google Scholar
Allnutt, T. R.et al. (1998). Genetic variation in Fitzroya cupressoides (Alerce), a threatened South American conifer. Molecular Ecology, 8, 975–87.Google Scholar
Ally, D., Ritland, K. & Otto, S. P (2008). Can clone size serve as a proxy for clone age? An exploration using microsatellite divergence in Populus tremuloides. Molecular Ecology, 17, 4897–911.CrossRefGoogle ScholarPubMed
Alonso-Blanco, C., Mendez-Vigo, B. & Koornneef, M. (2005). From phenotypic to molecular polymorphisms involved in naturally occurring variation of plant development. International Journal of Developmental Biology, 49, 717–32.CrossRefGoogle ScholarPubMed
Alsos, I. G.et al. (2007). Frequent long-distance plant colonization in the changing Arctic. Science, 316, 1606–9.CrossRefGoogle ScholarPubMed
Alsos, I. G.et al. (2009). Past and future range shifts and loss of diversity in Dwarf Willow (Salix herbacea L.) inferred from genetics, fossils, and modelling. Global Ecology and Biogeography, 18, 223–39.CrossRefGoogle Scholar
Alston, R. E. & Turner, B. L. (1963a). Biochemical systematics. London & New York: Prentice Hall.Google Scholar
Alston, R. E. & Turner, B. L. (1963b). Natural hybridization among four species of Baptisia (Leguminosae). American Journal of Botany, 50, 159–73.CrossRefGoogle Scholar
Altizer, S. M., Thrall, P. H. & Antonovics, J. (1998), Vector behaviour and transmission of anther-smut infection in Silene alba. American MidlandNaturalist, 139, 147–63.Google Scholar
Alvarez, L. W., Alvarez, F. A., Asaro, F. & Michel, H. V. (1980). Extraterrestrial cause for the Cretaceous–Tertiary extinction. Science, 208, 1095–108.CrossRefGoogle ScholarPubMed
Alvarez-Valin, F. (2002). Neutral theory. In Encyclopedia of evolution, vol. 2, 815–21, Oxford: Oxford University Press.Google Scholar
Amborella Genome Project (2013). The Amborella genome and the evolution of flowering plants. Science, 342, DOI:10.1126/science.1241089
Amsellen, L., Noyer, J. L, Le Bourgeois, T. & Hossaert-McKey, M. (2000). Comparison of genetic diversity of the invasive weed Rubus alceifolius Poir. (Rosaceae) in its native range and areas of introduction, using amplified fragment length polymorphism (AFLP) markers. Molecular Ecology, 9, 443–55.Google Scholar
Amsellem, L., Chevalier, M. H. & Hossaert-McKey, M. (2001). Ploidy level of the invasive weed Rubus alceifolius (Rosaceae) in its native range and in areas of introduction. Plant Systematics and Evolution, 228, 171–9.CrossRefGoogle Scholar
Amsellem, L., Noyer, J.-L. & Hossaert-McKey, M. (2001). Evidences of a switch in the reproductive biology of Rubus alceifolius (Rosaceae) towards apomixis, between its native range and its area of introduction. American Journal of Botany, 88, 2243–51.CrossRefGoogle Scholar
Anderson, D. L. & East, I. J. (2008).The latest buzz about colony collapse disorder. Science, 319, 724.CrossRefGoogle ScholarPubMed
Anderson, E. (1949). Introgressive hybridisation. London: Chapman & Hall; New York: Wiley.Google Scholar
Anderson, E. (1953). Introgressive hybridization. Biological Reviews, 28, 280–307.CrossRefGoogle Scholar
Anderson, E. & Abbe, L. B. (1933). A comparative anatomical study of a mutant Aquilegia. American Naturalist, 67, 380–4.CrossRefGoogle Scholar
Anderson, G.et al. (2006). Reproductive biology of the dioecious Canary Islands endemic Withania aristata (Solanaceae). American Journal of Botany, 93, 1295–1305.CrossRefGoogle Scholar
Anderson, J. T.et al. (2012). Phenotypic plasticity and adaptive evolution contribute to advancing flowering phenology in response to climate change. Proceedings of the Royal Society, B, 279, 3843–52.CrossRefGoogle ScholarPubMed
Andreasen, K., Manktelow, M. & Razafimandimbison, S. G. (2009). Successful DNA amplification of a more than 200-year-old herbarium specimen: recovering genetic material from the Linnaean era. Taxon, 58, 959–62.Google Scholar
Andrews, C. A. (2010). Natural selection, genetic drift, and gene flow do not act in isolation in natural populations. Nature Education Knowledge, 3, 5.Google Scholar
Andrus, N.et al. (2009). Phylogenetics of Darwiniothamnus (Asteraceae: Astereae) – molecular evidence for multiple origins in the endemic flora of the Galápagos Islands. Journal of Biogeography, 36, 1055–69.CrossRefGoogle Scholar
Angiosperm Phylogeny Group (2009). An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APGIII. Botanical Journal of the Linnean Society, 161, 105–21.
Anon. (1965). Iconographia Mendeliana. Brno: The Moravian Museum.
Anon. (1994). IUCN Red List categories. Gland, Switzerland: IUCN Council.
Anon. (2002). Seed banks receive vital cash boost. New Scientist, 30 August.
Anon. (2014). Protect and serve. Nations must keep expanding conservation efforts to avoid a biodiversity crisis. Editorial in Nature, 616, 144.
Antonovics, J. (1968). Evolution in closely adjacent plant populations. V. Evolution of self-fertility. Heredity, 23, 219–38.Google Scholar
Antonovics, J. & Bradshaw, A. D. (1970). Evolution in closely adjacent plant populations. VIII. Clinal patterns at a mine boundary. Heredity, 25, 349–62.CrossRefGoogle Scholar
Antonovics, J., Bradshaw, A. D. & Turner, R. G. (1971). Heavy metal tolerance in plants. Advances in Ecological Research, 7, 1–85.Google Scholar
Arabidopsis Genome Initiative (2000). Analysis of the genome sequence of the flowering plant Arabidopsis thaliana. Nature, 408, 796–815.
Araújo, M. B. & Peterson, T. (2011). Uses and misuses of bioclimatic envelope modeling. Ecology, 93, 1527–39.Google Scholar
Arber, A. (1938). Herbals: their origin and evolution. A chapter in the history of Botany 1470–1670, edn. Cambridge: Cambridge University Press.Google Scholar
Arbogast, B. S.et al. (2002). Estimating divergence times from molecular data on phylogenetic and population genetic timescales. Annual Review of Ecology & Systematics, 33, 707–40.CrossRefGoogle Scholar
Archibald, J. D. & Fastovsky, D. E. (2004). Dinosaur extinction. In The Dinosauria, ed. Weishampel, D. B.et al., pp. 672–684. Berkeley: University of California Press.Google Scholar
Armbruster, W. S. & Baldwin, B. G. (1998). Switch from specialized to generalized pollination. Nature, 394, 632.CrossRefGoogle Scholar
Armstrong, H. E., Armstrong, F. & Horton, E. (1912). Herbage studies 1. Lotus corniculatus, a cyanophoric plant. Proceedings of the Royal Society, B, 84, 471–84.CrossRefGoogle Scholar
Arnaud-Haond, S.et al. (2012). Implications of extreme life span in clonal organisms: millenary clones in meadows of the threatened seagrass Posidonia oceanica. PLOS ONE, 7, e30454, DOI:10.1371/journal.pone.0030454.CrossRefGoogle ScholarPubMed
Arnold, A. C.et al. (2005). Is there a future for wild grapevine (Vitis vinifera subsp. silvestris) in the Rhine Valley?Biodiversity and Conservation, 14, 1507–23.CrossRefGoogle Scholar
Arnold, A. E.et al. (2010). Interwoven branches of the plant and fungal trees of life. New Phytologist, 185, 874–8.CrossRefGoogle ScholarPubMed
Arnold, M. L. (1992). Natural hybridization as an evolutionary process. Annual Review of Ecology and Systematics, 23, 237–61.CrossRefGoogle Scholar
Arnold, M. L. (1993). Iris nelsonii (Iridaceae): origin and genetic composition of a homoploid hybrid species. American Journal of Botany, 80, 577–83.CrossRefGoogle Scholar
Arnold, M. L. & Bennett, B. D. (1993). Natural hybridization in Louisiana Irises: genetic variation and ecological determinants. In Hybrid zones and the evolutionary process, ed. Harrison, R. G., pp. 115–39. New York: Oxford University Press.Google Scholar
Arnold, M. L., Bouck, A. C. & Cornman, R. S. (2004). Verne Grant and Louisiana Irises: is there anything new under the sun?New Phytologist, 161, 143–9.Google Scholar
Arnold, M. L., Ballerini, E. S. & Brothers, A. N. (2012). Hybrid fitness, adaptation and evolutionary diversification: lessons learned from Louisiana Irises. Heredity, 108, 159–66.CrossRefGoogle ScholarPubMed
Aronne, G. & Wilcock, C. C. (1994). First evidence of myrmecochory in fleshy-fruited shrubs of the Mediterranean region. New Phytologist, 127, 781–8.CrossRefGoogle Scholar
Aronson, J.et al. (1993). Restoring natural capital: science, business and practice. Washington DC, & London: Island Press.Google Scholar
Ashley, M. V. (2010). Plant parentage, pollination, and dispersal: how DNA microsatellites have altered the landscape. Critical Reviews in Plant Sciences, 29, 148–61.CrossRefGoogle Scholar
Ashman, T. L.et al. (2004). Pollen limitation of plant reproduction: ecological and evolutionary causes and consequences. Ecology, 85, 2408–21.CrossRefGoogle Scholar
Ashman, T. L., Kwok, A. & Husband, B. C. (2013). Revisiting the dioecy–polyploidy association: alternate pathways and research opportunities. Cytogenetic and Genome Research, 140, 241–55.CrossRefGoogle ScholarPubMed
Ashton, P. A. & Abbott, R. J. (1992). Multiple origins and genetic diversity in the newly arisen allopolyploid species, Senecio cambrensis Rosser (Compositae). Heredity, 68, 25–32.CrossRefGoogle Scholar
Ashton, P. S. (1987). Biological considerations in in situ vs ex situ plant conservation. In Botanic gardens and the world conservation strategy, ed. Bramwell, D., Hamann, O., Heywood, V. & Synge, H., pp. 117–30. London: Academic Press for IUCN.Google Scholar
Asker, S. E. & Jerling, L. (1992). Apomixis in plants. Boca Raton, FL: CRC Press.Google Scholar
Assouad, M. W.et al., (1978). Reproductive capacities in the sexual forms of the gynodiecious species Thymus vulgaris. Botanical Journal of the Linnean Society, 77, 29–39.CrossRefGoogle Scholar
Assunçáo, A. G. L., Schat, H. & Aarts, M. G. M. (2003). Thlaspi caerulescens, an attractive model species to study heavy metal hyperaccumulation in plants. New Phytologist, 159, 351–60.CrossRefGoogle Scholar
Aston, J. L. & Bradshaw, A. D. (1966). Evolution in closely adjacent plant populations II. Agrostis stolonifera in maritime habitats. Heredity, 21, 649–64.CrossRefGoogle Scholar
Atwell, S.et al. (2010). Genome-wide association study of 107 phenotypes in Arabidopsis thaliana inbred lines. Nature, 465, 627–31.CrossRefGoogle ScholarPubMed
Atwood, S. S. & Sullivan, J. T. (1943). Inheritance of a cyanogenic glucoside and its hydrolysing enzyme in Trifolium repens. Journal of Heredity, 34, 311–20.CrossRefGoogle Scholar
Auffret, A. G. (2011). Can seed dispersal by human activity play a useful role for the conservation of European grasslands?Applied Vegetation Science, 14, 291–303.CrossRefGoogle Scholar
Avers, C. J. (1989). Process and pattern in evolution. New York: Oxford University Press.Google Scholar
Avise, J. C. (1994). Molecular markers, natural history and evolution. New York: Chapman & Hall.CrossRefGoogle Scholar
Avise, J. C. (2000). Phylogeography: the history and formation of species. Cambridge, MA: Harvard University Press.Google Scholar
Ayala, F. J. (2012). Evolution. London: Quercus.Google Scholar
Ayazloo, M. & Bell, J. N. B. (1981). Studies on the tolerance to sulphur dioxide of grass populations in polluted areas. 1. Identification of tolerant populations. New Phytologist, 88, 203–22.CrossRefGoogle Scholar
Ayre, D.et al. (2010). The accumulation of genetic diversity within a canopy-stored seed bank. Molecular Ecology, 19, 2640–50.CrossRefGoogle ScholarPubMed
Ayres, D. R. & Strong, D. R. (2001). Origin and genetic diversity of Spartina anglica (Poaceae) using nuclear DNA markers. American Journal of Botany, 88, 1863–7.CrossRefGoogle ScholarPubMed
Bacles, C. F. E., Lowe, A. J. & Ennos, R. A. (2006). Effective seed dispersal across a fragmented landscape. Science, 311, 628.CrossRefGoogle ScholarPubMed
Bailey, J. P., Bímová, K. & Mandák, B. (2009). Asexual spread versus sexual reproduction and evolution in Japanese Knotweeds sets the stage for the ‘Battle of the Clones’. Biological Invasions, 11, 1189–1203.CrossRefGoogle Scholar
Bailleul, D.et al. (2012). Seed spillage from grain trailers on road verges during oilseed rape harvest: an experimental survey. PLOS ONE, 7, 1–7.CrossRefGoogle Scholar
Baker, A. M., Barrett, S. C. H. & Thompson, J. D. (2000). Variation of pollen limitation in the early flowering Mediterranean geophyte Narcissus assoanus (Amaryllidaceae). Oecologia, 124, 529–35.CrossRefGoogle Scholar
Baker, H. (1966). The evolution, functioning and breakdown of heteromorphic incompatibility systems. I. The Plumbaginaceae. Evolution, 20, 349–68.CrossRefGoogle ScholarPubMed
Baker, H. G. (1947). Criteria of hybridity. Nature, 159, 1–5.Google ScholarPubMed
Baker, H. G. (1951). Hybridization and natural gene-flow between higher plants. Biological Reviews, 26, 302–37.CrossRefGoogle Scholar
Baker, H. G. (1954). Report of meeting of British Ecological Society, April 1953. Journal of Ecology, 42, 570–2.Google Scholar
Baker, H. G. (1955). Self-compatibility and establishment after ‘long-distance’ dispersal. Evolution, 9, 347–9.Google Scholar
Baker, H. G. (1965). Characteristics and modes of origin of weeds. In The genetics of colonizing species, ed. Baker, H. G. & Stebbins, G. L., pp. 147–72. New York: Academic Press.Google Scholar
Baker, H. G. (1967). Support for Baker's Law – as a rule. Evolution, 21, 853–6.CrossRefGoogle Scholar
Baker, H. G. (1974). The evolution of weeds. Annual Review of Ecology & Systematics, 5, 1–24.CrossRefGoogle Scholar
Baker, H. G. (1991). The continuing evolution of weeds. Economic Botany, 45, 445–9.CrossRefGoogle Scholar
Baker, H. G. & Cox, P. A. (1984). Further thoughts on dioecism and islands. Annals of the Missouri Botanical Garden, 71, 244–53.CrossRefGoogle Scholar
Bakker, E. G.et al. (2006). A genome-wide survey of R gene polymorphisms in Arabidopsis. Plant and Cell, 18, 1803–18.CrossRefGoogle ScholarPubMed
Balakrishnan, R. (2005). Species concepts, species boundaries and species identification: a view from the tropics. Systematic Biology, 54, 689–93.CrossRefGoogle ScholarPubMed
Balcombe, D. (2011). Invigorating plants. Research Horizons, 5, 9 and http://phys.org/news/2011-07-invigorating.htmlGoogle Scholar
Baldwin, B. G. (2006). Contrasting patterns and processes of evolutionary change in the tarweed-silversword lineage: revisiting Clausen, Keck, and Heisey's findings. Annals of the Missouri Botanical Garden, 93, 66–96.CrossRefGoogle Scholar
Baldwin, B. G. & Wagner, W. L. (2010). Hawaiian angiosperm radiations of North American origin. Annals of Botany, 105, 849–79.CrossRefGoogle ScholarPubMed
Balmford, A. & Long, A. (1994). Avian endemism and forest loss. Nature, 372, 623–4.CrossRefGoogle Scholar
Balmford, A. & Whitten, T. (2003). Who should pay for tropical conservation, and how could the costs be met?Oryx, 37, 238–50.CrossRefGoogle Scholar
Bannister, M. H. (1965). Variation in the breeding system of Pinus radiata. In The genetics of colonizing species, ed. Baker, H. G. & Stebbins, G. L., pp. 353–72. New York & London: Academic Press.Google Scholar
Bannister, P. (1976). Introduction to physiological plant ecology. Oxford: Blackwell.Google Scholar
Barcaccia, G. & Albertini, E. (2013). Apomixis in plant reproduction: a novel perspective on an old dilemma. Plant Reproduction, 26, 159–79.CrossRefGoogle ScholarPubMed
Barkley, T. M. (1966). A review of the origin and development of the florists’ Cineraria, Senecio cruentus. Economic Botany, 20, 386–95.CrossRefGoogle Scholar
Barkman, T. J., Lim, S.-H., Mat Salleh, K. & Nais, J. (2004). Mitochondrial DNA sequences reveal the photosynthetic relatives of Rafflesia, the world's largest flower. Proceedings of the National Academy of Sciences, USA, 101, 787–92.CrossRefGoogle ScholarPubMed
Barkman, T. J.et al. (2007). Mitochondrial DNA suggests at least 11 origins of parasitism in angiosperms and reveals genomic chimerism in parasitic plants. BMC Evolutionary Biology, 7, 248.CrossRefGoogle ScholarPubMed
Barling, D. M. (1955). Some population studies in Ranunculus bulbosus L. Journal of Ecology, 43, 207–18.
Barlow, B. A. & Wiens, D. (1977). Host parasite resemblance in Australian Mistletoes: the case for cryptic mimicry. Evolution, 31, 69–84.CrossRefGoogle ScholarPubMed
Barnes, D. K. A. (2002). Accumulation and fragmentation of plastic debris in global environments. Philosophical Transactions of the Royal Society, B, 364, 1985–98.Google Scholar
Barnosky, A. D.et al. (2011). Has the Earth's sixth mass extinction already arrived?Nature, 471, 51–7.CrossRefGoogle ScholarPubMed
Barnosky, A. D.et al. (2012). Approaching a state shift in Earth's biosphere. Nature, 486, 52–8.CrossRefGoogle ScholarPubMed
Barrett, P. H.et al. (1987). Charles Darwin notebooks 1836–1844. Cambridge: Cambridge University Press and the British Museum (Natural History).Google Scholar
Barrett, S. C. H. (1980a) Sexual reproduction in Eichhornia crassipes (Water Hyacinth). I. Fertility of clones from diverse regions. Journal of Applied Ecology, 17, 101–12.CrossRefGoogle Scholar
Barrett, S. C. H (1980b). Sexual reproduction in Eichhornia crassipes (Water Hyacinth). II. Seed production in natural populations. Journal of Applied Ecology, 17, 113–24.CrossRefGoogle Scholar
Barrett, S. C. H. (1983). Crop mimicry in weeds. Economic Botany, 37, 255–82.CrossRefGoogle Scholar
Barrett, S. C. H. (1985). Floral trimorphism and monomorphism in continental and island populations of Eichhornia paniculata (Spreng.) Sols. (Pontederiaceae). Biological Journal of the Linnean Society, 25, 41–60.CrossRefGoogle Scholar
Barrett, S. C. H. (1987). Mimicry in plants. Scientific American, 255, 76–83.Google Scholar
Barrett, S. C. H. (1988). Mating system evolution and speciation in heterostylous plants. In Speciation and its consequences, ed. Otte, D. & Endler, J. A., pp. 257–83. Sunderland, MA: Sinauer.Google Scholar
Barrett, S. C. H. (1989a). The evolutionary breakdown of heterostyly. In The evolutionary ecology of plants, ed. Bock, J. H. & Linhart, Y. B., pp.151–69. Boulder, CO: Westview Press.Google Scholar
Barrett, S. C. H. (1989b). Mating system evolution and speciation in heterostylous plants. In Speciation and its consequences, ed. Otte, D. & Endler, J. A., pp. 257–83. Sunderland, MA: SinauerGoogle Scholar
Barrett, S. C. H. (1992). Heterostylous genetic polymorphisms: model systems for evolutionary analysis. In Evolution and function of heterostyly, ed. Barrett, S. C. H., pp. 1–29. Heidelberg: Springer Verlag.CrossRefGoogle Scholar
Barrett, S. C. H. (2010a). Darwin's legacy: the forms, function and sexual diversity of flowers. Philosophical Transactions of the Royal Society of London, Ser. B., 365, 351–68.CrossRefGoogle ScholarPubMed
Barrett, S. C. H. (2010b). Understanding plant reproductive diversity. Philosophical Transactions of the Royal Society of London, Ser. B., 365, 99–109.CrossRefGoogle ScholarPubMed
Barrett, S. C. H. & Charlesworth, D. (2007). David Graham Lloyd 20 June 1937–30 May 2006. Biographical Memoirs of the Fellows of the Royal Society, 53, 203–21.Google Scholar
Barrett, S. C. H. & Harder, L. D. (1996). Ecology and evolution of plant mating. Trends in Ecology & Evolution, 11, 73–9.CrossRefGoogle ScholarPubMed
Barrett, S. C. H. & Kohn, J. R. (1991). Genetic and evolutionary consequences of small population size in plants: implications for conservation. In Genetics and conservation of rare plants, ed. Falk, D. A. & Holsinger, K. E., pp. 3–30. New York: Oxford University Press.Google Scholar
Barrett, S. C. H. & Richardson, B. J. (1986). Genetic attributes of invading species. In Ecology of biological invasions, ed. Groves, R. H. & Brown, J. J., pp. 21–33. Canberra: Australian Academy of Science.Google Scholar
Barrett, S. C. H. & Shore, J. S. (1990). Isozyme variation in colonising plants. In Isozymes in plant biology, ed. Soltis, D. E. & Soltis, P. S., pp. 106–26. London: Chapman & Hall.Google Scholar
Barrett, S. C. H. & Shore, J. S.. (2008). New insights on heterostyly: comparative biology, ecology and genetics. In Self-incompatibility in flowering plants: evolution, diversity and mechanisms, ed. Franklin-Tong, V., pp. 3–32. Berlin: Springer-Verlag.Google Scholar
Barrett, S. C. H., Dorken, M. E. and Case, A. L. (2001). A geographical context for the evolution of plant reproductive systems. In Integrating ecological and evolutionary processes in a spatial context, ed. Silvertown, J. & Antonovics, J., pp. 341–64. Oxford: Blackwell.Google Scholar
Barrett, S. C. H., Colautti, R. I. & Eckert, C. G. (2008). Reproductive systems and evolution during biological invasion. Molecular Ecology, 17: 373–83.CrossRefGoogle ScholarPubMed
Barrett, S. C. H., Ness, R. W. & Vallejo-Marín, M. (2009). Evolutionary pathways to self-fertilization in a tristylous plant species. New Phytologist, 183, 546–56.CrossRefGoogle Scholar
Barrett, S. C. H.et al. (2011). Plant reproductive systems and evolution during biological invasion. Molecular Biology, 17, 373–83.Google Scholar
Barringer, B. C. (2007). Polyploidy and self-fertilization in flowering plants. American Journal of Botany, 94, 1527–33.CrossRefGoogle ScholarPubMed
Barros, M. D. C. & Dyer, T. A. (1988). Atrazine resistance in the grass Poa annua is due to a single base change in a chloroplast gene for the D1 protein of photosystem II. Theoretical & Applied Genetics, 75, 610–16.CrossRefGoogle Scholar
Barton, N. H. (2000). Genetic hitchhiking. Philosophical Transactions of the Royal Society of London, B, 355, 1553–62.CrossRefGoogle ScholarPubMed
Barton, N. H. & Charlesworth, B. (1984). Genetic revolutions, founder effects, and speciation. Annual Review of Ecology and Systematics, 15, 133–64.CrossRefGoogle Scholar
Bashford, A. & Levine, P. (2012). The Oxford handbook of the history of eugenics. New York & Oxford: Oxford University Press.CrossRefGoogle Scholar
Bateman, R. M. (1999). Integrating molecular and morphological evidence of evolutionary radiations. In Molecular systematics and plant evolution, ed. Hollingsworth, P. M., Bateman, R. M. & Gornall, R. J., pp. 432–71. London & New York: Taylor & Francis. Systematics Association Special Volume Series 57.Google Scholar
Bateman, R. M. (2011). The perils of addressing long-term challenges in a short-term world: making descriptive taxonomy predictive. In Climate change, ecology and systematics, ed. Hodkinson, T. R.. et al., pp.67–95. Cambridge: Cambridge University Press.Google Scholar
Bateman, R. M. & DiMichele, W. A. (2002). Generating and filtering major phenotypic novelties: neoGoldschmidtian saltation revisited. In Developmental genetics and plant evolution, ed. Cronk, Q. C. B., Bateman, R. M. & Hawkin, J. A., pp. 109–59. London: Taylor & Francis.Google Scholar
Bateman, R. M. & Sexton, R. (2008). Is spur length of Platanthera species in the British Isles adaptively optimized or an evolutionary red herring?Watsonia, 27, 1–21.Google Scholar
Bateman, R. M., Pridgeon, A. M. & Chase, M. W. (1997). Phylogenetics of subtribe Orchidineae (Orchidoideae, Orchidaceae) based on nuclear ITS sequences. 2. Infrageneric relationships and reclassification to achieve monophyly of Orchis sensu stricto. Lindleyana, 12, 113–41.Google Scholar
Bateson, W. (1895a). The origin of the cultivated Cineraria. Nature, 51, 605–7.CrossRefGoogle Scholar
Bateson, W. (1895b). The origin of the cultivated Cineraria. Nature, 52, 29, 103–4.CrossRefGoogle Scholar
Bateson, W. (1897). Notes on hybrid Cinerarias produced by Mr Lynch and Miss Pertz. Proceedings of the Cambridge Philosophical Society, 9, 308–9.Google Scholar
Bateson, W. (1909). Mendel's principles of heredity. London: Cambridge University Press; New York: Macmillan.CrossRefGoogle Scholar
Bateson, W. (1913). Problems of genetics. London: Oxford University Press; New Haven, CT: Yale University Press.CrossRefGoogle Scholar
Bateson, W. & Punnett, R. C. (1911). On gametic series involving reduplication of certain terms. Journal of Genetics, 1, 293–302.CrossRefGoogle Scholar
Bateson, W. & Saunders, E. R. (1902). Experimental studies in the physiology of heredity. Report to the Evolution Committee of the Royal Society, 1, 1–160.Google Scholar
Bateson, W., Saunders, E. R. & Punnett, R. C. (1905). Experimental studies in the physiology of heredity. Report to the Evolution Committee of the Royal Society, 2, 1–55, 80–99.Google Scholar
Battaglia, E. (1963). Apomixis. In Recent advances in the embryology of Angiosperms, ed. Maheshwari, P., ch. 8, pp. 221–64. Delhi: University of Delhi Press.Google Scholar
Baucom, R. S. (2016). The remarkable repeated evolution of herbicide resistance. American Journal of Botany, 103, 181–3.CrossRefGoogle ScholarPubMed
Baum, D. A. & Donoghue, M. J. (2002). Transference of function, heterotopy, and the evolution of plant development. In Developmental genetics and plant evolution, ed. Cronk, Q. C. B., Bateman, R. M. & Hawkin, J. A., pp. 52–69. London: Taylor & Francis.Google Scholar
Baum, D. A. & Smith, S. D. (2012). Tree-thinking: an introduction to phylogenetic biology. Greenwood Village, CO: Roberts & Company.Google Scholar
Baum, D. A, Randall, L. S. & Wendel, J. F. (1998). Biogeography and floral evolution of Baobabs (Adansonia, Bombacaceae) as inferred from multiple data sets. Systematic Biology, 47, 181–207.CrossRefGoogle ScholarPubMed
Baumann, U.et al. (2000). Self-incompatibility in the grasses. Annals of Botany, 85, 203–9.CrossRefGoogle Scholar
Baumel, A., Ainouche, M. L. & Levasseur, J. E. (2001). Molecular investigations in populations of Spartina anglica C.E. Hubbard (Poaceae) invading coastal Brittany (France). Molecular Ecology, 10, 1689–1701.CrossRefGoogle Scholar
Bawa, K. S., Perry, D. R. & Beach, J. H. (1985). Reproductive biology of tropical lowland rain forest trees. I. Sexual systems and self-incompatibility mechanisms. American Journal of Botany, 72, 331–45.Google Scholar
Bayer, R. J., Ritland, K. & Purdy, B. G. (1990). Evidence of partial apomixis in Antennaria media (Asteraceae: Inuleae) detected by segregation of genetic markers. American Journal of Botany, 77, 1078–83.CrossRefGoogle Scholar
Beardsell, D. V.et al. (1993). Reproductive biology of Australian Myrtaceae. Australian Journal of Botany, 41, 511–26.CrossRefGoogle Scholar
Beattie, A. (1978). Plant–animal interactions affecting gene flow in Viola. In The pollination of flowers by insects, ed. Richards, A. J., pp. 151–64, Linnean Society Symposium Series 6. London: Academic Press.Google Scholar
Bebber, D. P.et al. (2010). Herbaria are a major frontier for species discovery. Proceedings of the National Academy of Sciences USA, 107, 22169–71.CrossRefGoogle ScholarPubMed
Beck, J. B., Schmuths, H. & Schaal, B. A. (2007). Native range genetic variation in Arabidopsis thaliana is strongly geographically structured and reflects Pleistocene glacial dynamics. Molecular Ecology, 17, 902–15.CrossRefGoogle ScholarPubMed
Beckage, B.et al. (2008). A rapid upward shift of a forest ecotone during 40 years of warming in the Green Mountains of Vermont. Proceedings of the National Academy of Sciences, USA, 104, 4197–202.Google Scholar
Becker, C.et al. (2011). Spontaneous epigenetic variation in the Arabidopsis thaliana methylome. Nature, 480, 245–9.CrossRefGoogle ScholarPubMed
Beckman, N. G. & Rogers, H. S. (2013). Consequences of seed dispersal for plant recruitment in tropical forests: interactions within the seedscape. Biotropica, 45, 666–81.CrossRefGoogle Scholar
Becqemont, D. (2009). Charles Darwin, 1837–1839: aux sources d'une découverte. Paris: Editions Kimé.Google Scholar
Beddall, B. G. (1957). Historical notes on avian classification. Systematic Zoology, 6, 129–36.CrossRefGoogle Scholar
Beddall, B. G. (1988). Darwin and divergence: the Wallace connection. Journal of the History of Biology, 21, 1–68.CrossRefGoogle Scholar
Beddie, A. D. (1942). Natural root grafts in New Zealand trees. Transactions and Proceedings of the Royal Society of New Zealand, 71, 199–203.Google Scholar
Behe, M. J. (1996). Darwin's black box: the biochemical challenge to evolution. New York: The Free Press.Google Scholar
Beilstein, M. A.et al. (2010). Dated molecular phylogenies indicate a Miocene origin for Arabidopsis thaliana. Proceedings of the National Academy of Sciences, USA, 107, 18724–8.CrossRefGoogle ScholarPubMed
Bell, C. D., Soltis, D. E. & Soltis, P. S. (2010). The age and diversification of the angiosperms re-revisited. American Journal of Botany, 97, 1296–1303.CrossRefGoogle ScholarPubMed
Bell, G. (1982). The masterpiece of nature: the evolution and genetics of sexuality. London: Croom Helm.Google Scholar
Belzer, N. F. & Ownbey, M. (1971). Chromatographic comparison of Tragopogon species and hybrids. American Journal of Botany, 58, 791–802.CrossRefGoogle Scholar
Ben-Ayed, R. I.et al. (2014). Genetic similarity among Tunisian olive cultivars and two unknown feral olive trees estimated through SSR markers. Biochemical Genetics, 52, 258–68.CrossRefGoogle ScholarPubMed
Bendiksby, M.et al. (2011). Allopolyploid origins of the Galeopsis tetraploids – revisiting Müntzing's classical textbook example using molecular tools. New Phytologist, 191, 1150–67.CrossRefGoogle ScholarPubMed
Bengtsson, B. O. (2009). Asex and evolution: a very large-scale overview. In Lost sex: the evolutionary biology of parthenogenesis, ed. Schön, I., Martens, K. & Dijk, P. van, pp. 1–19. Berlin: Springer Publications.Google Scholar
Bennett, J. H. (1983). Natural selection, heredity and eugenics. Oxford: Oxford University Press.Google Scholar
Bennett, K. D., Bhagwat, S. A. & Willis, K. J. (2012). Neotropical refugia. The Holocene, 22, 1207–14.CrossRefGoogle Scholar
Bennett, M. D. (1995). The development and use of genomic in situ hybridization (GISH) as a new tool in plant systematics. In Kew Chromosome Conference IV, ed. Brandham, P. E. & Bennett, M. D., pp.167–83. Richmond, UK: Royal Botanic Gardens, Kew.Google Scholar
Bennett, S. T., Kenton, A. Y. & Bennett, M. D. (1992). Genomic in situ hybridization reveals the allopolyploid nature of Milium montianum (Graminae). Chromosoma, 101, 420–4.CrossRefGoogle Scholar
Benson, L. (1962). Plant taxonomy. New York: Ronald Press.Google Scholar
Bento, M.et al. (2013). Retrotransposons represent the most labile fraction for genomic rearrangements in polyploid plant species. Cytogenetic and Genome Research, DOI:10.1159/000353308CrossRef
Benton, M. J. (2000). Stems, nodes, crown clades, and rank-free lists: is Linnaeus dead?Biological Reviews, 75, 633–48.Google ScholarPubMed
Benton, M. J., Donoghue, P. C. J. & Asher, R. J. (2009). Calibrating and constraining molecular clocks. In The timetree of life, ed. Hedges, S. B. and Kumar, S., pp. 35–86. Oxford: Oxford University Press.Google Scholar
Bergman, B. (1935). Zytologische Studien über sexuelles und asexuelles Hieracium umbellatum. Hereditas, 20, 47–64.Google Scholar
Bergman, B. (1941). Studies on the embryo sac mother cell and its development in Hieracium subg. Archieracium. Svensk botanisk Tidskrift, 35, 1–42.Google Scholar
Berlin, B., Breedlove, D. E. & Raven, P. H. (1974). Principles of Tzeltal plant classification. London & New York: Academic Press.Google Scholar
Berners-Lee, M. & Clark, D. (2013). The burning question: we can't burn half the world's oil, coal and gas, so how do we quit?London: Profile Books.Google Scholar
Berry, P. E., Tobe, H. & Gómez, J. A. (1991). Agamospermy and the loss of distyly in Erythroxylum undulatum (Erythroxylaceae) from Northern Venezuela. American Journal of Botany, 78, 595–600.CrossRefGoogle Scholar
Berry, R. J. (1977). Inheritance and natural history. London: Collins.Google Scholar
Bessey, C. E. (1908). The taxonomic aspect of the species questions. American Naturalist, 42, 218–24.CrossRefGoogle Scholar
Bessey, C. E. (1915). The phylogenetic taxonomy of flowering plants. Annals of Missouri Botanic Garden, 2, 109–64.CrossRefGoogle Scholar
Bhattachayya, M.et al. (1990). The wrinkled-seed character of pea described by Mendel is caused by a transposon-like insertion in a gene encoding starch-branching enzyme. Cell, 60, 115–22.Google Scholar
Bhattachayya, M., Martin, C. & Smith, A. (1993). The importance of starch biosynthesis in the wrinkled seed shape character of peas studied by Mendel. Plant Molecular Biology, 22, 525–31.Google Scholar
Bicknell, R. A. & Koltunow, A. M. (2004). Understanding apomixis: recent advances and remaining conundrums. The Plant Cell, 16, S228–S245.CrossRefGoogle ScholarPubMed
Bicknell, R. A., Lambie, S. C. & Butler, R. C. (2003). Quantification of progeny classes in two facultatively apomictic accessions of Hieracium. Hereditas, 138, 11–20.CrossRefGoogle ScholarPubMed
Billington, H. L. (1991). Effect of population size on genetic variability in a dioecious conifer. Conservation Biology, 5, 115–19.CrossRefGoogle Scholar
Bininda-Emonds, O. R. P., Gittleman, J. L. & Steel, M. A. (2002). The (super)tree of life. Annual Review of Ecology and Systematics, 33, 265–89.CrossRefGoogle Scholar
Birks, H. J. B. (1986). Late-Quaternary biotic changes in terrestrial and lacustrine environments, with particular reference to north-west Europe. In Handbook of Holocene palaeoecology and palaeohydrology, ed. Berglund, E., pp. 3–65. New York: Wiley.Google Scholar
Birks, H. J. B. & Willis, K. J. (2008) Alpine trees and refugia in Europe. Plant Ecology and Diversity, 1, 147–60.CrossRefGoogle Scholar
Bisby, F. A.et al. (2002). Taxonomy, at the click of a mouse. Nature, 418, 367.CrossRefGoogle ScholarPubMed
Bishop, J. A. & Cook, L. M. (1981) Genetic consequences of man made change. London: Academic Press.Google Scholar
Bishop, J. A. & Korn, M. E. (1969). Natural selection and cyanogenesis in White Clover, Trifolium repens. Heredity, 24, 423–30.Google Scholar
Bishop, O. (1971). Statistics for biology. A practical guide for the experimental biologist, edn. London: Longmans.Google Scholar
Bittencourt Júnior, N. S., Gibbs, P. E. & Semir, J. (2003). Histological study of post-pollination events in Spathodea campanulata Beauv. (Bignoniaceae), a species with late-acting self-incompatibility. Annals of Botany, 91, 827–34.Google Scholar
Bittrich, V. & Kadereit, J. (1988). Cytogenetical and geographical aspects of sterility in Lysimachia nummularia. Nordic Journal of Botany, 8, 325–8.CrossRefGoogle Scholar
Blackwell, W. H. (2002). One-hundred-year code déjà vu?Taxon, 51, 151–4.CrossRefGoogle Scholar
Blair, A. W. & Williamson, P. S. (2010). Pollen dispersal in Star Cactus (Astrophytum asterias). Journal of Arid Environments, 74, 525–7.CrossRefGoogle Scholar
Blakeslee, A. F. & Avery, A. G. (1937). Methods of inducing chromosome doubling in plants. Journal of Heredity, 28, 393–411.CrossRefGoogle Scholar
Blaxter, M. L. (2004). The promise of a DNA taxonomy. Philosophical Transactions of the Royal Society of London, B, 359, 669–79.CrossRefGoogle ScholarPubMed
Bleeker, W. & Hurka, H. (2001). Introgressive hybridization in Rorippa (Brassicaceae): gene flow and its consequences in natural and anthropogenic habitats. Molecular Ecology, 10, 2013–22.CrossRefGoogle ScholarPubMed
Bliege Bird, R.et al. (2008). The ‘fire stick farming’ hypothesis: Australian Aboriginal foraging strategies, biodiversity and anthropogenic fire mosaics. Proceedings of the National Academy of Sciences, USA, 105, 14796–801.CrossRefGoogle Scholar
Blois, J. L.et al. (2013). Climate change and the past, present, and future of biotic interactions. Science, 341. 499–504.CrossRefGoogle ScholarPubMed
Blunt, W. (2004). Linnaeus: The Compleat Naturalist. London: Frances Lincoln.Google Scholar
Bøcher, T. W. (1949). Racial divergences in Prunella vulgaris in relation to habitat and climate. New Phytologist, 48, 285–314.CrossRefGoogle Scholar
Bøcher, T. W. (1963). The study of ecotypical variation in relation to experimental morphology. Regnum Vegetabile, 27, 10–16.Google Scholar
Bøcher, T. W. & Larsen, K. (1958). Geographical distribution of initiation of flowering, growth habit and other factors in Holcus lanatus. Botaniska Notiser, 3, 289–300.Google Scholar
Bøcher, T. W. & Lewis, M. C. (1962). Experimental and cytological studies on plant species. 7, Geranium sanguineum. Biologiske Skrifter, 11, 1–25.Google Scholar
Bohm, W. (1979). Methods of studying root systems. Berlin, Heidelberg & New York: Springer-Verlag.CrossRefGoogle Scholar
Bolkhovskikh, Z., Grif, V., Matvejeva, T. & Zakharyeva, O. (1969). Chromosome numbers of flowering plants. Leningrad: Academy of Sciences of the USSR.Google Scholar
Bomblies, K. (2010). Doomed lovers: mechanisms of isolation and incompatibility in plant speciation. Annual Review of Plant Biology, 61, 109–24.CrossRefGoogle Scholar
Bomblies, K. & Weigel, D. (2010). Arabidopsis and relatives as models for the study of genetic and genomic incompatibilities. Philosophical Transactions of the Royal Society of London B, 365, 1815–23.CrossRefGoogle Scholar
Bomblies, K.et al. (2007). Autoimmune response as a mechanism for a Dobzhansky–Muller-type incompatibility syndrome in plants. PLoS Biology, 5, 1962–72.CrossRefGoogle ScholarPubMed
Bone, E. & Farres, A. (2001). Trends and rates of microevolution in plants. Genetica, 112–113, 165–82.Google ScholarPubMed
Bonin, A., Ehrich, D. & Manel, S. (2007). Statistical analysis of amplified fragment length polymorphism data: a toolbox for molecular ecologists and evolutionists. Molecular Ecology, 16, 3737–58.CrossRefGoogle ScholarPubMed
Bonnier, G. (1890). Cultures expérimentales dans les Alpes et les Pyrénées. Revue générale de Botanique, 2, 513–46.Google Scholar
Bonnier, G. (1895). Recherches expérimentales sur l'adaptation des plantes au climat Alpin. Annales des Sciences naturelles (Botanique), 20, 217–360.Google Scholar
Bonnier, G. (1920). Nouvelles observations sur les cultures expérimentales à diverses altitudes et cultures par semis. Revue générale de Botanique, 32, 305–26.Google Scholar
Bordacs, S.et al. (2002). Chloroplast DNA variation of white oaks in northern Balkans and in the Carpathian Basin. Forest Ecology and Management, 156, 197–209.CrossRefGoogle Scholar
Borg, S. J. (1972). Variability of Rhinanthus serotinus (Schönh.) Oborny in relation to environment. Unpublished thesis, Rijksuniversiteit te Groningen.
Borges, L. A., Sobrinho, M. S. & Lopes, A. V. (2009). Phenology, pollination, and breeding system of the threatened tree Caesalpinia echinata Lam. (Fabaceae), and a review of studies on the reproductive biology in the genus. Flora, 204, 111–30.CrossRefGoogle Scholar
Borgström, G. (1939). Formation of cleistogamic and chasmogamic flowers in Wild Violets as a photoperiodic response. Nature, 144, 514–15.CrossRefGoogle Scholar
Bosemark, N. O. (1954). On accessory chromosomes in Festuca pratensis. I. Cytological investigations. Hereditas, 40, 346–76.Google Scholar
Bossdorf, O.et al. (2005). Phenotypic and genetic differentiation in native versus introduced plant populations. Oecologia, 144, 1–11.CrossRefGoogle ScholarPubMed
Boswell Syme, J. T. (ed.) (1866). English botany; or Coloured figures of British plants, edn., vol. 6. London: Hardwicke.Google Scholar
Bothmer, R.et al. (1971). Clonal variation in populations of Anemone nemorosa L. Botaniska Notiser, 124, 505–19.
Bottin, L., Cadre, S. Le, Quilichini, A., Bardin, P., Moret, J.et al. (2007) Re-establishment trials in endangered plants: a review and the example of Arenaria grandiflora, a species on the brink of extinction in the Parisian region (France). Ecoscience, 14, 410–19, DOI:10.2980/1195-6860.CrossRefGoogle Scholar
Bouck, A. C.et al. (2005). Genetic mapping of species boundaries in Louisiana Irises using IRRE retrotransposon display markers. Genetics, 171, 1289–1303.CrossRefGoogle ScholarPubMed
Bowlby, J. (1990). Charles Darwin. A biography. London: Hutchinson.Google Scholar
Bowler, P. (2008). Foreword. In Natural selection and beyond: the intellectual legacy of Alfred Russel Wallace, ed. Smith, C. H. & Beccaloni, G.. Oxford: Oxford University Press.Google Scholar
Bowler, P. J. (1989a). The Mendelian revolution. London: The Athlone Press.Google Scholar
Bowler, P. J. (1989b). Evolution. The history of an idea. Revised edition. Berkeley: University of California Press.Google Scholar
Bowler, P. J. (1990). Charles Darwin; the man and his influence. Oxford: Basil Blackwell.Google Scholar
Bowles, M. L. & Whelan, C. J. (1994). Restoration of endangered species, conceptual issues, planning and implementation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Box, J. F. (1978). R. A. Fisher. The life of a scientist. New York, Chichester, Brisbane & Toronto: Wiley.Google Scholar
Box, M.S.et al. (2008). Floral ontogenetic evidence of repeated speciation via paedomorphosis in subtribe Orchidinae (Orchidaceae). Botanical Journal of the Linnean Society, 157, 429–54.CrossRefGoogle Scholar
Brackman, A. C. (1980). A delicate arrangement: the strange case of Charles Darwin and Alfred Russel Wallace. New York: Times Books.Google Scholar
Bradley, R. S. (1985). Quaternary paleoclimatology: methods of paleoclimatic reconstruction. Boston: Allen & Unwin.Google Scholar
Bradshaw, A. D. (1959a). Population differentiation in Agrostis tenuis Sibth. I. Morphological differentiation. New Phytologist, 58, 208–27.CrossRefGoogle Scholar
Bradshaw, A. D. (1959b). Population differentiation in Agrostis tenuis Sibth. II. The incidence and significance of infection by Epichloë typhina. New Phytologist, 58, 310–15.CrossRefGoogle Scholar
Bradshaw, A. D. (1959c). Studies of variation in bent grass species. II. Variation within Agrostis tenuis. Journal of the Sports Turf Research Institute, 10, 1–7.Google Scholar
Bradshaw, A. D. (1960). Population differentiation in Agrostis tenuis Sibth. III. Populations in varied environments. New Phytologist, 59, 92–103.CrossRefGoogle Scholar
Bradshaw, A. D. (1965). Evolutionary significance of phenotypic plasticity in plants. Advances in Genetics, 13, 115–55.Google Scholar
Bradshaw, A. D. (1976). Pollution and evolution. In Effects of air pollution on plants, ed. Mansfield, T. A., pp. 135–59. London, New York & Melbourne: Cambridge University Press.Google Scholar
Bradshaw, A. D. (1989). Is evolution fettered or free? Transactions of the Botanical Society ofEdinburgh, 45, 303–11.Google Scholar
Bradshaw, A. D. & McNeilly, T. (1981). Evolution and pollution. London: Arnold.Google Scholar
Bradshaw, A. D. & McNeilly, T. (1991). Evolutionary response to global climate change. Annals of Botany, 87 (Suppl.), 5–14.Google Scholar
Bradshaw, C. J. A. (2012). Little left to lose: deforestation and forest degradation in Australia since European colonization. Journal of Plant Ecology. 5, 109–20.CrossRefGoogle Scholar
Bradshaw, M. E. (1963a). Studies on Alchemilla filicaulis Bus., sensu lato and A. minima Walters. Introduction and I. Morphological variation in A. filicaulis, sensu lato. Watsonia, 5, 304–20.Google Scholar
Bradshaw, M. E. (1963b). Studies on Alchemilla filicaulis Bus., sensu lato, and A. minima Walters. II. Cytology of A. filicaulis, sensu lato. Watsonia, 5, 321–6.Google Scholar
Bradshaw, M. E. (1964). Studies on Alchemilla filicaulis Bus., sensu lato and A. minima Walters. III. Alchemilla minima. Watsonia, 6, 76–81.Google Scholar
Brady, K. U., Kruckeberg, A. R. & Bradshaw, H. D. Jr (2005). Evolutionary ecology of plant adaptation to serpentine soils. Annual Review of Ecology, Evolution, and Systematics, 36, 243–66.CrossRefGoogle Scholar
Brand, C. J. & Waldron, L. R. (1910). Cold resistance of Alfalfa and some factors influencing it. U.S. Department of Agriculture, Bureau of Plant Industry. Bulletin 185, 1–80.Google Scholar
Brannigan, A. (1979). The reification of Mendel. Social Studies of Science, 9, 423–54.CrossRefGoogle Scholar
Brauner, S. & Gottlieb, L. D. (1987). A self-compatible plant of Stephanomeria exigua subsp. coronaria (Asteraceae) and its relevance to the origin of its self-pollinating derivative S. malheurensis. Systematic Botany, 12, 299–304.CrossRefGoogle Scholar
Breed, M. F.et al. (2013). Which provenance and where? Seed sourcing strategies for revegetation in a changing environment. Conservation Genetics, 14, 1–10.CrossRefGoogle Scholar
Brehm, B. G. & Ownbey, M. (1965). Variation in chromatographic patterns in the Tragopogon dubius pratensis – porrifolius complex (Compositae). American Journal of Botany, 52, 81, 1–18.CrossRefGoogle Scholar
Breiman, A. & Graur, D. (1995). Wheat evolution. Israel Journal of Plant Sciences, 43, 85–98.CrossRefGoogle Scholar
Breinholt, J. W.et al. (2009). Population structure of an endangered Utah endemic Astragalus ampullaroides (Fabaceae). American Journal of Botany, 96, 661–7.CrossRefGoogle Scholar
Brenchley, W. E. & Warington, K. (1969). The Park Grass plots at Rothamsted, 1856–1949. Harpenden: Rothamsted Experimental Station.Google Scholar
Brennan, A. C. (2014). The genetic structure of Arabidopsis thaliana in the south-western Mediterranean range reveals a shared history between North Africa and southern Europe. BMC Plant Biology, 14:17, DOI:10.1186/1471-2229-14-17.CrossRefGoogle ScholarPubMed
Brennan, A. C.et al. (2011). Sporophytic self-incompatibility in Senecio squalidus (Asteraceae): S allele dominance interactions and modifiers of cross-compatibility and selfing rates. Heredity, 106, 113–23.CrossRefGoogle ScholarPubMed
Brenner, G. J. (1996). Evidence for the earliest stage of angiosperm pollen evolution: a paleoequatorial section from Israel. In Flowering plant origin, evolution and phylogeny, ed. Taylor, D. W. & Hickey, L. J., pp.91–115. New York: Chapman and Hall.Google Scholar
Breshears, D. D.et al. (2005). Regional vegetation die-off in response to global-change-type drought. Proceedings of the National Academy of Sciences, USA, 102, 15144–8.CrossRefGoogle ScholarPubMed
Bretagnolle, F. & Thompson, J. D. (1995). Gametes with the somatic chromosome number: mechanisms of their formation and role in the evolution in autoployploid plants. New Phytologist, 129, 1–22.CrossRefGoogle Scholar
Briand, F.et al. (1989). The Alps. A system under pressure. Chambéry: International Union for Conservation of Nature.Google Scholar
Brickell, C. D.et al. (2008) Do the views of users of taxonomic output count for anything?Taxon, 57, 1047–8.Google Scholar
Briggs, D. (2009). Plant microevolution and conservation in human-influenced ecosystems. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Briggs, D. & Block, M. (1981). An investigation into the use of the ‘-deme’ terminology. New Phytologist, 89, 729–35.CrossRefGoogle Scholar
Briggs, D. & Walters, S. M. (1997). Plant variation and evolution, edn. Cambridge: Cambridge University Press.Google Scholar
Briggs, J. C. (1987). Biogeography and plate tectonics. Amsterdam: Elsevier.Google Scholar
Broadhurst, L. M.et al. (2008). Seed supply for broadscale restoration: maximizing evolutionary potential. Evolutionary Applications, 1, 587–97.Google ScholarPubMed
Brochet, A. L.et al. (2009). The role of migratory ducks in the long-distance dispersal of native plants and the spread of exotic plants in Europe. Ecography, 32, 298–319.Google Scholar
Brochmann, C., Soltis, P. S. & Soltis, D. E. (1992a). Recurrent formation and polyphyly of Nordic polyploids in Draba (Brassicaceae). American Journal of Botany, 79, 673–88.CrossRefGoogle Scholar
Brochmann, C., Soltis, P. S. & Soltis, D. E. (1992b). Multiple origins of the octoploid Scandinavian endemic Draba cacuminum: electrophoretic and morphological evidence. Nordic Journal of Botany, 12, 257–72.CrossRefGoogle Scholar
Brochmann, C.et al. (1998). Molecular evidence for polyploid origins in Saxifraga (Saxifragaceae): the narrow arctic endemic S. svalbardensis and its widespread allies. American Journal of Botany, 85, 135–43.CrossRefGoogle ScholarPubMed
Brochmann, C.et al. (2000). Multiple diploid hybrid speciation of the Canary Island endemic Argyranthemum sundingii (Asteraceae). Plant Systematics and Evolution, 220, 77–92.CrossRefGoogle Scholar
Brochmann, C.et al. (2004). Polyploidy in arctic plants. Biological Journal of the Linnean Society, 82, 521–36.CrossRefGoogle Scholar
Brockway, L. H. (1979). Science and colonial expansion: the role of the British Royal Botanic Gardens. London: Academic Press.Google Scholar
Bronstein, J. L., Alarcón, R. & Geber, M. (2006). The evolution of plant–insect mutualisms. New Phytologist, 172, 412–28.CrossRefGoogle ScholarPubMed
Brooks, J. L. (1983). Just before the Origin: Alfred Russel Wallace's theory of evolution. New York: Columbia University Press.Google Scholar
Brooks, R. R.et al. (1977). Detection of nickeliferous rocks by analysis of herbarium specimens of indicator plants. Journal of Geochemistry Exploration, 7, 49–57.Google Scholar
Brougham, R. W. & Harris, W. (1967). Rapidity and extent of changes in genotypic structure induced by grazing in a Ryegrass population. New Zealand Journal of Agricultural Research, 10, 56–65.CrossRefGoogle Scholar
Brown, A. H. D. (1979). Enzyme polymorphism in plant populations. Theoretical Population Biology, 15, 1–42.CrossRefGoogle Scholar
Brown, A. H. D. & Burdon, J. J. (1983). Multilocus diversity in an outbreeding weed, Echium plantagineum L. Australian Journal of Biological Sciences, 36, 503–9.
Brown, A. H. D. & Marshall, D. R. (1981). Evolutionary changes accompanying colonization in plants. In Evolution today, ed. Sudder, G. C. E. & Reveal, J. L., pp. 351–63. Pittsburgh, PA: Hunt Institute for Botanical Documentation.Google Scholar
Brown, A. D. H. & Schoen, D. J. (1992). Plant population genetic structure and biological conservation. In Conservation of biodiversity for sustainable development, ed. Sandlund, O. T.et al., pp. 88–104. Oslo: Scandinavian University Press.Google Scholar
Brown, V. K. & Lawton, J. H. (1991). Herbivory and the evolution of leaf size and shape. Philosophical Transactions of the Royal Society of London, B, 333, 267–72.CrossRefGoogle Scholar
Browne, J. (1983). The Secular Ark; studies in the history of biogeography. New Haven, CT, & London: Yale University Press.CrossRefGoogle Scholar
Browne, J. (1995). Charles Darwin: Voyaging. Volume 1 of a biography. London: Pimlico. Random House.Google Scholar
Browne, J. (2002). Charles Darwin: The power of place. Volume 2 of a biography. London: Pimlico. Random House.Google Scholar
Brownfield, L. & Köhler, C. (2011).Unreduced gamete formation in plants: mechanisms and prospects. Journal of Experimental Botany, 62, 1659–68.CrossRefGoogle ScholarPubMed
Brucker, R. M. & Bordenstein, S. R. (2013).Speciation by symbiosis. Trends in Ecology & Evolution, 27, 443–51.Google Scholar
Brummitt, R. K. (1996). In defence of paraphyletic taxa. In The biodiversity of African plants, ed. Maesen, L. J. G. van der, Burgt, X. M. van der & Rooy, J. M. van Medenbach de, Proceedings XIVth AETFAT Congress, 22–27 August 1994, Wageningen, pp. 371–84. Dordrecht: Kluwer.Google Scholar
Brummitt, R. K. (1997). Taxonomy versus cladonomy, a fundamental controversy in biological systematics. Taxon, 46, 723–34.CrossRefGoogle Scholar
Brummitt, R. K. (2002). How to chop up a tree. Taxon, 51, 31–41.CrossRefGoogle Scholar
Brummitt, R. K. (2003). Further dogged defense of paraphyletic taxa. Taxon, 52, 803–4.CrossRefGoogle Scholar
Brummitt, R. K. (2006). The democratic processes of plant nomenclature. In Taxonomy and plant conservation: the cornerstone of the conservation and the sustainable use of plants, ed. Leadlay, E. & Jury, S, pp. 101–29. Cambridge: Cambridge University Press.Google Scholar
Brunsfeld, S. J. & Sullivan, J. (2005). A multi-compartmented glacial refugium in the northern Rocky Mountains: evidence from the phylogeography of Cardamine constancei (Brassicaceae). Conservation Genetics, 6, 895–904.Google Scholar
Brus, R. (2010). Growing evidence for the existence of glacial refugia of European beech (Fagus sylvatica L.) in the south-eastern Alps and north-western Dinaric Alps. Periodicum Biologorum, 112, 239–46.Google Scholar
Brush, S. G. (2009). Choosing selection. The revival of natural selection in Anglo-American evolutionary biology, 1930–1970. Transactions of the American Philosophical Society, 99 (3), 1–183. Philadelphia.Google Scholar
Brussard, P. F. (1997). A paradigm in conservation biology. Science, 277, 527–8.CrossRefGoogle Scholar
Bryant, H. N. & Cantino, P. D. (2002). A review of criticisms of phylogenetic nomenclature: is taxonomic freedom the fundamental issue?Biological Review, 77, 39–55.CrossRefGoogle ScholarPubMed
Brysting, A. K., Mathiesen, C. & Marcussen, T. (2011). Challenges in polyploid phylogenetic reconstruction: a case story from the arctic–alpine Cerastium alpinum complex. Taxon, 60, 333–47.Google Scholar
Budiansky, S. (1995). Nature's keepers. the new science of nature management. London: Weidenfeld & Nicolson.Google Scholar
Buggs, R. J. A. (2007). Empirical studies of hybrid zone movement. Heredity, 99, 301–12.CrossRefGoogle ScholarPubMed
Buggs, R. J. A.et al. (2008). Does phylogenetic distance between parental genomes govern the success of polyploids?Castanea, 73, 74–93.CrossRefGoogle Scholar
Buggs, R. J. A., Soltis, P. S. & Soltis, D. E. (2009). Does hybridization between divergent progenitors drive whole-genome duplication?Molecular Ecology, 18, 3334–9.CrossRefGoogle ScholarPubMed
Buggs, R. J. A.et al. (2009). Gene loss and silencing in Tragopogon miscellus (Asteraceae): comparison of natural and synthetic allotetraploids. Heredity, 103, 73–81.CrossRefGoogle ScholarPubMed
Buggs, R. J. A.et al. (2010a). Tissue-specific silencing of homoeologs in natural populations of the recent allopolyploid Tragopogon mirus. New Phytologist, 186, 175–83.CrossRefGoogle ScholarPubMed
Buggs, R. J. A.et al. (2010b). Characterization of duplicate gene evolution in the recent natural allopolyploid Tragopogon miscellus by next-generation sequencing and Sequenom iPLEX MassARRAY genotyping. Molecular Ecology, 19, 132–46.CrossRefGoogle ScholarPubMed
Buggs, R. J. A., Soltis, P. S. & Soltis, D. E. (2011). Biosystematic relationships and the formation of polyploids. Taxon, 60, 324–32.Google Scholar
Buggs, R. J. A.et al. (2011). Transcriptomic shock generates evolutionary novelty in a newly formed, natural allopolyploid plant. Current Biology, 21, 551–6.CrossRefGoogle Scholar
Buggs, R. J. A.et al. (2012). Next-generation sequencing and genome evolution in allopolyploids. American Journal of Botany, 99, 372–82.CrossRefGoogle ScholarPubMed
Bullock, J. M. (1998). Community translocation in Britain: setting objectives and measuring consequences. Biological Conservation, 6, 166–74.Google Scholar
Bulmer, M. (2003) Francis Galton: pioneer of heredity and biometry. Baltimore: John Hopkins University Press.Google Scholar
Bulmer, M. G. (1967). Principles of statistics, edn. London & Edinburgh: Oliver & Boyd.Google Scholar
Burchfield, J. D. (1990). Lord Kelvin and the age of the Earth. Chicago & London: University of Chicago Press.CrossRefGoogle Scholar
Burd, M. (1994). Bateman's principle and plant reproduction: the role of pollen limitation in fruit and seed set. Botanical Review, 60, 83–139.CrossRefGoogle Scholar
Burdon, J. J. (1980). Intraspecific diversity in a natural population of Trifolium repens. Journal of Ecology, 68, 717–35.CrossRefGoogle Scholar
Burdon, J. J. (1987). Disease and plant population biology. Cambridge: Cambridge University Press.Google Scholar
Burdon, J. J., Marshall, D. R. & Groves, R. H. (1980). Isozyme variation in Chondrilla juncea in Australia. Australian Journal of Botany, 28, 193–8.CrossRefGoogle Scholar
Burgess, K. S.et al. (2005). Asymmetrical introgression between two Morus species (M. alba, M. rubra) that differ in abundance. Molecular Ecology, 14, 3471–83.CrossRefGoogle ScholarPubMed
Burkhardt, F. & Smith, S. (1985–). The correspondence of Charles Darwin. Cambridge: Cambridge University Press.Google Scholar
Burkhardt, F. & Smith, S. (1991). The correspondence of Charles Darwin. Volume 7 1858–9. Cambridge: Cambridge University Press.Google Scholar
Burkill, I. H. (1895). On the variations in number of stamens and carpels. Journal of the Linnean Society (Botany), 31, 216–45.Google Scholar
Busch, J.et al. (2012). Climate change and the cost of conserving species in Madagascar. Conservation Biology, 26, 408–19.CrossRefGoogle ScholarPubMed
Bush, E. J. & Barrett, S. C. H. (1993). Genetics of mine invasions by Deschampsia cespitosa (Poaceae). Canadian Journal of Botany – Revue Canadienne de Botanique, 71, 1336–48.Google Scholar
Bush, M. B. & de Oliveira, P. E. (2006). The rise and fall of the Refugial Hypothesis of Amazonian Speciation: a paleo-ecological perspective. Biota Neotropica, 6, 1–17.CrossRefGoogle Scholar
Busi, R.et al. (2013). Herbicide-resistant weeds: from research and knowledge to future needs. Evolutionary Applications, 6, 1218–21.CrossRefGoogle ScholarPubMed
Cahn, M. A. & Harper, J. L. (1976a). The biology of the leaf mark polymorphism in Trifolium repens. 1. Distribution of phenotypes at a local scale. Heredity, 37, 309–25.CrossRefGoogle Scholar
Cahn, M. A. & Harper, J. L. (1976b). The biology of the leaf mark polymorphism in Trifolium repens. 2. Evidence for the selection of marks by rumen fistulated sheep. Heredity, 37, 327–33.CrossRefGoogle Scholar
Cain, A. J. (1958). Logic and memory in Linnaeus's system of taxonomy. Proceedings of the Linnean Society of London, 169, 144–63.CrossRefGoogle Scholar
Cain, A. J. (1996). John Ray on ‘Accidents’. Archives of Natural History, 23, 343–68.CrossRefGoogle Scholar
Cain, A. J. (1999). John Ray on species. Archives of Natural History, 26, 223–38.Google Scholar
Cain, A. J. (2008). The post-Linnaean development of taxonomy. Proceedings of the Linnean Society of London, 170, 234–44.Google Scholar
Cain, J. & Ruse, M. (2009). Descended from Darwin. Insights into the history of evolutionary studies, 1900–1970. Transactions of the American Philosophical Society, 99 (1), 1–386.Google Scholar
Cain, M. L., Milligan, B. G. & Strand, A. E. (2000). Long-distance seed dispersal in plant populations. American Journal of Botany, 87, 1217–27.CrossRefGoogle ScholarPubMed
Cairns, J. (1998). Ecological restoration. In Encyclopedia of ecology & environmental management, ed. Calow, P., pp. 217–19. Oxford: Blackwell.Google Scholar
Cairns, J. (2002). Rationale for restoration. In Handbook of ecological restoration, ed. Perrow, M. R. & Davy, A. J., pp.10–23. Cambridge: Cambridge University Press.Google Scholar
Caisse, M. & Antonovics, J. (1978). Evolution in closely adjacent plant populations. IX. Evolution of reproductive isolation in clinal populations. Heredity, 40, 371–84.CrossRefGoogle Scholar
Callender, L. A. (1988). Gregor Mendel: an opponent of descent with modification. History of Science, 26, 41–75.CrossRefGoogle Scholar
Calsbeek, B.et al. (2011). Comparing the genetic architecture and potential response to selection of invasive and native populations of Reed Canary Grass. Evolutionary Applications, 4, 726–35.CrossRefGoogle ScholarPubMed
Calsbeek, R., Thompson, J. R. & Richardson, J. E. (2003). A phylogenetic analysis of Rhamnaceae using rbcL and trnL-F plastid DNA sequences. Molecular Ecology, 12, 1021–9.Google Scholar
Caltagirone, L. E. (1981). Landmark examples in classical biological control. Annual Review of Entomology, 26, 213–32.CrossRefGoogle Scholar
Camacho, J. P. M. (2005). B chromosomes. In The evolution of the genome, ed. Gregory, T. R., pp. 223–86, London: Elsevier Academic Press.Google Scholar
Camp, W. H. & Gilly, C. L. (1943). The structure and origin of species. Brittonia, 4, 323–85.Google Scholar
Campbell, R. C. (1967). Statistics for biologists. London & New York: Cambridge University Press. [edn: 1974].Google Scholar
Campbell, S. A & Kessler, A. (2013). Plant mating system transitions drive the macroevolution of defence strategies. Proceedings of the National Academy of Sciences, USA, 110, 3973–8.CrossRefGoogle Scholar
Carlo, T. A., Tewksbury, J. J. & del Rio, C. M. (2009). A new method to track seed dispersal and recruitment using N-15 isotope enrichment. Ecology, 90, 3516–25.CrossRefGoogle Scholar
Carlquist, S. (1965). Island life. A natural history of the islands of the world. New York: Natural History Press.Google Scholar
Carlquist, S. (1974). Island biology. New York & London: Columbia University Press.CrossRefGoogle Scholar
Carlson, S., Mayer, V. & Michael, J. (2009). Phylogenetic relationships, taxonomy, and morphological evolution in Dipsacaceae (Dipsacales) inferred by DNA sequence data. Taxon, 58, 1075–91.Google Scholar
Carney, S. E., Cruzan, M. B. & Arnold, M. L. (1994). Reproductive interactions between hybridizing Irises: analyses of pollen-tube growth and fertilization success. American Journal of Botany, 81, 1169–75.CrossRefGoogle Scholar
Caron, G. E. & Leblanc, R. (1992). Pollen contamination on a small black spruce seedling seed orchard for 3 consecutive years. Forest Ecology & Management, 53, 587–92.CrossRefGoogle Scholar
Carr, G. D.et al. (1999). Chromosome numbers in Compositae. XVIII. American Journal of Botany, 86, 1003–13.CrossRefGoogle ScholarPubMed
Carson, R. (1962). Silent spring. London: Hamish Hamilton.Google Scholar
Catcheside, D. G. (1939). A position effect in Oenothera. Journal of Genetics, 38, 345–52.CrossRefGoogle Scholar
Catcheside, D. G. (1947). The P-locus position effect in Oenothera. Journal of Genetics, 48, 31–42.CrossRefGoogle ScholarPubMed
Caujapé-Castells, J. (2004). Boomerangs of biodiversity? The interchange of biodiversity between mainland north Africa and the Canary Islands as inferred from cpDNA RFLPs in genus Androcymbium. Botánica Macaronésica, 25, 53–69.Google Scholar
Caujapé-Castells, J.et al. (2010). Conservation of oceanic island floras: present and future global challenges. Perspectives in Plant Ecology, Evolution and Systematics, 12, 107–29.CrossRefGoogle Scholar
Cayouette, J. E. & Morisset, P. (1986). Chromosome studies on the Carex salina complex (Cyperaceae section Cryptocarpae) in northeastern North America. Cytologia, 51, 817–56.CrossRefGoogle Scholar
Chadwick, M. J. & Salt, J. K. (1969). Population differentiation within Agrostis tenuis L. in response to colliery spoil substrate factors. Nature, 224, 186.CrossRefGoogle Scholar
Chaing, G. C. K.et al. (2009). Major flowering time gene, FLOWERING LOCUS C, regulates seed germination in Arabidopsis thaliana. Proceedings of the National Academy of Sciences, USA, 106, 11661–6.Google Scholar
Chakraborty, R., Meagher, T. R. & Smouse, P. E. (1988). Parentage analysis with genetic markers in natural populations. I. The expected proportion of offspring with unambiguous paternity. Genetics, 118, 527–36.Google ScholarPubMed
Challice, J. & Kovanda, M. (1978). Chemotaxonomic survey of the genus Sorbus in Europe. Naturwissenschaften, 65, 111–12.CrossRefGoogle Scholar
Chandler, S. & Dunwell, J. M. (2008). Gene flow, risk assessment and the environmental release of transgenic plants. Critical Reviews in Plant Sciences, 27, 25–49.CrossRefGoogle Scholar
Chanway, C. P.Holl, F. B. & Turkington, R. (1989). Effect of Rhizobium leguminosarum biovar. trifolii genotype on specificity between Trifolium repens and Lolium perenne. Journal of Ecology, 77, 1150–60.CrossRefGoogle Scholar
Chapman, M. A. & Abbott, R. J. (2009). Introgression of fitness genes across a ploidy barrier. New Phytologist, 186, 63–71.Google ScholarPubMed
Chapman, M. A. & Burke, J. M. (2006). Letting the gene out of the bottle: the population genetics of genetically modified crops. New Phytologist, 170, 429–43.CrossRefGoogle ScholarPubMed
Charlesworth, B. (2009). Effective population size and patterns of molecular evolution and variation. Nature Reviews of Genetics, 10, 195–205.CrossRefGoogle ScholarPubMed
Charlesworth, D. (1985). Distribution of dioecy and self-incompatibility in angiosperms. In Evolution – essays in honour of John Maynard Smith, ed. Greenwood, P. J. & Slatkin, M., pp. 237–68. Cambridge: Cambridge University Press.Google Scholar
Charlesworth, D. (2002). Plant sex determination and sex chromosomes. Heredity, 88, 94–101.CrossRefGoogle ScholarPubMed
Charlesworth, D. (2006). Evolution of plant breeding systems. Current Biology, 16, R726–R735.CrossRefGoogle ScholarPubMed
Charlesworth, D. (2010). Self-incompatibility. F1000 Report Biology, 2,68. Published online, DOI:10.3410/B2-68.CrossRef
Charlesworth, D. & Charlesworth, B. (1979). The evolutionary genetics of sexual systems in flowering plants. Proceedings of the Royal Society London, B, 205, 513–30.CrossRefGoogle ScholarPubMed
Charlesworth, D. & Charlesworth, B. (1987). Inbreeding depression and its evolutionary consequences. Annual Review of Ecology and Systematics, 18, 237–68.CrossRefGoogle Scholar
Charlesworth, D. & Willis, J. H. (2009). The genetics of inbreeding depression. Nature, 10, 783–96.Google ScholarPubMed
Charlesworth, D.et al. (2005). Plant self-incompatibility systems: a molecular evolutionary perspective. New Phytologist, 168, 61–9.CrossRefGoogle ScholarPubMed
Charnov, E. L. (1988). Foreword. In Plant reproductive ecology, ed. Doust, J. Lovett & Doust, L. Lovett, pp. ix–x. New York: Oxford University Press.Google Scholar
Chase, M. W.et al. (1993). Phylogenetics of seed plants: an analysis of nucleotide-sequences from the plastid gene rbcL. Annals of the Missouri Botanical Garden, 80, 528–80.CrossRefGoogle Scholar
Chase, M. W., Fay, M. F. & Savolainen, V. (2000). Higher-level classification in the angiosperms: new insights from the perspective of DNA sequence data. Taxon, 49, 685–704.CrossRefGoogle Scholar
Chase, M. W.et al. (2005). Land plants and DNA barcodes: short-term and long-term goals. Philosophical Transactions of the Royal Society of London, B, 359, 1889–95.Google Scholar
Chase, M. W.et al. (2009). Murderous plants: Victorian Gothic, Darwin and modern insight into vegetable carnivory. Botanical Journal of the Linnean Society, 161, 329–56.CrossRefGoogle Scholar
Chen, J.-Q.et al. (2008). Over-expression of OsDREB genes leads to enhanced drought tolerance in rice. Biotechnology Letters, 30, 2191–8.CrossRefGoogle Scholar
Chen, J.et al. (2012). Disentangling the roles of history and local selection in shaping clinal variation in allele frequencies and gene expression in Norway Spruce (Picea abies). Genetics, 191, 865–81.CrossRefGoogle Scholar
Chen, J.et al. (2014). Clinal variation at phenology-related genes in Spruce; parallel evolution in FTL2 and Gigantea?Genetics, 197, 1025–38.CrossRefGoogle ScholarPubMed
Chen, Q. & Armstrong, K. (1994). Genomic in situ hybridization in Avena sativa. Genome, 37, 607–12.CrossRefGoogle ScholarPubMed
Cheplick, G. P. & Quinn, J. A. (1982). Amphicarpum purshii and the ‘pessimistic strategy’ in amphicarpic annuals with subterranean fruit. Oecologia, 52, 327–32.CrossRefGoogle Scholar
Cheplick, G. P. & Quinn, J. A. (1983). The shift in aerial/subterranean fruit ratio in Amphicarpum purshii: causes and significance. Oecologia, 57, 374–9.CrossRefGoogle ScholarPubMed
Cheptou, P. O. (2012). Clarifying Baker's Law. Annals of Botany, 109, 633–41.CrossRefGoogle ScholarPubMed
Chester, M.et al. (2007). Parentage of endemic Sorbus L. (Rosaceae) species in the British Isles. Botanical Journal of the Linnean Society, 154, 291–304.CrossRefGoogle Scholar
Chester, M.et al. (2010a). Review of the application of modern cytological methods (FISH/GISH) to the study of reticulation (Polyploidy/Hybridisation). Genes (Basel), 1, 166–92, DOI:10.3390/genes1020166.CrossRefGoogle Scholar
Chester, M.et al. (2010b). Extensive chromosomal variation in a recently formed natural allopolyploid species, Tragopogon miscellus (Asteraceae). Proceedings of the National Academy of Sciences, USA, 109, 1176–81.Google Scholar
Choi, Y. D. (2004). Theories for ecological restoration in changing environment: towards ‘futuristic’ restoration. Ecological Research, 19, 75–81.CrossRefGoogle Scholar
Choi, Y. D. (2007). Restoration ecology to the future: a call for a new paradigm. Restoration Ecology, 15, 351–3.CrossRefGoogle Scholar
Chomorro, S.et al. (2012). Pollination patterns and plant breeding systems in the Galápagos: a review. Annals of Botany, 110, 1489–1501.Google Scholar
Chung, M. G., Hamrick, J. L., Jones, S. B. & Derda, G. S. (1991). Isozyme variation within and among populations of Hosta (Liliaceae) in Korea. Systematic Botany, 16, 667–84.CrossRefGoogle Scholar
Church, S. A. & Taylor, D. R. (2007). The evolution of reproductive isolation in spatially structured populations. Evolution, 56, 1859–62.Google Scholar
Cicerone, R. J. & Nurse, P. (2014). Climate change: evidence and causes. US National Academy of Sciences and the Royal Society.Google Scholar
Claessen, D.et al. (2005a). Which traits promote persistence of feral GM crops? Part 1: implications of environmental stochasticity. Oikos, 110, 20–29.Google Scholar
Claessen, D.et al. (2005b). Which traits promote persistence of feral GM crops? Part 2: implications of metapopulation structure. Oikos, 110, 30–42.Google Scholar
Clapham, A. R., Tutin, T. G. & Warburg, E. F. (1981). Excursion flora of the British Isles. London: Cambridge University Press.Google Scholar
Clark, B. R.et al. (2009). Taxonomy as an escience. Philosophical Transactions of the Royal Society of London, B, 367, 953–66.Google ScholarPubMed
Clark, D. (2005). Molecular biology. Burlington, MA, & London: Elsevier.Google Scholar
Clark, L. V. & Jasieniuk, M. (2012). Spontaneous hybrids between native and exotic Rubus in the Western United States produce offspring both by apomixis and by sexual recombination. Heredity, 109, 320–8.CrossRefGoogle ScholarPubMed
Clark, P.et al. (2009). The Last Glacial Maximum. Science, 325, 710–14.CrossRefGoogle ScholarPubMed
Clarke, G. M. (1980). Statistics and experimental design. London: Edward Arnold.Google Scholar
Clarke, J. T., Warnock, R. C. M. & Donoghue, P. C. J. (2011). Establishing a timescale for plant evolution. New Phytologist, 192, 266–301.CrossRefGoogle Scholar
Clausen, J. (1951). Stages in the evolution of plant species. London: Oxford University Press; Ithaca, NY: Cornell University Press.Google Scholar
Clausen, J. & Hiesey, W. M. (1958). Experimental studies on the nature of species. IV. Genetic structure of ecological races. Carnegie Institution of Washington Publication No. 615, Washington DC.Google Scholar
Clausen, J., Keck, D. D. & Hiesey, W. M. (1939). The concept of species based on experiment. American Journal of Botany, 26, 103–6.CrossRefGoogle Scholar
Clausen, J., Keck, D. D. & Hiesey, W. M. (1940). Experimental studies on the nature of species. I. The effect of varied environments on Western North American plants. Carnegie Institution of Washington Publication No. 520, pp. 1–452, Washington DC.Google Scholar
Clausen, J., Keck, D. D. & Hiesey, W. M. (1941). Experimental taxonomy. Carnegie Institution of Washington Year Book, 40, 160–70.Google Scholar
Clausen, J., Keck, D. D. & Hiesey, W. M. (1945). Experimental studies on the nature of species. 11. Plant evolution through amphiploidy and autoploidy, with examples from the Madiinae. Carnegie Institute of Washington Publication 564. Washington DC.Google Scholar
Clausen, R. E. & Goodspeed, T. H. (1925). Interspecific hybridization in Nicotiana. II. A tetraploid glutinosa-tabacum hybrid, an experimental verification of Winge's hypothesis. Genetics, 10, 279–84.Google Scholar
Clements, D. R. & Ditommosa, A. (2011). Climate change and weed adaptation: can evolution of invasive plants lead to greater range expansion than forecasted?Weed Research, 51, 227–40.CrossRefGoogle Scholar
Clements, E. J. & Foster, M. C. (1994). Alien plants in the British Isles. London: Botanical Society of the British Isles.Google Scholar
Clements, F. E., Martin, E. V. & Long, F. L. (1950). Adaptation and origin in the plant world. The role of environment in evolution. Waltham, MA: Chronica Britanica Co.Google Scholar
Coates, D. J. & Byrne, M. (2005). Genetic variation in plant populations: assessing cause and pattern. In Plant diversity and evolution: genotypic and phenotypic variation in higher plants, ed. Henry, R. J., pp.139–64. Cambridge, MA, & Wallingford, UK: CABI Publishing.Google Scholar
Coates, D. J, Williams, M. R. & Madden, S. (2013). Temporal and spatial mating-system variation in fragmented populations of Banksia cuneata, a rare bird-pollinated long-lived plant. Australian Journal of Botany, 61, 235–42.CrossRefGoogle Scholar
Cochran, W. G. (1963). Sampling techniques, edn. New York: Wiley.Google Scholar
Cock, A. & Forsdyke, D. R. (2008). Treasure your exceptions. The science and life of William Bateson. New York: Springer-Verlag Inc.CrossRefGoogle Scholar
Cockburn, A. (1991). An introduction to evolutionary ecology. Oxford: Blackwell.Google Scholar
Cody, S.et al. (2010). The great American biotic interchange revisited. Ecography, 33, 326–32.Google Scholar
Coen, E. S. (1991). The role of homeotic genes in flower development and evolution. Annual Review of Plant Physiology and Plant Molecular Biology, 42, 241–79.CrossRefGoogle Scholar
Coen, E. S. & Meyerwitz, E. M. (1991). The war of the whorls: genetic interactions controlling flower development. Nature, 353, 31–7.CrossRefGoogle ScholarPubMed
Coen, E. S. & Nugent, J. (1994). Evolution of flowers and inflorescences. Development, 107 (Suppl.), 107–16.Google Scholar
Cogoni, D.et al. (2013). The effectiveness of plant conservation measures: the Dianthus morisianus reintroduction. Oryx, 47, 203–6.CrossRefGoogle Scholar
Colautti, R., Lee, C. R. & Mitchell-Olds, T. (2012). Origin, fate, and architecture of ecologically relevant genetic variation. Current Opinion in Plant Biology, 15, 199–204.CrossRefGoogle ScholarPubMed
Colautti, R. I. & Barrett, S. C. H. (2011). Population divergence along lines of genetic variance and covariance in the invasive plant Lythrum salicaria in eastern North America. Evolution, 65, 2514–29.CrossRefGoogle ScholarPubMed
Colautti, R. I., Maron, J. L. & Barrett, S. C. H. (2009). Common garden comparisons of native and introduced plant populations: latitudinal clines can obscure evolutionary inferences. Evolutionary Applications, 2, 187–99.CrossRefGoogle ScholarPubMed
Colautti, R. I., White, N. A. & Barrett, S. C. H. (2010). Variation of self-incompatibility within invasive populations of purple loosestrife (Lythrum salicaria L.) from eastern North America. International Journal of Plant Sciences, 171, 158–66.CrossRefGoogle Scholar
Colautti, R. I., Eckert, C. G & Barrett, S. C. H. (2010). Evolutionary constraints on adaptive evolution during range expansion in an invasive plant. Proceedings of the Royal Society of London, B, 277, 1799–1806.CrossRefGoogle Scholar
Colbach, N. & Sache, I. (2001). Blackgrass (Alopecurus myosuroides Huds.) seed dispersal from a single plant and its consequences on weed infestation. Ecological Modelling, 139, 201–19.CrossRefGoogle Scholar
Cole, C. T. (2003). Genetic variation in rare and common plants. Annual Review of Ecology, Evolution and Systematics, 34, 213–37.CrossRefGoogle Scholar
Collen, B., Purvis, A. & Mace, G. M. (2010). When is a species really extinct? Testing extinction inference from a sighting record to inform conservation assessment. Diversity and Distributions, 16, 755–64.CrossRefGoogle Scholar
Collins, J. L. (1927). A low temperature type of albinism in Barley. Journal of Heredity, 33, 82–6.Google Scholar
Coltman, D. W. (2008). Molecular ecological approaches to studying the evolutionary impact of selective harvesting in wildlife. Molecular Ecology, 17, 221–35.CrossRefGoogle ScholarPubMed
Comai, L. (2005). The advantages and disadvantages of being polyploid. Nature Reviews Genetics, 6, 836–46.CrossRefGoogle ScholarPubMed
Comes, H. P. & Kadereit, J. W. (1998). The effect of Quaternary climatic changes on plant distribution and evolution. Trends in Plant Science, 3, 432–8.CrossRefGoogle Scholar
Constanza, R.et al. (1997). The value of the World's ecosystem services and natural capital. Nature, 387, 253–60.Google Scholar
Cook, C. D. K. (1968). Phenotypic plasticity with particular reference to three amphibious plant species. In Modern methods in plant taxonomy, ed. Heywood, V. H., pp.97–111. London: Academic Press.Google Scholar
Cook, J.et al. (2013). Quantifying the consensus on anthropogenic global warming in the scientific literature. Environmental Research Letters, 8, 1–7.CrossRefGoogle Scholar
Cook, L. M.et al. (1998). Multiple independent formations of Tragopogon tetraploids (Asteraceae): evidence from RAPD markers. Molecular Ecology, 7, 1293–1302.CrossRefGoogle Scholar
Cook, S. A. (1962). Genetic system, variation and adaptation in Eschscholzia californica. Evolution, 16, 278–99.CrossRefGoogle Scholar
Cook, S. A. & Johnson, M. P. (1968). Adaptation to heterogeneous environments. I. Variation in heterophylly in Ranunculus flammula L. Evolution, 22, 496–516.Google ScholarPubMed
Cooke, T.J. (2006). Do Fibonacci numbers reveal the involvement of geometric imperatives or biological interactions in phyllotaxis?Botanical Journal of the Linnean Society, 150, 3–24.CrossRefGoogle Scholar
Corcos, A. F. & Monaghan, F. V. (1990). Mendel's work and its rediscovery: a new perspective. Critical Reviews in Plant Sciences, 9, 197–212.CrossRefGoogle Scholar
Corkhill, L. (1942). Cyanogenesis in white clover (Trifolium repens L.) V. The inheritance of cyanogenesis. New Zealand Journal of Science & Technology, B, 23, 178–93.Google Scholar
Corrado, P. & Magri, D. (2011). A late Early Pleistocene pollen record from Fontana Ranuccio (central Italy). Journal of Quaternary Science, 26, 335–44.CrossRefGoogle Scholar
Correns, C. (1909). Vererbungsversuche mit blass (gelb) grunen und buntblättrigen Sippen bei Mirabilis jalapa, Urtica pilulifera, und Lunularia annua. Zeitschrift für Vererbungslehre, 1, 291–329.Google Scholar
Correns, C. (1913). Selbststerilität und Individualstoffe. Biologisches Zentralblatt, 33, 389–443.Google Scholar
Corsi, P. (1988). The age of Lamarck: evolutionary theories in France 1790–1830. Berkeley, Los Angeles & London: University of California Press.Google Scholar
Costello, M. J., May, R. M. & Stork, N. E. (2013). Can we name Earth's species before they go extinct?Science, 339, 413–16.CrossRefGoogle ScholarPubMed
Cott, H. B. (1940). Adaptive coloration in animals. London: Methuen.Google Scholar
Coughtrey, P. J. & Martin, M. H. (1978). Tolerance of Holcus lanatus to lead, zinc and cadmium in factorial combination. New Phytologist, 81, 147–54.CrossRefGoogle Scholar
Cousens, R. & Mortimer, M. (1995). Dynamics of weed populations. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Cowan, R. S. (1972). Francis Galton's statistical ideas: the influence of eugenics. Isis, 63, 509–28.CrossRefGoogle ScholarPubMed
Cox, A. V.et al. (1998). Genome size and karyotype evolution in the Slipper Orchids (Cypripedioideae: Orchidaceae). American Journal of Botany, 85, 681–7.CrossRefGoogle Scholar
Cox, G. W. (1999). Alien species in North America and Hawaii. Washington DC: Island Press.Google Scholar
Cox, G. W. (2004). Alien species and evolution: the evolutionary ecology of exotic plants, animals, microbes, and interacting native species. Washington DC: Island Press.Google Scholar
Cox, P. A. (1988). Monomorphic and dimorphic sexual strategies: a modular approach. In Plant reproductive ecology, ed. Lovett Doust, J. & Doust, L. Lovett, pp. 80–97. New York: Oxford University Press.Google Scholar
Cox-Foster, D. L.et al. (2007). A metagenomic survey of microbes in honey bee colony collapse disorder. Science, 318, 283–7.CrossRefGoogle ScholarPubMed
Coyne, J. A. (1994). Ernst Mayr and the origin of species. Evolution, 48, 19–30.CrossRefGoogle Scholar
Coyne, J. A. (1994). Recognising species. Nature, 364, 298.CrossRefGoogle Scholar
Coyne, J. A. & Lande, R. (1985). The genetic basis of species differences in plants. The American Naturalist, 126, 141–5.CrossRefGoogle Scholar
Coyne, J. A. & Orr, H. A. (2004). Speciation. Sunderland, MA: Sinauer Associates.Google Scholar
Cracraft, J. (1983). Species concepts and speciation analysis. Current Ornithology, 1, 159–87.Google Scholar
Craig, N. L.et al. (2010). Molecular biology: principles of genome function. Oxford & New York: Oxford University Press.Google Scholar
Craig, P. (2005). Centennial history of the Carnegie Institution of Washington. Vol. IV. The Department of Plant Biology. Cambridge: Cambridge University Press.Google Scholar
Crawford, D. J. (1989). Enzyme electrophoresis and plant systematics. In Isozymes in plant biology, ed. Soltis, D. E. & Soltis, P. S., pp. 146–64. London: Chapman & Hall.Google Scholar
Crawford, D. J. (1990). Plant molecular systematics. New York: Wiley.Google Scholar
Crawford, D. J. & Smith, E. B. (1982). Allozyme variation in Coreopsis nuecensoides and C. nucensis (Compositae), a progenitor-derivative species pair. Evolution, 36, 379–86.CrossRefGoogle Scholar
Crawford, D. J., Ornduff, R. & Vasey, M. C. (1985). Allozyme variation within and between Lasthenia minor and its derivative species, Lasthenia maritima (Asteraceae). American Journal of Botany, 72, 1177–84.CrossRefGoogle Scholar
Crawford, D. J.et al. (2010). Mixed mating in the ‘obligately outcrossing’ Tolpis (Asteraceae) of the Canary Islands. Plant Species Biology, 25, 114–19.CrossRefGoogle Scholar
Crawford, D. J.et al. (2014). Contemporary and future studies in plant speciation, morphological/floral evolution and polyploidy: honouring the scientific contributions of Leslie D. Gottlieb to plant evolutionary biology. Philosophical Transactions of the Royal Society, B, 369, http://rstb.royalsocietypublishing.org/content/369/1648/20130341.full.html#ref-list-1CrossRefGoogle Scholar
Crawford, R. M. M. (1985). Studies in plant survival. Oxford: Blackwell.Google Scholar
Crawford, R. M. M. (2014). Gaps in maps: disjunctions in European plant distributions. New Journal of Botany, 4, 64–75.CrossRefGoogle Scholar
Crawford, T. J. (1984). What is a population? In Evolutionary ecology, ed. Sharrocks, B., pp. 135–73. Oxford: Blackwell.Google Scholar
Crawford, T. J. & Jones, D. A. (1986). Variation in the colour of the keel petals in Lotus corniculatus L., 2. Clines in Yorkshire and adjacent counties. Watsonia, 16, 15–19.Google Scholar
Crawford-Sidebotham, T. J. (1971). Studies of aspects of slug behaviour and the relation between molluscs and cyanogenic plants. Unpublished PhD thesis, University of Birmingham.
Cressey, D. (2010). Linnaeus meets the Internet. Nature online 5 May 2010, DOI:10.1038/news.2010.221.CrossRef
Crew, F. A. E. (1966). Mendelism comes to England. In G. Mendel Memorial Symposium. 1865–1965, ed. Sosna, M., pp. 15–30. Prague: Academia Publishing House of the Czechoslovak Academy of Sciences.Google Scholar
Crick, F. (1988). What mad pursuit: a personal view of scientific discovery. London: Weidenfeld & Nicolson.Google Scholar
Crofts, A. & Jefferson, R. G. (1994). The lowland grassland management handbook. English Nature/The Wildlife Trusts.Google Scholar
Cronin, T. N. (2010). Paleoclimates: understanding climate change past and present. New York: Columbia University Press.Google Scholar
Cronk, Q. C. B. (2009) The molecular organography of plants. Oxford: Oxford University Press.CrossRefGoogle Scholar
Cronk, Q. C. B. & Ojeda, I. (2008). Bird-pollinated flowers in an evolutionary and molecular context. Journal of Experimental Botany, 59, 715–27.CrossRefGoogle Scholar
Cronk, Q. C. B., Bateman, R. M. and Hawkins, J. A. (2002). Developmental genetics and plant evolution. London: Taylor & Francis.Google Scholar
Cronquist, A. (1987). A botanical critique of cladism. Botanical Review, 53, 1–52.CrossRefGoogle Scholar
Cronquist, A. (1988). The evolution and classification of flowering plants, edn. New York: New York Botanic Garden.Google Scholar
Crooks, J. A. & Soulé, M. E. (1999). Lag times in population explosions of invasive species: causes and implications. In Invasive species and biodiversity management, ed. Sundlund, O. T., Schei, P. J. & Viken, Å., pp.103–25. Dordrecht, Boston & London: Kluwer.Google Scholar
Crosby, A. W. (1986). Ecological imperialism: the biological expansion of Europe, 900–1900. Cambridge: Cambridge University Press.Google Scholar
Crouzet, P. & Hohn, B. (2007). Transgenic plants. In Handbook of plant science, ed. Roberts, K., pp. 612–18. Chichester: Wiley.Google Scholar
Cruden, R. W. & Hermann-Parker, S. M. (1977). Temporal dioecism: an alternative to dioecism?Evolution, 31, 863–6.Google ScholarPubMed
Cullen, J. & Walters, S. M. (2006). Flowering plant families: how many do we need? In Taxonomy and plant conservation, ed. Leadley, E. & Jury, S. L., pp. 45–90. Cambridge: Cambridge University Press.Google Scholar
Culley, T. M. & Klooster, M. R. (2007). The cleistogamous breeding system: a review of its frequency, evolution and ecology in angiosperms. The Botanical Review, 73, 1–30.CrossRefGoogle Scholar
Curtis, O. F. & Clark, D. G. (1950). An introduction to plant physiology. London, New York & Toronto: McGraw-Hill.Google Scholar
d'Erfurth, I.et al. (2008). Mutations in AtPS1 (Arabidopsis thaliana Parallel Spindle 1) lead to the production of diploid pollen grains. PLoS Genet, 4: e1000274, DOI:10.1371/journal.pgen.1000274.CrossRefGoogle ScholarPubMed
Da Silva, J. M.et al. (2012). Population genetics and conservation of critically small cycad populations: a case study of the Albany Cycad, Encephalartos latifrons (Lehmann). Biological Journal of the Linnean Society, 105, 293–308.CrossRefGoogle Scholar
Daday, H. (1954a). Gene frequencies in wild populations of Trifolium repens. I. Distribution by latitude. Heredity, 8, 61–78.Google Scholar
Daday, H. (1954b). Gene frequencies in wild populations of Trifolium repens. II. Distribution by altitude. Heredity, 8, 377–84.Google Scholar
Daday, H. (1958). Gene frequencies in wild populations of Trifolium repens L. III. World distribution. Heredity, 12, 169–84.CrossRefGoogle Scholar
Daday, H. (1965). Gene frequencies in wild populations of Trifolium repens L. IV. Mechanism of natural selection. Heredity, 20, 355–65.CrossRefGoogle Scholar
Dafni, A. (1992). Pollination ecology: a practical approach. Oxford: Oxford University Press.Google Scholar
Dahlgren, G. (1987). An updated angiosperm classification. Botanical Journal of the Linnean Society, 100, 197–203.Google Scholar
Dahlgren, K. V. O. (1922). Selbststerilität interhalb Klonen von Lysimachia nummularia. Hereditas, 3, 200–10.Google Scholar
Dalgleish, H. J. & Swihart, R. K. (2012). American chestnut past and future: implications of restoration for resource pulses and consumer populations in the eastern U.S. Restoration Ecology, 20, 490–7.CrossRefGoogle Scholar
Darimont, C. T.et al. (2009). Human predators outpace other agents of trait change in the wild. Proceedings of the National Academy of Sciences, USA, 106, 952–4.CrossRefGoogle ScholarPubMed
Darlington, C. D. (1937). Recent advances in cytology, edn. London: Churchill.Google Scholar
Darlington, C. D. (1939). The evolution of genetic systems. Cambridge: Cambridge University Press.Google Scholar
Darlington, C. D. (1956). Chromosome botany. London: Allen & Unwin.Google Scholar
Darlington, C. D. (1963). Chromosome botany and the origins of cultivated plants. London: Allen & Unwin.Google Scholar
Darlington, C. D. & Mather, K. (1949). Elements of genetics. London: Allen & Unwin.Google Scholar
Darlington, C. D. & Wylie, A. P. (1955). Chromosome atlas of flowering plants, edn. London: Allen & Unwin.Google Scholar
Darmency, H. & Gasquez, J. (1997). Spontaneous hybridization of the putative ancestors of the allotetraploid Poa annua. New Phytologist, 136, 497–501.CrossRefGoogle Scholar
Darwin, C. (1859). On the origin of species by means of natural selection, edn. London: Murray. [6th edn: 1872]Google Scholar
Darwin, C. (1862). On the various contrivances by which British and foreign orchids are fertilised by insects and on the good effects of crossing. London: Murray.Google Scholar
Darwin, C. (1868). The variation of plants and animals under domestication. London: Murray.Google Scholar
Darwin, C. (1871a). Pangenesis. Nature, 3, 502–3.CrossRefGoogle Scholar
Darwin, C. (1871b). The descent of man, and selection in relation to sex. Part 2, ed. Barrett, P. H. & Freeman, R. B. (1989). London: William Pickering.Google Scholar
Darwin, C. (1875). Insectivorous plants. London: John Murray.CrossRefGoogle Scholar
Darwin, C. (1876). The effects of cross- and self-fertilisation in the vegetable kingdom. London: Murray.CrossRefGoogle Scholar
Darwin, C. (1877a). The different forms of flowers of the same species. London: Murray.CrossRefGoogle Scholar
Darwin, C. (1877b). The various contrivances by which orchids are fertilised by insects, edn. London: Murray.Google Scholar
Darwin, C. & Wallace, A. (1858). On the tendency of species to form varieties; and on the perpetuation of varieties and species by natural means of selection. Proceedings of the Linnean Society of London, 3, 45–62.Google Scholar
Darwin, F. (ed.) (1909a) The foundations of the origin of species. A sketch written in 1842 by Charles Darwin. Cambridge: Cambridge University Press.Google Scholar
Darwin, F. (ed.) (1909b) The foundations of the origin of species. Two essays written in 1842 and 1844 by Charles Darwin. London: Cambridge University Press.Google Scholar
Darwin, F. & Seward, A. C. (1903). More letters of Charles Darwin, 2 vols. London: Murray.Google Scholar
Davenport, C. B. (1904). Statistical methods with special reference to biological variation, edn. London: Chapman & Hall; New York: Wiley.Google Scholar
David, F. N. (1971). A first course in statistics, edn. London: Griffin.Google Scholar
Davidson, J. F. (1947). The polygonal graph for simultaneous portrayal of several variables in population analysis. Madroño, 9, 105–10.Google Scholar
Davies, H. M. (2010). Review article: Commercialization of whole-plant systems for biomanufacturing of protein products: evolution and prospects. Plant Biotechnology Journal, 8, 845–61.CrossRefGoogle ScholarPubMed
Davies, M. S. (1975). Physiological differences among populations of Anthoxanthum odoratum collected from the Park Grass experiment. IV. Response to potassium and magnesium. Journal of Applied Ecology, 12, 953–64.CrossRefGoogle Scholar
Davies, M. S. (1993). Rapid evolution in plant populations. In Evolutionary patterns and processes, ed. Lees, D. R. & Edwards, D.. Linnean Society Symposium Series, 14, pp. 172–88. London: Published for the Linnean Society by Academic Press.Google Scholar
Davies, M. S. & Snaydon, R. W. (1973a). Physiological differences among populations of Anthoxanthum odoratum collected from the Park Grass experiment. I. Response to calcium. Journal of Applied Ecology, 10, 33–45.Google Scholar
Davies, M. S. & Snaydon, R. W. (1973b). Physiological differences among populations of Anthoxanthum odoratum collected from the Park Grass experiment. II. Response to aluminium. Journal of Applied Ecology, 10, 47–55.Google Scholar
Davies, M. S. & Snaydon, R. W. (1974). Physiological differences among populations of Anthoxanthum odoratum collected from the Park Grass experiment. III. Response to phosphate. Journal of Applied Ecology, 11, 699–707.CrossRefGoogle Scholar
Davies, M. S. & Snaydon, R. W. (1976). Rapid population differentiation in a mosaic environment. III. Measures of selection pressures. Heredity, 36, 59–66.CrossRefGoogle Scholar
Davies, R. (2008). The Darwin conspiracy. origins of a scientific crime. London: Golden Square Books.Google Scholar
Davies, T. M. & Snaydon, R. W. (1989). An assessment of the spaced-plant trial technique. Heredity, 63, 37–45.CrossRefGoogle Scholar
Davies, W. E. (1963). Leaf markings in Trifolium repens. In Teaching genetics in school and university, ed. Darlington, C. D. & Bradshaw, A. D., pp. 94–8. Edinburgh: Oliver & Boyd.Google Scholar
Davis, H.et al. (2004) An Allee effect at the front of a plant invasion: Spartina in a Pacific estuary. Journal of Ecology, 92, 321–7.CrossRefGoogle Scholar
Davis, J. I. (1995). Species concepts and phylogenetic analysis – introduction. Systematic Botany, 20, 555–9.Google Scholar
Davis, M. B. & Shaw, R. G. (2001). Range shifts and adaptive responses to Quaternary climate change. Science, 292, 673–9.CrossRefGoogle ScholarPubMed
Davis, P. H. & Heywood, V. H. (1963). Principles of angiosperm taxonomy. Edinburgh: Oliver & Boyd; New York: Van Nostrand.Google Scholar
Davison, A. W. & Reiling, K. (1995). A rapid change in ozone resistance of Plantago major after summers with high ozone concentrations. New Phytologist, 131, 337–44.CrossRefGoogle Scholar
Davy, A. J. & Jeffries, R. L. (1981). Approaches to the monitoring of rare plant populations. In The biological aspects of rare plant conservation, ed. Synge, H., pp. 219–32. London: Wiley.Google Scholar
Dawkins, R. (2003). A devil's chaplain. London: Weidenfeld & Nicolson.Google Scholar
Dawkins, R. (2005). The ancestor's tale: a pilgrimage to the dawn of life. London: Weidenfeld & Nicolson.Google Scholar
Dawkins, R. (2009). The greatest show on Earth: the evidence for evolution. London: Bantam Press.Google Scholar
Dawson, C. D. R. (1941). Tetrasomic inheritance in Lotus corniculatus L. Journal of Genetics, 42, 49–72.
De Beer, G. (ed.) (1960–1). Darwin's notebooks on transmutation of species I–IV. Bulletin of British Museum (Natural History), Historical Series 2, Nos. 2–6.
De Beer, G. (1963). Charles Darwin. London: Nelson.Google Scholar
De Beer, G. (1964). Atlas of evolution. London: Nelson.Google Scholar
De Bodt, S., Maere, S. & Van de Peer, Y. (2005). Genome duplication and the origin of angiosperms. Trends in Ecology and Evolution, 20, 591–7.Google ScholarPubMed
De Haan, A.et al. (1992). Production of 2n gametes in diploid subspecies of Dactylis glomerata L. 2. Occurrence and frequency of 2n eggs. Annals of Botany, 69, 345–50.Google Scholar
De Nettancourt, D. (1977). Incompatibility in angiosperms. Berlin, Heidelberg & New York: Springer Verlag.CrossRefGoogle Scholar
De Pamphilis, C. W. & Palmer, J. D. (1990) Loss of photosynthetic and chlororespiratory genes from the plastid genome of a parasitic plant. Nature, 348, 337–9.Google Scholar
De Queiroz, A. (2005a). The resurrection of oceanic dispersal in historical biogeography. Annual Review of Ecology and Systematics, 26, 373–401.Google Scholar
De Queiroz, A. (2005b). The resurrection of oceanic dispersal in historical biogeography. Trends in Ecology and Evolution, 20, 68–73.CrossRefGoogle ScholarPubMed
de Queiroz, K. & Donoghue, M. J. (2013). Phylogenetic nomenclature, hierarchical information, and testability. Systematic Biology, 62, 167–74.CrossRefGoogle ScholarPubMed
de Queiroz, K. & Gauthier, J. (1994). Toward a phylogenetic system of biological nomenclature. Trends in Research in Ecology and Evolution, 9, 27–31.Google Scholar
de Vilmorin, P. (1910). Recherches sur l'héredité Mendélienne. Compte Rendu Hebdomadaire des Séances de l Académie des Sciences, Paris, 151, 548–51.Google Scholar
de Vilmorin, P. (1911). Etude sur la caractère adhérence des grains entre eux chez ‘le Pois, Chenille’. 4th International Conference on Genetics, Paris, 368–72.Google Scholar
de Vilmorin, P. & Bateson, W. (1911). A case of gametic coupling in Pisum. Proceedings of the Royal Society, B, 84, 9–11.CrossRefGoogle Scholar
De Vries, H. (1894). Uber halbe Galton-Kurven als Zeichnen diskontinurlichen Variation. Bericht der Deutschen Botanischen Gesellschaft, 12, 197–207.Google Scholar
De Vries, H. (1897). Monstruosités héréditaires offertes en échange aux jardins botaniques. Botanisch Jaarboek, 9, 80–93.Google Scholar
De Vries, H. (1905). Species and varieties: their origin by mutation. Chicago: Open Court Publishing Co.Google Scholar
de Wet, J. M. J. (1968), Diploid–tetraploid–haploid cycles and the origin of variability in Dichanthium agamospecies. Evolution, 22, 394–7.CrossRefGoogle ScholarPubMed
de Wet, J. M. J. (1971). Reversible tetraploidy as an evolutionary mechanism. Evolution, 25, 545–8.CrossRefGoogle ScholarPubMed
de Wet, J. M. J. (1980). Origins of polyploids. In Polyploidy, ed. Lewis, W. H., pp. 3–15. New York & London: Plenum Press.Google Scholar
de Wet, J. M. J. & Harlan, J. R. (1970). Apomixis, polyploidy and speciation in Dichanthium. Evolution, 24, 270–7.CrossRefGoogle ScholarPubMed
de Wet, J. M. J. & Harlan, J. R. (1972). Chromosome pairing and phylogenetic affinities. Taxon, 21, 67–70.CrossRefGoogle Scholar
Witte, L. C. De & Stöcklin, J. (2010). Longevity of clonal plants: why it matters and how to measure it. Annals of Botany, 106, 849–57.CrossRefGoogle Scholar
Decruse, S. W.et al. (2003). Micropropagation and ecorestoration of Vanda spathulata, an exquisite orchid. Plant Cell, Tissue & Organ Culture, 72, 199–202.Google Scholar
Demauro, M. M. (1993). Relationship of breeding system to rarity in the Lakeside Daisy (Hymenoxys acaulis var. glabra). Conservation Biology, 7, 542–50.CrossRefGoogle Scholar
Demauro, M. M. (1994). Development and implementation of a recovery program for the federal threatened Lakeside Daisy (Hymenoxys acaulis var. glabra). In Restoration of endangered species, ed. Bowles, M. L. & Whelan, C. J., pp. 298–312. Cambridge: Cambridge University Press.Google Scholar
Dembski, W. A. & Ruse, M. (eds.) (2007a). Debating design. From Darwin to DNA. Cambridge: Cambridge University Press.Google Scholar
Dembski, W. A. & Ruse, M. (2007b). General introduction. In Debating design. From Darwin to DNA, ed. Dembski, W. A. & Ruse, M., pp. 3–12. Cambridge: Cambridge University Press.Google Scholar
Dennis, A. J., Green, R. J. & Schupp, E. W. (2007). Seed dispersal: theory and its application in a changing world. Egham, Surrey: CABI.CrossRefGoogle Scholar
Depew, D. J. & Weber, B. H. (1996). Darwinism evolving: systems dynamics and the genealogy of natural selection. Cambridge, MA: MIT Press, A Bradford Book.Google Scholar
Des Marais, D. L. & Rauscher, M. D. (2010). Parallel evolution at multiple levels in the origin of hummingbird pollinated flowers of Ipomoea. Evolution, 64, 2044–54.Google ScholarPubMed
DeSalle, R. (2006). Species discovery versus species identification: response to Rubinoff. Conservation Biology, 20, 1545–7.CrossRefGoogle ScholarPubMed
DeSalle, R., Egan, M. G. & Siddall, M. (2005). The unholy trinity: taxonomy, species delimiting and DNA barcoding. Philosophical Transactions of the Royal Society of London, B, 360, 1905–16, I.CrossRefGoogle Scholar
Desmond, A. & Moore, J. (1991). Darwin. London: Michael Joseph.Google Scholar
Desmond, A. & Moore, J. (2009). Darwin's sacred cause. race, slavery and the quest for human origins. London: Allen Lane, an imprint of Penguin Books.Google Scholar
Dettner, K. & Liepert, C. (1994). Chemical mimicry and camouflage. Annual Review of Entomology, 39, 129–54.CrossRefGoogle Scholar
Devos, N.et al. (2006). On the monophyly of Dactylorhiza Necker ex Nevski (Orchidaceae): is Coeloglossum viride (L.) Hartman a Dactylorhiza?Botanical Journal of the Linnean Society, 152, 261–9.CrossRefGoogle Scholar
DeWoody, J. A.et al. (2010). Molecular approaches in natural resource conservation and management. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Di Cesnola, A. P. (1904). Preliminary note on the protective value of colour in Mantis religiosa. Biometrika, 3, 58–9.CrossRefGoogle Scholar
Diamond, J. (1992). The rise and fall of the third chimpanzee: how our animal heritage affects the way we live. London: Vantage.Google Scholar
Diaz, A. & Macnair, M. R. (1998.) The effect of plant size on the expression of cleistogamy in Mimulus nasutus. Functional Ecology, 12, 92–8.CrossRefGoogle Scholar
Dick, C. W. (2001). Genetic rescue of remnant tropical trees by an alien pollinator. Proceedings of the Royal Society of London, B, 268, 2391–6.CrossRefGoogle ScholarPubMed
Digby, L. (1912). The cytology of Primula kewensis and of other related Primula hybrids. Annals of Botany, 26, 357–88.Google Scholar
Dirzo, R. & Harper, J. L. (1982a). Experimental studies of slug–plant interactions. III. Differences in acceptability of individual plants of Trifolium repens to slugs and snails. Journal of Ecology, 70, 101–17.CrossRefGoogle Scholar
Dirzo, R. & Harper, J. L. (1982b). Experimental studies of slug–plant interactions. IV. The performance of cyanogenic and acyanogenic morphs of Trifolium repens in the field. Journal of Ecology, 70, 119–38.CrossRefGoogle Scholar
Ditta, G.et al. (2004). The SEP4 gene of Arabidopsis thaliana functions in floral organ and meristem identity. Current Biology, 14, 1935–40.CrossRefGoogle ScholarPubMed
Dlugosch, K. M. & Parker, I. M. (2007). Molecular and life history trait variation across the native range of the invasive species Hypericum canariense: evidence for ancient patterns of colonization via pre-adaptation?Molecular Ecology, 16, 4269–83.CrossRefGoogle ScholarPubMed
Dlugosch, K. M. & Parker, I. M. (2008a). Founding events in species invasions: genetic variation, adaptive evolution, and the role of multiple introductions. Molecular Ecology, 17, 431–49.CrossRefGoogle ScholarPubMed
Dlugosch, K. M. & Parker, I. M. (2008b). Invading populations of an ornamental shrub show rapid life history evolution despite genetic bottlenecks. Ecology Letters, 11, 701–9.CrossRefGoogle ScholarPubMed
Dobzhansky, T. (1935). A critique of the species concept in biology. Philosophy of Science, 2, 344–55.Google Scholar
Dobzhansky, T. G. (1937). Genetics and the origin of species. New York: Columbia University Press.Google Scholar
Dobzhansky, T. G. (1941). Genetics and the origin of species, edn. New York: Columbia University Press.Google Scholar
Dobzhansky, T. G. (1951). Genetics and the origin of species, edn. New York: Columbia University Press.Google Scholar
Dodd, M.et al. (1995). Community stability – a 60-year record of trends and outbreaks in the occurrence of species in the Park Grass Experiment. Journal of Ecology, 83, 277–85.CrossRefGoogle Scholar
Dodds, W. K. (2008). Humanity's footprint: momentum, impact, and our global environment. New York: Columbia University Press.Google Scholar
Doebley, J., Stec, A. & Hubbard, L. (1997). The evolution of apical dominance in maize. Nature, 386, 485–8.CrossRefGoogle ScholarPubMed
Dommée, B., Assouad, M. W. & Valdeyron, G. (1978). Natural selection and gynodioecy in Thymus vulgaris L. Botanical Journal of the Linnean Society, 77, 17–28.
Donoghue, M. J. & Catino, P. D. (1988). Paraphyly, ancestors, and the goals of taxonomy: a botanical defense of cladism. The Botanical Review, 54, 107–28.CrossRefGoogle Scholar
Donoghue, M. J. & Kadereit, J. W. (1992). Walter Zimmermann and the growth of phylogenetic theory. Systematic Biology, 41, 74–85.CrossRefGoogle Scholar
Donoghue, M. J. & Sandeson, M. J. (1992). The suitability of molecular and morphological evidence in reconstructing plant phylogeny. In Molecular systematics of plants, ed. Soltis, P. S., Soltis, D. E. & Doyle, J. J., pp.340–68. New York: Chapman & Hall.Google Scholar
Donoghue, M. J. & Smith, S. A. (2004). Patterns in the assembly of temperate forests around the Northern Hemisphere. Philosophical Transactions of the Royal Society, B, 359, 1633–44.CrossRefGoogle ScholarPubMed
Donoghue, M. J., Ree, R. H. & Baum, D. A. (1998). Phylogeny and the evolution of flower symmetry in the Asteridae. Trends in Plant Science, 3, 311–17.CrossRefGoogle Scholar
Donohue, K. (2002). Germination timing influences natural selection on life-history characters in Arabidopsis thaliana. Ecology, 83, 1006–16.CrossRefGoogle Scholar
Doorenbos, J. (1965). Juvenile and adult phases in woody plants. In Encyclopedia of plant physiology, XV/1, ed. Ruhland, W., pp. 1222–35. Berlin: Springer.Google Scholar
Doran, P. T. & Zimmerman, M. K. (2009). Examining the scientific consensus on climate change. Eos, Transactions, American Geophysical Union, 90, 22–3.Google Scholar
Douhovnikoff, V. & Dodd, R. S. (2003). Intra-clonal variation and a similarity threshold for identification of clones: application to Salix exigua using AFLP molecular markers. Theoretical and Applied Genetics, 106, 1307–15.CrossRefGoogle Scholar
Doyle, J. A. (1978). Origin of angiosperms. Annual Review of Ecology & Systematics, 9, 365–92.CrossRefGoogle Scholar
Doyle, J. A. (2012). Molecular and fossil evidence on the origin of angiosperms. Annual Review of Earth and Planetary Sciences, 40, 301–26.CrossRefGoogle Scholar
Doyle, J. A. & Donoghue, M. J. (1986). Phylogeny and the origin of angiosperms: an experimental cladistic approach. Botanical Review, 52, 321–431.CrossRefGoogle Scholar
Doyle, J. J.et al. (2003). Diploid and polyploid reticulate evolution throughout the history of the perennial soybeans (Glycine subgenus Glycine). New Phytologist, 161, 121–32.CrossRefGoogle Scholar
Drayton, B. & Primack, R. B. (2012). Success rate for reintroductions of eight perennial species after 15 years. Restoration Ecology, 20, 299–303.CrossRefGoogle Scholar
Dubois, A. (2007). Naming taxa from cladograms: some confusions, misleading statements, and necessary clarifications. Cladistics, 23, 390–402.CrossRefGoogle Scholar
Dukes, J. S. (2011). Climate change. In Encyclopedia of biological invasions, ed. Simberloff, D. and Rejmanke, M., pp. 113–17. Encyclopedias of the Natural World. No 3. Berkeley, CA: University of California Press.Google Scholar
Dunn, G. & Everitt, B. S. (1982). An introduction to mathematical taxonomy. Cambridge: Cambridge University Press.Google Scholar
Dunstan, W. R. & Henry, T. A. (1901). The nature and origin of the poison of Lotus arabicus. Proceedings of the Royal Society of London, 68, 374–8.CrossRefGoogle Scholar
Dupré, J. (2010). The polygenic organism. In Nature after the genome, ed. Parry, S. & Dupré, J., pp. 19–31. Oxford: Blackwell.Google Scholar
Durka, W.et al. (2005). Molecular evidence for multiple introductions of Garlic Mustard (Alliaria petiolata, Brassicaceae) to North America. Molecular Ecology, 14, 1697–1706.CrossRefGoogle Scholar
Dyer, G. A.et al. (2009). Dispersal of transgenes through maize seed systems in Mexico. PLoS ONE, 4, e5734, DOI:10.1371/journal.pone.0005734.CrossRefGoogle ScholarPubMed
East, E. M. (1913). Inheritance of flower size in crosses between species of Nicotiana. Botanical Gazette, 55, 177–88.CrossRefGoogle Scholar
East, E. M. & Mangelsdorf, A. J. (1925). A new interpretation of the hereditary behaviour of self-sterile plants. Proceedings of the National Academy of Sciences, Washington, 11, 166–83.CrossRefGoogle ScholarPubMed
Ebling, S. K.et al. (2011). Multiple common garden experiments suggest lack of local adaptation in an invasive ornamental plant. Journal of Plant Ecology, 4, 209–20.Google Scholar
Eckert, C. G. (2000). Contributions of autogamy and geitonogamy to self-fertilization in a mass-flowering, clonal plant. Ecology, 81, 532–42.CrossRefGoogle Scholar
Eckert, C. G. & Barrett, S. C. H. (1993). Clonal reproduction and patterns of genotypic diversity in Decodon verticillatum (Lythraceae). American Journal of Botany, 80, 1175–82.CrossRefGoogle Scholar
Eckert, C. G., Manicacci, D. & Barrett, S. C. H. (1996). Genetic drift and founder effect in native versus introduced populations of an invading plant, Lythrum salicaria (Lythraceae). Evolution, 50, 1512–19.CrossRefGoogle Scholar
Eddy, S. R. (2012). The C-value paradox, junk DNA and ENCODE. Current Biology, 22, R896.CrossRefGoogle ScholarPubMed
Edwards, A. W. F. (1986). Are Mendel's results really too close?Biological Reviews, 61, 295–312.CrossRefGoogle ScholarPubMed
Edwards, M. & Richardson, A. J. (2004). Impact of climate change on marine pelagic phenology and trophic mismatch. Nature, 430, 881–4.CrossRefGoogle ScholarPubMed
Ehlers, B. K., & Bataillon, T. (2007). Inconsistent males and the maintenance of labile sex expression in subdioecious plants. New Phytologist, 174, 194–211.CrossRefGoogle Scholar
Ehlers, B. K., Maurice, S. & Bataillon, T. (2005). Sex inheritance in gynodioecious species: a polygenic view. Proceedings of the Royal Society, B, 272, 1795–1802.CrossRefGoogle ScholarPubMed
Ehlers, J., Gibbard, P. L. & Hughes, P. D. (2011). Quaternary glaciations – extent and chronology: a closer look. Amsterdam: Elsevier.Google Scholar
Ehrendorfer, F. (1980). Polyploidy and distribution. In Polyploidy: biological relevance, ed. Lewis, W. H., pp. 45–60. New York: Plenum Press.Google Scholar
Ehrenreich, I. M. & Purugganan, M. D. (2006). The molecular genetic basis of plant adaptation. American Journal of Botany, 93, 953–62.CrossRefGoogle ScholarPubMed
Ehrich, D.et al. (2007). Genetic consequences of Pleistocene range shifts: contrast between the Arctic, the Alps and the East African mountains. Molecular Ecology, 16, 2542–59.CrossRefGoogle ScholarPubMed
Ehrlich, P. R. & Raven, P. H. (1964). Butterflies and plants: a study in coevolution. Evolution, 18, 586–608.CrossRefGoogle Scholar
Ehrlich, P. R. & Raven, P. H. (1969). Differentiation of populations. Science, 165, 1228–32.CrossRefGoogle ScholarPubMed
Ehrlich, P. R, Kareiva, P. M. & Daily, G. C. (2012). Securing natural capital and expanding equity to rescale civilization. Nature, 486, 68–73.CrossRefGoogle ScholarPubMed
Eidesen, P. B. (2007). Nuclear vs. plastid data: complex Pleistocene history of a circumpolar key species. Molecular Ecology, 16, 3902–25.CrossRefGoogle ScholarPubMed
Eigsti, C. J. & Dustin, P. (1955). Colchicine in agriculture, medicine, biology and chemistry. Ames, IA: Iowa State College Press.Google Scholar
Elam, D. R.et al. (2007). Population size and relatedness affect fitness of a self-incompatible invasive plant. Proceedings of the National Academy of Sciences, USA, 104, 549–52.CrossRefGoogle ScholarPubMed
Elkington, T. T. (1968). Introgressive hybridization between Betula nana L. and B. pubescens Ehrh. in North-west Iceland. New Phytologist, 67, 109–18.CrossRefGoogle Scholar
Elliot, E. (1914), see Lamarck.
Ellis, W. M., Keymer, R. J. & Jones, D. A. (1977a). The effect of temperature on the polymorphism of cyanogenesis in Lotus corniculatus L. Heredity, 38, 339–47.Google Scholar
Ellis, W. M., Keymer, R. J. & Jones, D. A. (1977b). On the polymorphism of cyanogenesis in Lotus corniculatus L. VIII. Ecological studies in Anglesey. Heredity, 39, 45–65.CrossRefGoogle Scholar
Ellstrand, N. C. & Elam, D. R. (1993). Population genetic consequences of small population size: implications for plant conservation. Annual Review of Ecology & Systematics, 24, 217–42.CrossRefGoogle Scholar
Ellstrand, N. C. & Marshall, D. L. (1985a). Interpopulation gene flow by pollen in wild radish, Raphanus sativus. The American Naturalist, 126, 606–16.CrossRefGoogle Scholar
Ellstrand, N. C. & Marshall, D. L. (1985b). Variation in extent of multiple paternity among plants and populations of wild radish. American Journal of Botany, 72, 876.Google Scholar
Ellstrand, N. C. & Schierenbeck, K. (2000). Hybridization as a stimulus for the evolution of invasiveness in plants?Proceedings of the National Academy of Sciences, USA, 97, 7043–50.CrossRefGoogle ScholarPubMed
Ellstrand, N. C. & Schierenbeck, K. A. (2006). Hybridization as a stimulus for the evolution of invasiveness in plants?Euphytica, 48, 35–46.Google Scholar
Ellstrand, N. C., Whitkus, R. & Rieseberg, L. H. (1996). Distribution of spontaneous plant hybrids. Proceedings of the National Academy of Sciences, USA, 93, 5090–3.CrossRefGoogle ScholarPubMed
Ellstrand, N. C., Prentice, H. C. & Hancock, J. F. (1999). Gene flow and introgression from domesticated plants into their wild relatives. Annual Review of Ecology & Systematics, 30, 539–63.CrossRefGoogle Scholar
Ellstrand, N. C.et al. (2010). Crops gone wild: evolution of weeds and invasives from domesticated ancestors. Evolutionary Applications, 3, 494–504.CrossRefGoogle ScholarPubMed
Ellstrand, N. C.et al. (2013). Introgression of crop alleles into wild or weedy populations. Annual Review of Ecology, Evolution & Systematics, 44, 325–45.CrossRefGoogle Scholar
Elmqvist, T. (2000). Pollinator extinction in the Pacific Islands. Conservation Biology, 14, 1237–9.Google Scholar
Elwell, A. L.et al. (2011). Separating parental environment from seed size effects on next generation growth and development in Arabidopsis. Plant, Cell and Environment, 34, 291–301.CrossRefGoogle ScholarPubMed
Emerson, B. C. (2002). Evolution on oceanic islands: molecular phylogenetic approaches to understanding pattern and process. Molecular Ecology, 11, 951–66.CrossRefGoogle ScholarPubMed
Emms, S. K. & Arnold, M. L. (1997). The effect of habitat on parental and hybrid fitness: transplant experiments with Louisiana Irises. Evolution, 51, 1112–19.CrossRefGoogle ScholarPubMed
Endersby, J. ed. (2009). Introduction, appendix and explanatory notes. In On the origin of species by Charles Darwin. Cambridge: Cambridge University Press.Google Scholar
Engeldow, F. L. (1950). Rowland Harry Biffin 1874–1949. Obituary Notices of Fellows of the Royal Society, 7, 9–25.Google Scholar
Ennos, R. A. (1982). Association of the cyanogenic loci in White clover. Genetic Research, 40, 65–72.CrossRefGoogle Scholar
Ensslin, A., Sandner, T. M. & Matthies, D. (2011). Consequences of ex situ cultivation of plants: genetic diversity, fitness and adaptation of the monocarpic Cynoglossum officinale L. in botanic gardens. Biological Conservation, 144, 272–8.CrossRefGoogle Scholar
Epperson, B. K. (2007). Plant dispersal, neighbourhood size and isolation by distance. Molecular Ecology, 16, 3854–65.CrossRefGoogle ScholarPubMed
Ereshefsky, M. (2001). The poverty of the Linnaean hierarchy: a philosophical study of biological taxonomy. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Erfmeier, A. & Bruelheide, H. (2011). Maintenance of high genetic diversity during invasion of Rhododendron ponticum in the British Isles. International Journal of Plant Sciences, 172, 795–806.CrossRefGoogle Scholar
Erikkson, G. (1983). Linnaeus the botanist. In Linnaeus. The man and his work, ed. Frangsmyr, T., pp.63–109. Berkeley: University of California Press.Google Scholar
Ernst, A. (1955). Self-fertility in monomorphic primulas. Genetica, 27, 91–148.CrossRefGoogle Scholar
Ernst, W. H. O. (1990). Mine vegetation in Europe. In Heavy metal tolerance in plants: evolutionary aspects, ed. Shaw, A. J., pp. 21–37. Boca Raton, FL: CRC Press.Google Scholar
Ernst, W. H. O. (1998a). Evolution of plants on soils anthropogenically contaminated by heavy metals. In Plant evolution in man-made habitats, ed. Raamsdonk, L. W. D. van & Nijs, J. C. M. den, pp. 13–27. Amsterdam: Hugo de Vries Laboratory.Google Scholar
Ernst, W. H. O. (1998b). Invasion, dispersal and ecology of the South African neophyte Senecio inaequidens in The Netherlands: from wool alien to railway and road alien. Acta Botanica Neerlandica, 47, 131–51.Google Scholar
Ernst, W. H. O. (2006). Evolution of metal tolerance in higher plants. Forest Snow and Landscape Research, 80, 251–74.Google Scholar
Escaravage, N.et al. (1998). Clonal diversity in a Rhododendron ferrugineum L. (Ericaceae) population inferred from AFLP markers. Molecular Ecology, 7, 975–82.CrossRefGoogle Scholar
Evans, R. C. & Turkington, R. (1988). Maintenance of morphological variation in a biotically patchy environment. New Phytologist, 109, 369–76.CrossRefGoogle Scholar
Evenari, M. (1989). The history of research on white-green variegated plants. Botanical Review, 55, 106–39.CrossRefGoogle Scholar
Ewel, J. J. (1986). Invasibility: lessons from southern California. In Ecology of biological invasions of North America and Hawaii, ed. Mooney, H. A. & Drake, J. A., pp. 214–39. New York: Springer Verlag.Google Scholar
Fagerlind, F. (1937). Embryologische, zytologische und bestäubungs-experimentelle Studien in der Familie Rubiaceae nebst Bemerkungen über einige Polyploiditätsprobleme. Acta Horti Bergiani, 11, 195–470.Google Scholar
Fairbanks, D. J. (2008). Mendelian controversies: an update. In Ending the Mendel–Fisher controversy, ed. Franklin, A.et al., pp. 302–11. Pittsburgh, PA: University of Pittsburgh Press.Google Scholar
Falahati-Anbaran, M.et al. (2011). Genetic consequences of seed banks in the perennial herb Arabidopsis lyrata subsp. petraea (Brassicaceae). American Journal of Botany, 98, 1475–85.CrossRefGoogle Scholar
Falconer, D. S. (1952). Introduction to quantitative genetics, edn. London: Longmans.Google Scholar
Falconer, D. S. (1981). Introduction to quantitative genetics. edn. London: Longmans.Google Scholar
Falk, D. A. & Holsinger, K. E. (1991). Genetics and conservation of rare plants. Oxford: Oxford University Press.Google Scholar
Falk, D. A. & Olwell, P. (1992). Scientific and policy considerations in restoration and reintroduction of endangered species. Rhodora, 94, 287–315.Google Scholar
Falk, D. A., Millar, C. I. & Olwell, M. (1966). Restoring diversity: strategies for reintroduction of endangered species. Washington DC: Island Press.Google Scholar
Fant, D. A.et al. (2007). Genetic structure of threatened native populations and propagules used for restoration in a clonal species, American Beachgrass (Ammophila breviligulata Fern.)Restoration Ecology, 16, 594–603.Google Scholar
Fant, J. B.et al. (2013). Genetics of reintroduced populations of the narrowly endemic thistle, Cirsium pitcheri (Asteraceae). Botany-Botanique, 91, 301–8.CrossRefGoogle Scholar
Farjon, A. (2007). In defence of a conifer taxonomy which recognises evolution. Taxon, 56, 639–41.CrossRefGoogle Scholar
Favarger, C. (1967). Cytologie et distribution des plantes. Botanical Review, 42, 163–206.Google Scholar
Favarger, C. (1984). Cytogeography and biosystematics. In Plant biosystematics, ed. Grant, W. F., pp. 453–75. Toronto: Academic Press.Google Scholar
Favarger, C. & Villard, M. (1965). Nouvelles récherches cytotaxinomiques sur Chrysanthemum leucanthemum L. sens. lat. Bericht der Schweizerischen Botanischen Gesellschaft, 75, 57–79.Google Scholar
Fawcett, J. A., Maerea, S. & Van de Peera, Y. (2009). Plants with double genomes might have had a better chance to survive the Cretaceous–Tertiary extinction event. Proceedings of the National Academy of Sciences, USA, 106, 5737–42.CrossRefGoogle ScholarPubMed
Fay, M. & Cowan, R. S. (2001). Plant microsatellites in Cypripedium caleolus (Orchidaceae): genetic fingerprints from herbarium specimens. Lindleyana, 16, 151–6.Google Scholar
Fazekas, A. J.et al. (2010). Stopping the stutter: improvements in sequence quality from regions with mononucleotide repeats can increase the usefulness of non-coding regions for DNA barcoding. Taxon, 59, 694–7.Google Scholar
Fearnside, P. M. & Ferraz, J. (1995). A conservation gap analysis of Brazil's Amazonian vegetation. Conservation Biology, 9, 1134–47.CrossRefGoogle Scholar
Federici, S.et al. (2013). DNA barcoding to analyse taxonomically complex groups in plants: the case of Thymus (Lamiaceae). Botanical Journal of the Linnean Society, 171, 687–99.CrossRefGoogle Scholar
Fenger, J. & Tjell, J. C. (2009). Air pollution – from a local to global perspective. Lyngby: Polyteknisk.Google Scholar
Ferguson, A. (1980). Biochemical systematics and evolution. Glasgow & London: Blackie.Google Scholar
Ferrari, D. M. & Guiseppi, A. R. (2011). Geomorphology and plate tectonics. New York: Nova Publishers.Google Scholar
Ferris, C.et al. (1995). Chloroplast DNA recognizes three refugial sources of European oaks and suggests independent eastern and western immigrations to Finland. Heredity, 80, 584–93.Google Scholar
Ferris, C., King, R. A. & Gray, A. J. (1997). Molecular evidence for maternal parentage in the hybrid origin of Spartina anglica C. E. Hubbard. Molecular Ecology, 6, 185–7.CrossRefGoogle Scholar
Ferris-Kaan, R. (ed.) (1995). The ecology of woodland creation. Chichester & New York: Wiley.Google Scholar
Feuillet, C.et al. (2011). Crop genome sequencing: lessons and rationales. Trends in Plant Science, 16, 77–88.CrossRefGoogle ScholarPubMed
Feulner, M.et al. (2013). Floral scent and its correlation with AFLP data in Sorbus. Organisms Diversity and Evolution, 14, 339–48.Google Scholar
Field, D. L.et al. (2010). Patterns of hybridization and asymmetrical gene flow in hybrid zones of the rare Eucalyptus aggregata and common E. rubida. Heredity, 106, 841–53.Google ScholarPubMed
Fincham, J. R. S. (1983). Genetics. Bristol: Wright.Google Scholar
Fisher, K. (2006). Rank-free monography: a practical example from the moss clade Leucophanella (Calymperaceae). Systematic Botany, 31, 13–30.CrossRefGoogle Scholar
Fisher, R. A. (1929). The genetical theory of natural selection, edn, reprinted 1958. London: Constable; New York: Dover Books.Google Scholar
Fisher, R. A. (1935). The design of experiments. Edinburgh & London: Oliver & Boyd.Google Scholar
Fisher, R. A. (1936). Has Mendel's work been rediscovered?Annals of Science, 1, 115–37.CrossRefGoogle Scholar
Fisher, R. A. & Yates, F. (1963). Statistical tables for biological, agricultural and medical research, edn. Edinburgh: Longman (Oliver & Boyd).Google Scholar
Flake, R. H., von Rudloff, E., & Turner, B. L. (1969). Quantitative study of clinal variation in Juniperus virginiana using terpenoid data. Proceedings of the National Academy of Sciences, USA, 64, 487–94.CrossRefGoogle ScholarPubMed
Flann, C., Turland, N. M. & Monro, A. (2014). Report on botanical nomenclature – Melbourne 2011 XVIII International Botanical Congress, Melbourne: Nomenclature Section, 18–22 July 2011. PhytoKeys, 41, 1–289, DOI:10.3897/phytokeys.41.8398CrossRef
Flowers, J.et al. (2009). Population genomics of the Arabidopsis thaliana flowering time gene network. Molecular Biology and Evolution, 26, 2475–86.CrossRefGoogle ScholarPubMed
Flowers, J. M. & Purugganan, M. D. (2008). The evolution of plant genomes – scaling up from a population perspective. Current Opinion in Genetics & Development, 18, 565–70.CrossRefGoogle ScholarPubMed
Focke, W. O. (1881). Die Pflanzen-Mishlinge ein Beitrag zur Biologie der Gewächse. Berlin: Gebrüder Borntraeger.CrossRefGoogle Scholar
Foden, W.et al. (2007). A changing climate is eroding the geographical range of the Namib Desert tree Aloe through population declines and dispersal lags. Diversity and Distributions, 13, 645–53.CrossRefGoogle Scholar
Ford, V. S. & Gottlieb, L. D. (1992). Bicalyx is a natural homeotic floral variant. Nature, 358, 671–3.CrossRefGoogle Scholar
Forest, F. & Chase, M. W. (2009). Eudicots. In The timetree of life, ed. Hedges, S. B. and Kumar, S., pp.169–176. New York: Oxford University Press.Google Scholar
Forey, P. L., Humphries, C. J. & Kitching, I. J. (1993). Cladistics: a practical approach in systematics. Oxford: Oxford University Press.Google Scholar
Forister, M. L. & Feldman, C. R. (2011). Phylogenetic cascades and the origins of tropical diversity. Biotropica, 43, 270–8.CrossRefGoogle Scholar
Forrest, C. N.et al. (2011). Tests for inbreeding and outbreeding depression and estimation of population differentiation in the bird-pollinated shrub Grevillea mucronulata. Annals of Botany, 108, 185–95.CrossRefGoogle ScholarPubMed
Foster, S. A.et al. (2007). Parallel evolution of dwarf ecotypes in the forest tree Eucalyptus globulus. New Phytologist, 175, 370–80.CrossRefGoogle ScholarPubMed
Fournier-Level, A.et al. (2011). A map of local adaptation in Arabidopsis thaliana. Science, 334, 86–9.CrossRefGoogle ScholarPubMed
Fowler, N. & Levin, D. A. (1984). Ecological constraints on the establishment of a novel polyploid in competition with its diploid progenitor. American Naturalist, 124, 703–11.CrossRefGoogle Scholar
Foxe, J. P.et al. (2010). Reconstructing origins of loss of self-incompatibility and selfing in North American Arabidopsis lyrata: a population genetic context. Evolution, 64, 3495–510.CrossRefGoogle ScholarPubMed
Francico-Ortega, J.et al. (1999). Internal transcribed spacer sequence phylogeny of Crambe L. (Brassicaceae): molecular data reveal two Old World disjunctions. Molecular Phylogenetics and Evolution, 11, 361–80.Google Scholar
Frankel, O. H. & Soulé, M. E. (1981). Conservation and evolution. London: Cambridge University Press.Google Scholar
Frankel, O. H., Brown, A. H. D. & Burdon, J. J. (1995). The conservation of plant biodiversity. Cambridge: Cambridge University Press.Google Scholar
Frankham, R. (1995) Effective population size/adult population size ratios in wildlife: a review. Genetics Research Cambridge, 66, 95–107.CrossRefGoogle Scholar
Frankham, R. (2008). Genetic adaptation to captivity in species conservation programs. Molecular Ecology, 17, 325–33.CrossRefGoogle ScholarPubMed
Frankham, R., Ballou, J. D. & Briscoe, D. A. (2002). Introduction to conservation genetics. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Frankham, R. J. D.et al. (2011). Predicting the probability of outbreeding depression. Conservation Biology, 25, 465–75.CrossRefGoogle ScholarPubMed
Franklin, A. (2008) The Mendel–Fisher controversy. An overview. In Franklin, A.et al., Ending the Mendel–Fisher controversy, pp. 1–77. Pittsburgh, PA: University of Pittsburgh Press.Google Scholar
Franklin, A.et al. (2008). Ending the Mendel–Fisher controversy. Pittsburgh, PA: University of Pittsburgh Press.Google Scholar
Franklin, I. A. (1980). Evolutionary change in small populations. In Conservation biology: an evolutionary-ecological perspective, ed. Soulé, M. E. & Wilcox, B. A., pp. 135–50. Sunderland, MA: Sinauer.Google Scholar
Franklin, I. R. & Frankham, R. (1998). How large must populations be to retain evolutionary potential?Animal Conservation, 1, 69–70.CrossRefGoogle Scholar
Franklin-Tong, V. E. & Franklin, F. C. H. (2003). The different mechanisms of gametophytic self-incompatibility. Philosophical Transactions of the Royal Society of London, B, 358, 1025–32.CrossRefGoogle ScholarPubMed
Franks, S. J. (2010). Genetics, evolution and conservation of island plants. Journal of Plant Biology, 2, 481–8.Google Scholar
Franks, S. J. and Weis, A. E. (2008). A change in climate causes rapid evolution of multiple life-history traits and their interactions in an annual plant. Journal of Evolutionary Biology, 21, 1321–34.CrossRefGoogle Scholar
Franks, S. J., Sim, S. & Weis, A. E. (2007). Rapid response of flowering time by an annual plant in response to a climate fluctuation. Proceedings of the National Academy of Sciences, USA, 104, 1278–82.CrossRefGoogle ScholarPubMed
Franzke, A. & Mummenhoff, K. (1999). Recent hybrid speciation in Cardamine (Brassicaceae). Conversion of nuclear ribosomal ITS sequences in statu nascendi. Theoretical and Applied Genetics, 98, 831–4.CrossRefGoogle Scholar
Freckleton, R. P. & Watkinson, A. R. (2003). Are all plant populations metapopulations?Journal of Ecology, 91, 321–4.CrossRefGoogle Scholar
Frey-Klett, P., Garbaye, J. & Tarkka, M. (2007). The mycorrhiza helper bacteria revisited. New Phytologist, 176, 22–36.CrossRefGoogle ScholarPubMed
Friedman, J. & Barrett, S. C. H. (2009). Wind of change: new insights on the ecology and evolution of pollination and mating in wind-pollinated plants. Annals of Botany, 103, 1515–27.CrossRefGoogle ScholarPubMed
Friedman, S. T. & Adams, W. T. (1985). Estimation of gene flow into two seed orchards of Loblolly-pine (Pinus taeda L.). Theoretical & Applied Genetics, 69, 609–15.Google Scholar
Friis, E. M., Pedersen, K.R. & Crane, P. R. (2010) Diversity in obscurity: fossil flowers and the early history of angiosperms. Philosophical Transactions of the Royal Society, 365, 369–82.CrossRefGoogle ScholarPubMed
Friis, E. M., Crane, P. R. & Pedersen, K. R. (2011). Early flowers and angiosperm evolution. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Frisch, W., Meschede, M. & Blakey, R. C. (2011). Plate tectonics – continental drift and mountain building. Heidelberg: Springer.Google Scholar
Fritsch, P. & Rieseberg, L. H. (1992). High outcrossing rates maintain male and hermaphrodite individuals in populations of the flowering plant Datisca glomerata. Nature, 359, 633–6.CrossRefGoogle Scholar
Frohlich, M. W. & Parker, D. S. (2000). The mostly male theory of flower evolutionary origins: from genes to fossils. Systematic Botany, 25, 155–70.CrossRefGoogle Scholar
Frost, H. B. (1938). Nuclear and juvenile characters in clonal varieties of Citrus. Journal of Heredity, 29, 423–32.CrossRefGoogle Scholar
Fröst, S. & Ising, G. (1968). An investigation into the phenolic compounds in Vaccinium myrtillus L. (Bilberries), Vaccinium vitis-idaea L. (Cowberries), and the hybrid between them V. intermedium Ruthe employing thin layer chromatography. Hereditas, 60, 72–6.Google Scholar
Fryer, J. D. & Chancellor, R. J. (1979). Evidence of changing weed populations in arable land. Proceedings of the 14th British Weed Control Conference, pp. 958–64.
Fuentes, I.et al. (2014). Horizontal genome transfer as an asexual path to the formation of new species. Nature, 511, 232–5.CrossRefGoogle ScholarPubMed
Futuyma, D. J. & Agrawal, A. A. (2009). Macroevolution and the biological diversity of plants and herbivores. Proceedings of the National Academy of Sciences, USA, 106, 18054–61.CrossRefGoogle ScholarPubMed
Gaeta, R. T.et al. (2007). Genomic changes in resynthesized Brassica napus and their effect on gene expression and phenotype. The Plant Cell, 19, 3403–17.CrossRefGoogle ScholarPubMed
Gajewski, W. (1957). A cytogenetic study on the genus Geum. Monographiae Botanicae, No. 4.
Galeuchet, D. J. & Holderegger, R. (2005). (Erhaltung und Wiederansiedlung des Kleinen Rohrkolbens (Typha minima) – Vegetationsaufnahmen, Monitoring und genetische Herkunftsanalysen. Botanica Helvetica, 115, 15–32.CrossRefGoogle Scholar
Gallez, G. P. & Gottlieb, L. D. (1982). Genetic evidence for the hybrid origin of the diploid plant Stephanomeria diegensis. Evolution, 36, 1158–67.CrossRefGoogle ScholarPubMed
Galton, F. (1871). Experiments in pangenesis, by breeding from rabbits of a pure variety, into whose circulation blood taken from other varieties had previously been largely transfused. Proceedings of the Royal Society of London, 19, 393–410.Google Scholar
Galton, F. (1876). The history of twins, as a criterion of the relative powers of nature and nurture. Journal of the Royal Anthropological Institute, 5, 391–406.Google Scholar
Galton, F. (1889). Natural inheritance. London & New York: Macmillan.CrossRefGoogle Scholar
Galton, F. (1904). Eugenics. Its definition, scope and aims. In Essays in eugenics, pp. 35–43. London: Eugenics Education Society.Google Scholar
Galton, F. (1908). Memories of my life. London: Methuen.CrossRefGoogle Scholar
Ganders, F. R. (1979). The biology of heterostyly. New Zealand Journal of Botany, 17, 607–35.CrossRefGoogle Scholar
Ganders, F. R. (1990). Altitudinal clines for cyanogenesis in introduced populations of white clover near Vancouver, Canada. Heredity, 64, 387–90.CrossRefGoogle Scholar
Garbelotto, M. & Pautasso, M. (2012). Impacts of exotic forest pathogens on Mediterranean ecosystems: four case studies. European Journal of Plant Pathology, 133, 101–16.CrossRefGoogle Scholar
Garcia, P.et al. (1989). Allelic and genotypic composition of ancestral Spanish and colonial Californian gene pools of Avena barbata: evolutionary implications. Genetics, 122, 687–94.Google ScholarPubMed
Garcia-Robledo, C.et al. (2013). Tropical plant–herbivore networks: reconstructing species interactions using DNA barcodes. PLOS ONE, 8, 1–10.CrossRefGoogle ScholarPubMed
Garnier, A.et al. (2008). Measuring and modelling anthropogenic secondary seed dispersal along road verges for feral oilseed rape. Basic and Applied Ecology, 9, 533–41.CrossRefGoogle Scholar
Gartside, D. W. & McNeilly, T. (1974). The potential for evolution of metal tolerance in plants. III. Copper tolerance in normal populations of different species. Heredity, 32, 335–48.CrossRefGoogle Scholar
Gates, R. R. (1909). The stature and chromosomes of Oenothera gigas de Vries. Archiv für Zellforschung, 3, 525–52.Google Scholar
Gay, P. A. (1960). A new method for the comparison of populations that contain hybrids. New Phytologist, 59, 219–26.CrossRefGoogle Scholar
Gayon, J. (2009). From Darwin to today in evolutionary biology. In The Cambridge companion to Darwin, ed. Hodge, J. & Radick, G., pp. 227–301. Cambridge: Cambridge University Press.Google Scholar
Ge, Y. Z., Cheng, X. F., Hopkins, A. & Wang, Z. Y. (2007). Generation of transgenic Lolium temulentum plants by Agrobacterium tumefaciens-mediated transformation. Plant Cell Reports, 26, 783–9.CrossRefGoogle ScholarPubMed
Gealy, D. R.et al. (2007). Implications of gene flow in the scale-up and commercial use of biotechnology-derived crop. CAST Issue Paper 37. Ames, IA: Council for Agricultural Science and Technology.Google Scholar
Geiger, R. (1965). The climate near the ground. Cambridge, MA: Harvard University Press.Google Scholar
Genton, B. J, Shykoff, J. A. & Giraud, T. (2005). High genetic diversity in French invasive populations of common ragweed, Ambrosia artemisiifolia, as a result of multiple sources of introduction. Molecular Ecology, 10, 4275–85.Google Scholar
Gerstel, D. U. (1950). Self-incompatibility studies in Guayule. II. Inheritance. Genetics, 35, 482–506.Google Scholar
Ghazoul, J. (2005). Pollen and seed dispersal among dispersed plants. Biological Reviews, 80, 413–43.CrossRefGoogle ScholarPubMed
Giakountis, A.et al. (2010). Distinct patterns of genetic variation alter flowering responses of Arabidopsis accessions to different daylengths. Plant Physiology, 152, 177–91.CrossRefGoogle ScholarPubMed
Gibby, M. (1981). Polyploidy and its evolutionary significance. In The evolving biosphere, ed. Forey, P. L., pp. 87–96. Cambridge: British Museum (Natural History) & Cambridge University Press.Google Scholar
Gilbert, B. & Levine, J. M. (2013). Plant invasions and extinction debts. Proceedings of the National Academy of Sciences, USA, 110, 1744–9.CrossRefGoogle ScholarPubMed
Gilbert, N. (2013) Case studies: a hard look at GM crops. Superweeds? Suicides? Stealthy genes? The true, the false and the still unknown about transgenic crops. Nature, 497, 24–6.Google Scholar
Giles, B. E. & Goudet, J. (1997). Genetic differentiation in Silene dioica metapopulations: estimation of spatiotemporal effects in a successional plant species. American Naturalist, 149, 507–26.CrossRefGoogle Scholar
Giles, B. E., Lundqvist, E. & Goudet, J. (1998). Restricted gene flow and subpopulation differentiation in Silene dioica. Heredity, 80, 715–23.CrossRefGoogle Scholar
Gill, A. M. & Williams, J. E. (1996). Fire regimes and biodiversity: the effects of fragmentation of southeastern Australian eucalypt forests by urbanisation, agriculture and pine plantations. Forest Ecology and Management, 85, 261–78CrossRefGoogle Scholar
Gill, B. S. & Kimber, G. (1974). Giemsa C-banding and the evolution of Wheat. Proceedings of the National Academy of Sciences, USA, 71, 4086–90.CrossRefGoogle ScholarPubMed
Gill, L.et al. (2004). Phylogeography: English elm is a 2,000-year-old Roman clone. Nature, 431, 1053.CrossRefGoogle Scholar
Gillespie, R. G.et al. (2011). Long-distance dispersal: a framework for hypothesis testing. Trends in Ecology and Evolution, 27, 47–56.Google ScholarPubMed
Gillham, N. W. (2001). A life of Sir Francis Galton: from African exploration to the birth of eugenics. Oxford: Oxford University Press.Google Scholar
Gilmour, J. S. L. (1936). Two early papers on classification. [Reprint 1989: Plant Systematics and Evolution, 167, 97–107.]
Gilmour, J. S. L. (1937). A taxonomic problem. Nature, 139, 1040.CrossRefGoogle Scholar
Gilmour, J. S. L. (1940). Taxonomy and philosophy. In The new systematics, ed. Huxley, J., pp. 461–74. Oxford: Clarendon Press.Google Scholar
Gilmour, J. S. L. (1951). The development of taxonomic theory since 1851. Nature, 168, 400–2.CrossRefGoogle ScholarPubMed
Gilmour, J. S. L. & Gregor, J. W. (1939). Demes: a suggested new terminology. Nature, 144, 333–4.CrossRefGoogle Scholar
Gilmour, J. S. L. & Heslop-Harrison, J. (1954). The deme terminology and the units of micro-evolutionary change. Genetica, 27, 147–61.Google ScholarPubMed
Gilmour, J. S. L. & Walters, S. M. (1963). Philosophy and classification. Vistas in Botany, 4, 1–22.Google Scholar
Gilpin, M. E. & Soulé, M. E. (1986). Minimum viable populations: processes of species extinction. In Conservation biology: the science of scarcity and diversity, pp. 19–34. Sunderland, MA: Sinauer.Google Scholar
Given, D. R. (1994). Principles and practice of plant conservation. London: Chapman & Hall.Google Scholar
Givnish, T. (1979). On the adaptive significance of leaf form. In Topics in plant population biology, ed. Solbrig, O. T., Jain, S., Johnson, G. B. & Raven, P. H., pp. 375–407. London: Columbia University Press.Google Scholar
Givnish, T. J. (2010). Ecology of plant speciation. Taxon, 59, 1326–66.Google Scholar
Givnish, T. J. & Renner, S. S. (2004). Tropical intercontinental disjunctions: Gondwana breakup, immigration from the Boreotropics, and transoceanic dispersal. International Journal of Plant Sciences, 165(4 Suppl.), S1–S6.CrossRefGoogle Scholar
Givnish, T. J. & Vermeij, G. J. (1976). Sizes and shapes of Liane leaves. The American Naturalist, 110, 743–76.CrossRefGoogle Scholar
Glass, B. (1959). Heredity and variation in the eighteenth century concept of the species. In Forerunners of Darwin 1745–1859, ed. Glass, B., Temkin, O. & Straus, W. I., pp. 144–72. London: Oxford University Press.Google Scholar
Glen, W. (ed.) (1994). The mass extinction debates. Stanford, CA: Stanford University Press.Google Scholar
Glenn, T. C. I. & Schable, N. A. (2005). Isolating microsatellite DNA loci. Methods in Enzymology, 395, 202–22.Google ScholarPubMed
Gliddon, C. & Saleem, M. (1985). Gene flow in Trifolium repens – an expanding genetic neighbourhood. In Genetic differentiation and dispersal in plants, ed. Jacquard, P., Heim, G. & Antonovics, J., pp. 293–309. Berlin: Springer–Verlag.Google Scholar
Godefroid, S.et al. (2011a). How successful are plant species reintroductions?Biological Conservation, 144, 672–82.CrossRefGoogle Scholar
Godefroid, S.et al. (2011b). To what extent are threatened European plant species conserved in seed banks?Biological Conservation, 144, 1494–8.CrossRefGoogle Scholar
Godfray, H. C. J. (2002). Challenges for taxonomy. Nature, 417, 17–19.CrossRefGoogle ScholarPubMed
Godfray, H. C. J. & Knapp, S. (2004). Introduction. In ‘Taxonomy for the twenty-first century’. Philosophical Transactions of the Royal Society of London, B, 359, 559–69.Google Scholar
Godfray, H. C. J.et al. (2007). The web and the structure of taxonomy. Systematic Biology, 56. 943–55.CrossRefGoogle ScholarPubMed
Godt, M. J.Caplow, W. F. & Hamrick, J. L. (2005). Allozyme diversity in the federally threatened golden paintbrush, Castilleja levisecta (Scrophulariaceae). Conservation Genetics, 6, 87–99.CrossRefGoogle Scholar
Godward, M. B. E. (1985). The kinetochore. International Review of Cytology, 94, 77–105.Google ScholarPubMed
Goebel, K. (1897). Uber Jugendformen von Pflanzen und deren künstliche Wiederhervorrufung. Sitzungsbericht der Mathematisch-Physikalischen Class der Königlich-Bayerischen Akademie der Wissenschaften, München, 26.
Goerke, H. (1989). Carl von Linné. Stuttgart: Wissenschaftliche Verlagsgesellschaft.Google Scholar
Göhre, V. & Paszkowski, U. (2006). Contribution of arbuscular mycorrhizal symbiosis to heavy metal phytoremediation. Planta, 223, 1115–22.CrossRefGoogle ScholarPubMed
Goldblatt, P. (1980). Polyploidy in angiosperms: monocotyledons. In Polyploidy, ed. Lewis, W. H., pp. 219–39. New York & London: Plenum Press.Google Scholar
Goldblatt, P. (2007). Index to plant chromosome numbers. Taxon, 56, 984–986.CrossRefGoogle Scholar
Goldblatt, P. & Johnson, D. E. (1990) Index to plant chromosome numbers 1986–1987. St Louis, MO: Missouri Botanical Garden.Google Scholar
Goldblatt, P., Hoch, P. C. & McCook, L. M. (1992). Documenting scientific data: the need for voucher specimens. Annals of the Missouri Botanical Garden, 79, 969–70.CrossRefGoogle Scholar
Goldschmidt, R. B. (1940). The material basis of evolution. New Haven, CT: Yale University Press.Google Scholar
Goldstein, P. Z. & DeSalle, R. (2011). Integrating DNA barcode data and taxonomic practice: determination, discovery, and description. BioEssays, 33, 135–47.CrossRefGoogle ScholarPubMed
Gordo, O. & Sanz, J. J. (2005). Phenology and climate change: a long-term study in a Mediterranean locality. Oecologia, 146, 484–95.CrossRefGoogle Scholar
Gore, A. (2013). Earth in the balance. London: Random House.Google Scholar
Gornall, R. J. (1999). Population genetic structure in agamospermous plants. In Molecular systematics and plant evolution, ed. Hollingsworth, P. M., Bateman, R. M. & Gornall, R. J., pp. 118–38. London: Taylor & Francis.Google Scholar
Gornitz, V., Rosenzweig, D. & Hillel, D. (1997). Effects of anthropogenic intervention in the land hydrologic cycle on global sea level rise. Global and Planetary Change, 14, 147–61.CrossRefGoogle Scholar
Gottlieb, L. C. & Pilz, G. (1976). Genetic similarity between Gaura longifolia and its obligately outcrossing derivative G. demareei. Systematic Botany, 1, 181–7.CrossRefGoogle Scholar
Gottlieb, L. D. (1972). Levels of confidence in the analysis of hybridization in plants. Annals of Missouri Botanic Garden, 59, 435–46.CrossRefGoogle Scholar
Gottlieb, L. D. (1973). Genetic differentiation, sympatric speciation, and the origin of a diploid species of Stephanomeria. American Journal of Botany, 60, 545–53.CrossRefGoogle Scholar
Gottlieb, L. D. (1979). The origin of phenotype in a recently evolved species. In Topics in plant population biology, ed. Solbrig, O. T., Jain, S., Johnson, G. B. & Raven, P. H., pp. 264–86. New York: Columbia University Press.Google Scholar
Gottlieb, L. D. (1981a). Electrophoretic evidence and plant populations. Progress in Phytochemistry, 7, 1–46.Google Scholar
Gottlieb, L. D. (1981b). Gene number in species of Asteraceae that have different chromosome numbers. Proceedings of the National Academy of Sciences USA, 78, 3726–9.CrossRefGoogle Scholar
Gottlieb, L. D. (1982). Conservation and duplication of isozymes in plants. Science, 216, 373–80.CrossRefGoogle ScholarPubMed
Gottlieb, L. D. (1984). Isozyme evidence and problem solving in plant systematics. In Plant biosystematics, ed. Grant, W. F., pp. 343–57. Toronto: Academic Press.Google Scholar
Gottlieb, L. D. (1986). Genetic differentiation, speciation and phylogeny in Clarkia (Onagraceae). In Modern aspects of species, ed. Iwatsuki, K., Raven, P. H. & Bock, W. J., pp. 145–60. Tokyo: University of Tokyo Press.Google Scholar
Gottlieb, L. D. (2003). Rethinking classic examples of recent speciation in plants. New Phytologist, DOI: 10.1046/j.1469-8137.2003.00922.xCrossRef
Gottlieb, L. D., Warwick, S. I. & Ford, V. S. (1985). Morphological and electrophoretic divergence between Layia discoidea and L. glandulosa. Systematic Botany, 10, 484–95.CrossRefGoogle Scholar
Goudie, A. (2009). The human impact: man's role in environmental change. Oxford: Blackwell.Google Scholar
Goudriaan, J.et al. (1999). Use of models in global change studies. In The terrestrial biosphere and global change, ed. Walker, B., Steffen, W., Canadell, J. & Ingram, J., pp. 106–40. Cambridge: Cambridge University Press.Google Scholar
Gould, S. J. (1979). Species are not specious. New Scientist, 83, 374–6.Google Scholar
Gould, S. J. (2007). Punctuated equilibrium. Cambridge, MA: Belknap Press.Google Scholar
Gould, S. J. & Eldredge, N. (1977). Punctuated equilibria: the tempo and mode of evolution reconsidered. Paleobiology, 3, 115–51.CrossRefGoogle Scholar
Gould, S. J. & Eldredge, N. (1993). Punctuated equilibrium comes of age. Nature, 366, 223–7.CrossRefGoogle ScholarPubMed
Gould, S. J. & Lewontin, R. G. (1979). The spandrels of San Marco and the panglossian paradigm: a critique of the adaptionist programme. Proceedings of the Royal Society of London, B, 205, 581–8.CrossRefGoogle Scholar
Govindaraju, D. R. (1988). Relationship between dispersal ability and levels of gene flow in plants. Oikos, 52, 31–5.CrossRefGoogle Scholar
Gradstein, F. M. & Ogg, J. G. (2009). The geologic time scale. In Tree of life, ed. Hedges, S. B. & Kumar, S., pp. 26–34. Oxford: Oxford University Press.Google Scholar
Graetz, R. D., Olsson, L. & Wilson, M. A. (1995). Interpreting a 9 year time-series of satellite observations for the Australian continent. Proceedings of the Satellite Colloquium of the 10th International Congress of Photosynthesis, Montpellier: International Society for Photogrammetry and Remote Sensing, 395–406.Google Scholar
Graham, B. F Jr & Bormann, F. H. (1966). Natural root grafts. The Botanical Review, 32, 255–92.CrossRefGoogle Scholar
Grandont, L., Jenczewski, E. & Lloyd, A. (2013). Meiosis and its deviations in polyploidy plants. Cytogenetic and Genome Research, 140, 171–84.CrossRefGoogle Scholar
Grant, V. (1950). The protection of the ovules in flowering plants. Evolution, 4, 179–201.CrossRefGoogle Scholar
Grant, V. (1966). The selective origin of incompatibility barriers in the plant genus Gilia. The American Naturalist, 100, 99–118.CrossRefGoogle Scholar
Grant, V. (1971). Plant speciation. New York & London: Columbia University Press.Google Scholar
Grant, V. (1975). Genetics of flowering plants. New York & London: Columbia University Press.Google Scholar
Grant, V. (1981). Plant speciation. edn. New York: Columbia University Press.Google Scholar
Grant, V. & Grant, K. A. (1965). Flower pollination in the Phlox family. New York: Columbia University Press.Google Scholar
Grant-Downton, R. T. & Dickinson, H. G. (2006). Epigenetics and its implications for plant biology. 2. The ‘epigenetic epiphany’: epigenetics, evolution and beyond. Annals of Botany, 97, 11–27.CrossRefGoogle ScholarPubMed
Graur, D. & Li, W.-H. (1999). Fundamentals of molecular evolution, edn. Sunderland, MA: Sinauer Associates.Google Scholar
Graur, D. & Li, W.-H. (2000). Fundamentals of molecular evolution, edn. Sunderland, MA: Sinauer Associates.Google Scholar
Graur, D. & Martin, W. (2004). Reading the entrails of chickens: molecular timescales of evolution and the illusion of precision. Trends in Genetics, 20, 80–6.CrossRefGoogle ScholarPubMed
Graur, D.et al. (2013). On the immortality of television sets: ‘function’ in the human genome according to the evolution-free gospel of ENCODE. Genome Biology & Evolution, 5, 578–90.CrossRefGoogle ScholarPubMed
Gray, A. (2004). Will Spartina anglica invade northwards with changing climate? In Third International Conference on Invasive Spartina. San Francisco, CA, 8–10 November 2004. San Francisco: Invasive Spartina Project.Google Scholar
Gray, A. J., Marshall, D. F. & Raybould, A. F. (1991). A century of evolution in Spartina anglica. Advances in Ecological Research, 21, 1–62.Google Scholar
Green, M. R. & Sambrook, J. (2012). Molecular cloning: a laboratory manual. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.Google Scholar
Green, R. H. (1979). Sampling design and statistical methods for environmental biologists. New York, Chichester, Brisbane & Toronto: Wiley.Google Scholar
Greene, E. L. (1909). Linnaeus as an evolutionist. Proceedings of the Washington Academy of Sciences, 11, 17–26.Google Scholar
Greenwood, M. S. (1986). Gene exchange in loblolly pine: the relation between pollination mechanism, female receptivity and pollen availability. American Journal of Botany, 73, 1443–51.CrossRefGoogle Scholar
Gregor, J. W. (1930). Experiments on the genetics of wild populations, I. Plantago maritima. Journal of Genetics, 22, 15–25.CrossRefGoogle Scholar
Gregor, J. W. (1931). Experimental delimitation of species. New Phytologist, 30, 204–17.Google Scholar
Gregor, J. W. (1938). Experimental taxonomy. 2. Initial population differentiation in Plantago maritima in Britain. New Phytologist, 37, 15–49.CrossRefGoogle Scholar
Gregor, J. W. (1939). Experimental taxonomy. 4. Population differentiation in North American and European Sea Plantains allied to Plantago maritima L. New Phytologist, 38, 293–322.
Gregor, J. W. (1944). The ecotype. Biological Reviews, 19, 20–30.CrossRefGoogle Scholar
Gregor, J. W. (1946). Ecotypic differentiation. New Phytologist, 45, 254–70.CrossRefGoogle Scholar
Gregor, J. W. & Lang, J. M. S. (1950). Intra-colonial variation in plant size and habit in Sea Plantains. New Phytologist, 49, 135–41.CrossRefGoogle Scholar
Gregor, J. W., Davey, V. McM. & Lang, J. M. S. (1936). Experimental taxonomy. 1. Experimental garden technique in relation to the recognition of small taxonomic units. New Phytologist, 35, 323–50.CrossRefGoogle Scholar
Greig-Smith, P. (1964). Quantitative plant ecology, edn. London: Butterworth.Google Scholar
Greiner, S.et al. (2011). The role of plastids in plant speciation. Molecular Ecology, 20, 671–91.CrossRefGoogle ScholarPubMed
Gressel, J. (2005a). Introduction: the challenges of ferality. In Crop ferality and volunteerism, ed. Gressel, J., pp. 1–7. Boca Raton, FL: Taylor & Francis.CrossRefGoogle Scholar
Gressel, J. (ed.) (2005b).Crop ferality and volunteerism. Boca Raton, FL: Taylor & Francis.CrossRefGoogle Scholar
Griffiths, M. (1994). Index of garden plants. London: Macmillan.Google Scholar
Groot, J. & Boschuizen, R. (1970). A preliminary investigation into the genecology of Plantago major L. Journal of Experimental Botany, 21, 835–41.
Gross, B. L. & Rieseberg, L. H. (2005). The ecological genetics of homoploid hybrid speciation. Journal of Heredity, 96, 241–52.CrossRefGoogle ScholarPubMed
Grotewold, E., Chappell, J. & Kellogg, E. (2015). Plant genes, genomes and genetics. Chichester: Wiley-Blackwell.CrossRefGoogle Scholar
Grove, A. T. & Rackham, O. (2001). The nature of Mediterranean Europe: an ecological history. New Haven, CT, & London: Yale University Press.Google Scholar
Groves, R. H. & di Castri, F. (1991). Biogeography of Mediterranean invasions. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Gruber, S. & Claupein, W. (2007). Fecundity of volunteer oilseed rape and estimation of potential gene dispersal by a practice-related model. Agriculture, Ecosystems & Environment, 119, 401–8.CrossRefGoogle Scholar
Gübeli, A., Hochuli, P. A. & Wildi, W. (1984). Lower Cretaceous turbiditic sediment from the Rif chain (northern Morocco), palynology, stratigraphy setting. Geologisches Rundschau, 73, 1081–114.CrossRefGoogle Scholar
Guerrant, E. O. (1992). Genetic and demographic considerations in the sampling and reintroduction of rare plants. In Conservation biology, ed. Fiedler, P. L. & Jain, S. K., pp. 321–44. New York: Chapman & Hall.Google Scholar
Guerrant, E. O. (2013). The value and propriety of reintroduction as a conservation tool for rare plants. Botany-Botanique, 91, v–x.CrossRefGoogle Scholar
Gugerli, F., Parducci, L. & Petit, R. J. (2005). Ancient plant DNA: review and prospects. New Phytologist, 166, 409–18.CrossRefGoogle ScholarPubMed
Guggisberg, A.et al. (2008). Genomic origin and organization of the allopolyploid Primula egaliksensis investigated by in situ hybridization. Annals of Botany, 101, 919–27.CrossRefGoogle ScholarPubMed
Guggisberg, A.et al. (2013). Transcriptome divergence between introduced and native populations of Canada thistle, Cirsium arvense. New Phytologist, 199, 595–608.CrossRefGoogle ScholarPubMed
Guignard, L. (1891). Nouvelles études sur la fécondation. Annales des Sciences naturelles (Botanique), 14, 163–296.Google Scholar
Gunther, R. W. T. (1928). Further correspondence of John Ray. London & New York: Oxford University Press.Google Scholar
Guo, L.-L.et al. (2011). Evolution of the S-locus region in Arabidopsis relatives. Plant Physiology, 157, 937–46.CrossRefGoogle ScholarPubMed
Gupta, P. K. & Varshney, R. K. (1999) Molecular markers for genetic fidelity during micropropagation and conservation. Current Science, 76, 1308–10.Google Scholar
Gustafsson, Å. (1946). Apomixis in higher plants. 1. The mechanism of apomixis. Acta Universitatis Lundensis, 42 (3), 1–67.Google Scholar
Gustafsson, Å. (1947a). Apomixis in higher plants. 2. The causal aspect of apomixis. Acta Universitatis Lundensis, 43(3), 69–179.Google Scholar
Gustafsson, Å. (1947b). Apomixis in higher plants. 3. Biotype and species formation. Acta Universitatis Lundensis, 43(12), 183–370.Google Scholar
Gustafsson, L. & Gustafsson, P. (1994). Low genetic variation in Swedish populations of the rare species Vicia pisiformis (Fabaceae) revealed with RFLP & RAPD. Plant Systematics & Evolution, 189, 133–48.CrossRefGoogle Scholar
Gwynn-Jones, D., Lee, J. A. & Callaghan, T. V. (1997) Effects of enhanced UV-B and elevated carbon dioxide concentrations on a sub-arctic forest ecosystem. Plant Ecology, 128, 243–49.CrossRefGoogle Scholar
Habel, J. C. & Zachos, F. E. (2012). Habitat fragmentation versus fragmented habitats. Biodiversity and Conservation, 21, 2987–90.CrossRefGoogle Scholar
Haeckel, E. H. P. A. (1876). The history of creation. (Translation revised by Lankester, E. R..) London: Routledge.Google Scholar
Haffer, J. (1969). Speciation in Amazonian forest birds. Science, 165, 131–7.CrossRefGoogle ScholarPubMed
Haffer, J. (2007). Ornithology, evolution, and philosophy: the life and science of Ernst Mayr 1904–2005. San Francisco: Ignatius Press.Google Scholar
Haffer, J. & Prance, G. T. (2001). Climatic forcing of evolution in Amazonia during the Cenozoic: on the refuge theory of biotic differentiationAmazoniana, 16, 579–607.Google Scholar
Hagerup, O. (1932). Uber polyploidie in Beziehung zu Klima, Okologie, und Phyogenie – Chromosomenzahlen aus Timbuktu. Hereditas, 16, 19–40.Google Scholar
Håkansson, S. (1983). Competition and production in short-lived crop-weed stands: density effects. Uppsala: Department of Plant Husbandry. Report No. 127. Swedish University of Agricultural Sciences.Google Scholar
Haldane, J. B. S. (1932). The causes of evolution. London: Longmans.Google Scholar
Hall, B. G. (2011). Phylogeny for physiologists: a pragmatic guide to building trees: phylogenetic trees made easy. A how-to manual, edn. Sunderland, MA: Sinauer.Google Scholar
Hall, M. (2000). Comparing damages: Italian and American concepts of restoration. In Methods and approaches in forest history, ed. Agnoletti, M. & Anderson, S., pp. 165–72. Wallingford, UK: CAB International.Google Scholar
Hall, O. (1972). Oxygen requirements of root meristems in diploid and autotetraploid Rye. Hereditas, 70, 69–74.Google Scholar
Hallam, A. (1973). A revolution in the earth sciences: from continental drift to plate tectonics. Oxford: Clarendon Press.Google Scholar
Hampe, A. (2011). Plants on the move: the role of seed dispersal and initial population establishment for climate-driven range expansions. Acta Oecologica – International Journal of Ecology, 37, 666–73.CrossRefGoogle Scholar
Hampe, A. & Petit, R. J. (2007). Ever deeper phylogeographies: trees retain the genetic imprint of Tertiary plate tectonics. Molecular Ecology, 16, 5113–14.CrossRefGoogle ScholarPubMed
Hamrick, J. L. (1990). Isozymes and the analysis of genetic structure in plant populations. In Isozymes in plant biology, ed. Soltis, D. E. & Soltis, P. S., pp. 87–105. London: Chapman & Hall.Google Scholar
Hamrick, J. L. & Godt, M. J. W. (1989). Allozyme diversity in plant species. In Plant population genetics, breeding and genetic resources, ed. Brown, A. D. H., Clegg, M. T., Kahler, A. L. & Weir, B. S., pp. 43–63. Sunderland, MA: Sinauer.Google Scholar
Hamrick, J. L. & Loveless, M. D. (1989). Associations between the breeding system and the genetic structure of tropical tree populations. In The evolutionary ecology of plants, ed. Bock, J. & Linhart, Y. B., pp. 129–46. Boulder, CO: Westview Press.Google Scholar
Hamrick, J. L., Linhart, Y. B. & Mitton, J. B. (1979). Relationships between life history characteristics and electrophoretically detectable genetic variation in plants. Annual Review of Ecology & Systematics, 10, 173–200.CrossRefGoogle Scholar
Han, F.et al. (2005). Rapid and repeatable elimination of a parental genome-specific DNA repeat (pGc1R–1a) in newly synthesized wheat allopolyploids. Genetics, 170, 1239–45.CrossRefGoogle ScholarPubMed
Hancock, A. M.et al. (2011). Adaptation to climate across the Arabidopsis thaliana genome. Science, 334, 83–6.CrossRefGoogle ScholarPubMed
Hancock, C. N.et al. (2003). The S-locus and unilateral incompatibility. Transactions of the Royal Society, London, 358, 1133–40.Google ScholarPubMed
Hancock, J. F. (2012). Plant evolution and the origin of crop species, edn. Wallingford: CABI.CrossRefGoogle Scholar
Hannah, L. (2011). Climate change, connectivity, and conservation success. Conservation Biology, 25, 1139–42.CrossRefGoogle ScholarPubMed
Hansen, M. M.et al. (2012). Monitoring adaptive genetic responses to environmental change. Molecular Ecology, 21, 1311–29.CrossRefGoogle ScholarPubMed
Hao, Y.-Q.et al. (2012). The role of late-acting self-incompatibility and early-acting inbreeding depression in governing female fertility in monkshood, Aconitum kusnezoffii. PLOS ONE, DOI:10.1371/journal.pone.0047034eyCrossRef
Harberd, D. J. (1957). The within population variance in genecological trials. New Phytologist, 56, 269–80.CrossRefGoogle Scholar
Harberd, D. J. (1958). Progress and prospects in genecology. Record of the Scottish Breeding Station, 1958, 52–60.Google Scholar
Harberd, D. J. (1961). Observations on population structure and longevity of Festuca rubra. L. New Phytologist, 60, 184–206.
Harberd, D. J. (1962). Some observations on natural clones in Festuca ovina. New Phytologist, 61, 85–100.CrossRefGoogle Scholar
Harberd, D. J. (1963). Observations on natural clones of Trifolium repens L. New Phytologist, 62, 198–204.
Harborne, J. B. (1993). Introduction to ecological biochemistry, edn. San Diego, CA: Academic Press.Google Scholar
Harborne, J. B. & Turner, B. L. (1984). Plant chemosystematics. Orlando, FL: Academic Press.Google Scholar
Harborne, J. B., Williams, C. A. & Smith, D. M. (1973). Species-specific Kaempferol derivatives in ferns of the Appalachian Asplenium complex. Biochemical Systematics, 1 (5), 1–4.CrossRefGoogle Scholar
Hardesty, B. D.et al. (2006) Genetic evidence of frequent long-distance recruitment in a vertebrate-dispersed tree. Ecology Letters, 9, 516–25.CrossRefGoogle Scholar
Hardin, G. (1966). Biology: its principles and implications, edn. London & San Francisco: Freeman.Google Scholar
Hardin, J.Bertoni, G. & Kleinsmith, L. J. (2009). Becker's world of the cell, edn. Boston: Benjamin Cummings.Google Scholar
Hardwick, K. A.et al. (2011). The role of botanic gardens in science and practice of ecological restoration. Conservation Biology, 25, 265–75.Google ScholarPubMed
Hargreaves, A. L., Harder, L. D. & Johnson, S. D. (2009). Consumptive emasculation: the ecological and evolutionary consequences of pollen theft. Biological Reviews of the Cambridge Philosophical Society, 84, 259–76.CrossRefGoogle ScholarPubMed
Harlan, H. V. & Martini, M. L. (1938). The effect of natural selection on a mixture of barley varieties. Journal of Agricultural Research, 57, 189–99.Google Scholar
Harlan, J. R. & de Wet, J. M. J. (1975). On Ö Winge and a prayer: The origins of polyploidy. The Botanical Review, 41, 361–90.CrossRefGoogle Scholar
Harley, J. L. & Harley, E. L. (1987). A check-list of mycorrhiza in the British Flora. New Phytologist (Supplement), 105, 1–102.Google Scholar
Harnesk, H. (2007). Linnaeus: genius of Uppsala. Uppsala: Hallgren & Fallgren.Google Scholar
Harper, J. L. (1977). Population biology of plants. London & New York: Academic Press.Google Scholar
Harper, J. L. (1978). The demography of plants with clonal growth. In Structure and functioning of plant populations, ed. Freyson, A. H. J. & Waldendorp, J. W., pp. 27–48. Amsterdam, Oxford & New York: North Holland Publishing Company.Google Scholar
Harper, J. L. (1983). A Darwinian plant ecology. In Evolution from molecules to men, ed. Bendall, D. S., pp. 323–45. Cambridge: Cambridge University Press.Google Scholar
Harris, J.et al. (2009). Soil microbial communities and restoration ecology: facilitators or followers?Science 325, 573–4.CrossRefGoogle ScholarPubMed
Harris, R. (2005). Attacks on taxonomy. American Scientist, 93, 311.Google Scholar
Harris, S. A. & Ingram, R. (1991). Chloroplast DNA and biosystematics: the effects of intraspecific diversity and plastid transmission. Taxon, 40, 393–401.CrossRefGoogle Scholar
Harris, S. A. & Ingram, R. (1992). Molecular systematics of the genus Senecio L. I. Hybridization in a British polyploid complex. Heredity, 69, 1–10.CrossRefGoogle Scholar
Harrison, C. J. & Langdale, J. A. (2006). A step by step guide to phylogeny reconstruction. The Plant Journal, 45, 561–72.Google ScholarPubMed
Harshberger, J. W. (1901). The limits of variation in plants. Proceedings of the National Academy of Sciences, USA, 53, 303–19.Google Scholar
Hartl, D. L. & Clark, A. G. (2004). Principles of population genetics, edn. Sutherland, MA: Sinauer.Google Scholar
Harvey, W. H. (1860). Darwin on the origin of species. The Gardeners’ Chronicle and Agricultural Gazette, 18 February, 145–6.CrossRef
Hathaway, W. H. (1962). Weighted hybrid index. Evolution, 16, 1–10.Google Scholar
Hauser, T. P. & Loeschcke, V. (1994). Inbreeding depression and mating-distance dependent offspring fitness in large and small populations of Lychnis flos-cuculi. Journal of Evolutionary Biology, 7, 609–22.CrossRefGoogle Scholar
Hayes, W. (1964). The genetics of bacteria and their viruses. Oxford & Edinburgh: Blackwell.Google Scholar
Hayward, I. M. & Druce, G. C. (1919). Adventive flora of Tweedside. Arbroath, UK: Buncle.Google Scholar
Hazarika, M. H. & Rees, H. (1967). Genotypic control of chromosome behaviour in Rye. X. Chromosome pairing and fertility in autotetraploids. Heredity, 22, 317–32.CrossRefGoogle Scholar
He, S.et al. (2009). Dynamics of the evolution of the genus Agrostis revealed by GISH/FISH. Crop Science, 49, 2285–90.CrossRefGoogle Scholar
He, X.-J., Chen, T. & Zhu, J.-K. (2011). Regulation and function of DNA methylation in plants and animals. Cell Research, 21, 442–65.CrossRefGoogle ScholarPubMed
Heap, I. (2008). International survey of herbicide resistant weeds. www.weedscience.org/in.asp.
Hebert, P. D. N. (2003). Biological identification through DNA barcodes. Philosophical Transactions of the Royal Society of London, B, 270, 313–21.Google ScholarPubMed
Heckmann, S. & Houben, A. (2012). Holokinetic centromeres. In Plant centromere biology, ed. Jiang, J. & Birchler, J., pp. 83–94. Ames, IA: Wiley-Blackwell.Google Scholar
Heckmann, S.et al. (2013). The holocentric species Luzula elegans shows interplay between centromere and large-scale genome organization. Plant Journal, 73, 555–65.CrossRefGoogle ScholarPubMed
Hedges, S. B. & S. Kumar, S. (2009). The time tree of life. New York: Oxford University Press.Google Scholar
Hegarty, M. J. and Hiscock, S. J. (2005). Hybrid speciation in plants: new insights from molecular studies. New Phytologist, 165, 411–23.Google ScholarPubMed
Hegarty, M. J.et al. (2009). Extreme changes to gene expression associated with homoploid hybrid speciation. Molecular Ecology, DOI:10.1111/j.1365-294X.2008.04054.xCrossRef
Hegarty, M. J.et al. (2011). Nonadditive changes in cytosine methylation as a consequence of hybridization and genome duplication in Senecio (Asteraceae). Molecular Ecology, 20, 105–13.CrossRefGoogle Scholar
Hegarty, M. J., Abbott, R. J. & Hiscock, S. J. (2012). Allopolyploid speciation in action: the origins and evolution of Senecio cambrensis. In Polyploidy and genome evolution, ed. Soltis, P. S. & Soltis, D. E., pp. 245–70. Heidelberg: Springer.Google Scholar
Hegarty, M.et al. (2013). Lessons from natural and artificial polyploids in higher plants. Cytogenetic and Genome Research, 140, 204–25.CrossRefGoogle ScholarPubMed
Hegland, S. J.et al. (2009). How does climate warming affect plant–pollinator interactions?Ecological Letters, 12, 184–95.CrossRefGoogle ScholarPubMed
Heiser, C. B. Jr (1949a). Natural hybridization with particular reference to introgression. The Botanical Review, 15, 645–87.CrossRefGoogle Scholar
Heiser, C. B. Jr (1949b). Studies in the evolution of sunflower species Helianthus annuus and H. bolanderi. University of California Publication in Botany, 23, 157–96.Google Scholar
Heiser, C. B. Jr (1973). Introgression re-examined. The Botanical Review, 39, 347–66.CrossRefGoogle Scholar
Heiser, C. B. Jr (1979). Hybrid populations of Helianthus divaricatus and H. microcephalus after 22 years. Taxon, 28, 71–5.CrossRefGoogle Scholar
Heiser, C. B. Jret al. (1969). The North American Sunflower (Helianthus). Memoirs of the Torrey Botanical Club, 22, 1–218.Google Scholar
Helmann, J. J.et al. (2008). Five potential consequences of climate change for invasive species. Conservation Biology, 22, 534–43.CrossRefGoogle Scholar
Hemming, M. N. & Trevaskis, B. (2011). Make hay while the sun shines: the role of MADS-box genes in temperature dependant seasonal flowering responses. Plant Science, 180, 447–53.CrossRefGoogle ScholarPubMed
Hendry, A. P. (2009). Evolutionary biology: speciation. Nature, 458, 162–4.CrossRefGoogle ScholarPubMed
Hennig, W. (1950). Grundzüge einer Theorie der phylogenetischen Systematik. Berlin: Deutscher Zentralverlag.Google Scholar
Hermanutz, L. A., Innes, D. J. & Weis, I. M. (1989). Clonal structure of Arctic Dwarf Birch (Betula glandulosa) at its northern limit. American Journal of Botany, 76, 755–61.CrossRefGoogle Scholar
Herrera, J. C., D'Hont, A. & Lashermes, P. (2007). Use of fluorescence in situ hybridization as a tool for introgression analysis and chromosome identification in coffee (Coffea arabica L.)Genome, 50, 619–26.CrossRefGoogle Scholar
Heslop-Harrison, J. (1953). New concepts in flowering-plant taxonomy. London: Heinemann.Google Scholar
Heslop-Harrison, J. (1964). Forty years of genecology. Advances in Ecological Research, 2, 159–247.Google Scholar
Heslop-Harrison, J. (1978). Cellular recognition systems in plants. Studies in Biology, No. 100. London: Arnold.Google Scholar
Heslop-Harrison, J. S. & Schmidt, T. (2007). Plant nuclear genome composition. In Handbook of plant science, ed. Roberts, K., pp. 565–72. Chichester: Wiley.Google Scholar
Heslop-Harrison, J. S. & Schwarzacher, T. (2011). Organisation of the plant genome in chromosomes. The Plant Journal, 66, 18–33.CrossRefGoogle Scholar
Hesselman, H. (1919). Iakttagelser øver skogsträdpollens spridningsførmagå. Meddelanden från Statens Skogsførsoksanstalt, 16, 27.Google Scholar
Hewitt, G. (1996). Some genetic consequences of ice ages, and their role in divergence and speciation. Biological Journal of the Linnean Society, 58, 247–76.CrossRefGoogle Scholar
Hewitt, G. (2000). The genetic legacy of the Quaternary ice ages. Nature, 405, 907–13.CrossRefGoogle ScholarPubMed
Heywood, V. H. (1976). Plant taxonomy, edn. Studies in Biology, No. 5. London: Arnold.Google Scholar
Heywood, V. H. (1980). The impact of Linnaeus on botanical taxonomy – past, present and future. Veröffentlichungen der Joachim Jungius-Gesellschaft der Wissenschaften Hamburg, 43, 97–115.Google Scholar
Heywood, V. H. (1983). Botanic gardens and taxonomy: their economic role. Bulletin of the Botanical Survey of India, 25, 134–47.Google Scholar
Heywood, V. H. (2001). Floristics and monography – an uncertain future?Taxon, 50, 361–80.CrossRefGoogle Scholar
Heywood, V. H. (2011). The role of botanic gardens as resource and introduction centres in the face of global change. Biodiversity & Conservation, 20, 221–39.CrossRefGoogle Scholar
Heywood, V. H. & Stuart, S. N. (1992). Species extinctions in tropical forests. In Tropical deforestation, ed. Whitmore, T. C. & Sayer, J., pp. 91–117. London: Chapman & Hall.Google Scholar
Heywood, V. H. & Wyse Jackson, P. S. (1991). Tropical botanic gardens: their role in conservation and development. London: Academic Press.Google Scholar
Hierro, J. L., Maron, J. L. & Callaway, R. M. (2005). A biogeographical approach to plant invasions: the importance of studying exotics in their introduced and native range. Journal of Ecology, 93, 5–15.CrossRefGoogle Scholar
Hiesey, W. M. & Milner, H. W. (1965). Physiology of ecological races and species. Annual Review of Plant Physiology, 16, 203–16.CrossRefGoogle Scholar
Hillis, D. M., Moritz, C. & Mable, B. K. (1996). Molecular systematics, edn, Sunderland, MA: Sinauer.Google Scholar
Hinton, W. F. (1976). The evolution of insect-mediated self-pollination from an outcrossing system in Calyptridium (Portulacaceae). American Journal of Botany, 63, 979–86.CrossRefGoogle Scholar
Hipp, A. L., Rothrock, P. E. & Roalson, E. H. (2009). The evolution of chromosome arrangements in Carex (Cyperaceae). The Botanical Review, 75, 96–109.CrossRefGoogle Scholar
Ho, P., Zhao, J. H. & Dayuan, X. (2009). Rethinking agro-biotechnological innovations in emerging economies: the case of Bt Cotton in China. Journal of Peasant Studies, 36, 345–64.Google Scholar
Hoballah, M. E.et al. (2007). Single gene-mediated shift in pollinator attraction in Petunia. The Plant Cell, 19, 779–90.CrossRefGoogle ScholarPubMed
Hobbs, R. J. & Cramer, V. A. (2008). Restoration ecology: interventionist approaches for restoring and maintaining ecosystem function in the face of rapid environmental change. Annual Review of Environment and Resources, 33, 39–61.CrossRefGoogle Scholar
Hochuli, P. & Feist-Burkhardt, S. (2013). Angiosperm-like pollen and Afropollis from the Middle Triassic (Anisian) of the Germanic Basin (Northern Switzerland). Frontiers in Plant Science, 4: 344.CrossRefGoogle Scholar
Hodge, J. (2009). The notebook programmes and projects of Darwin's London years. In The Cambridge companion to Darwin, edn, ed. Hodge, J. and Radick, G., pp. 44–72. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Hodge, J. (2013). The origins of the Origin: Darwin's first thoughts about the tree of life and natural selection, 1837–1839. In The Cambridge encyclopedia of Darwin and evolutionary thought, ed. Ruse, M., pp. 64–71. Cambridge: Cambridge University Press.Google Scholar
Hodgins, K. A. & Barrett, S. C. H. (2008). Natural selection on floral traits through male and female function in wild populations of the heterostylous daffodilNarcissus triandrus. Evolution, 62, 1751–63.Google ScholarPubMed
Hodkinson, T. R. & Parnell, J. A. N. (2006). Reconstructing the Tree of Life: taxonomy and systematics of species rich taxa. Boca Raton, FL: CRC Press.CrossRefGoogle Scholar
Hoebee, S. E.et al. (2007). Mating patterns and contemporary gene flow by pollen in a large continuous and a small isolated population of the scattered forest tree Sorbus torminalis. Heredity, 99, 47–55.CrossRefGoogle Scholar
Hoegh-Guldberg, O.et al. (2008). Assisted colonization and rapid climate change. Science, 321, 345–6.CrossRefGoogle ScholarPubMed
Hoffmann, H. (1881). Rückblick auf meine Variations-Versuche von 1855–80. Botanische Zeitung, 1881, 345–51, 361–7, 377–83, 393–9, 409–15, 424–31.
Holderegger, R., Kamm, U. & Gugerli, F. (2006). Adaptive vs. neutral genetic diversity: implications for landscape genetics. Landscape Ecology, 21, 797–807.CrossRefGoogle Scholar
Holliday, R. J. & Putwain, P. D. (1977). Evolution of resistance to simazine in Senecio vulgaris L. Weed Research, 17, 291–6.
Holliday, R. J. & Putwain, P. D. (1980). Evolution of herbicide resistance in Senecio vulgaris: variation in susceptibility to simazine between and within populations. Journal of Applied Ecology, 17, 779–91.CrossRefGoogle Scholar
Hollingsworth, M. L. & Bailey, J. P. (2000). Evidence for massive clonal growth in the invasive weed Fallopia japonica (Japanese Knotweed). Botanical Journal of the Linnean Society, 133, 463–72.CrossRefGoogle Scholar
Hollingsworth, P. M., Graham, S. W. & Little, D. P. (2011). Choosing and using a plant barcode. PLoS ONE, 6, 1–13.CrossRefGoogle Scholar
Holm, L. G.et al. (1977). The world's worst weeds: distribution and biology. Honolulu, HI: University of Honolulu Press.Google Scholar
Holmes, D. S. & Bougourd, S. M. (1991). B chromosome selection in Allium schoenoprasum. II. Experimental populations. Heredity, 67, 117–22.CrossRefGoogle Scholar
Holsinger, K. E. (1984). The nature of biological species. Philosophy of Science, 51, 293–307.CrossRefGoogle Scholar
Holsinger, K. E. & Gottlieb, L. D. (1991). Conservation of rare and endangered plants. In Genetics and conservation of rare plants, ed. Falk, D. A. & Holsinger, K. E., pp. 195–208. New York: Oxford University Press.Google Scholar
Holttum, R. E. (1967). Comparative morphology, taxonomy and evolution. Phytomorphology, 17, 36–41.Google Scholar
Honjo, M.et al. (2008). Tracing the origins of stocks of the endangered species Primula sieboldii using nuclear microsatellites and chloroplast DNA. Conservation Genetics, 9, 1139–47.CrossRefGoogle Scholar
Honnay, O.et al. (2005). Forest fragmentation effects on patch occupancy and population viability of herbaceous plant species. New Phytologist, 166, 723–36.CrossRefGoogle ScholarPubMed
Honnay, O.et al. (2008). Can a seed bank maintain the genetic variation in the aboveground plant populations?Oikos, 117, 1–5.CrossRefGoogle Scholar
Hooglander, N., Lumaret, R. & Bos, M. (1993). Inter-intraspecific variation of chloroplast DNA of European Plantago species. Heredity, 70, 322–34.CrossRefGoogle Scholar
Hooykaas, P. J. J. (2007). Plant transformation. In Handbook of plant science, ed. Roberts, K., pp.665–70. Chichester: Wiley.Google Scholar
Hoquet, T. (2013). The evolution of the Origin (1859–1872). In The Cambridge encyclopedia of Darwin and evolutionary thought, ed. Ruse, M., pp 158–64. Cambridge: Cambridge University Press.Google Scholar
Hörandl, E. (2010). Beyond cladistics: extending evolutionary classifications into deeper time levels. Taxon, 59, 45–350.Google Scholar
Hörandl, E. & Paun, O. (2007). Patterns and sources of genetic diversity in apomictic plants: implications for evolutionary potentials. In Apomixis: evolution, mechanisms and perspectives, ed. Hörandl, E., Grossniklaus, U., Dijk, P. J. Van & Sharbel, T., pp. 169–94. Ruggell, Liechtenstein: Gantner.Google Scholar
Hörandl, E. & Stuessy, T. F. (2010). Paraphyletic groups as natural units of biological classification. Taxon, 59, 1641–53.Google Scholar
Hörandl, E.et al. (2007). Apomixis: evolution, mechanisms and perspectives. Ruggell, Liechtenstein: Gantner.Google Scholar
Horsman, D. C., Roberts, T. M. & Bradshaw, A. D. (1979). Studies on the effect of sulphur dioxide on perennial ryegrass (Lolium perenne, L.). II. Evolution of sulphur dioxide tolerance. Journal of Experimental Botany, 30, 495–501.CrossRefGoogle Scholar
Hort, A. (1938). The Critica botanica of Linnaeus. London: Ray Society, British Museum.Google Scholar
Houben, A.et al. (2001). The genomic complexity of micro B chromosomes of Brachycome dichromosomatica. Chromosoma, 110, 451–9.CrossRefGoogle ScholarPubMed
Houben, A., Nasuda, S. & Endo, T. R. (2011). Plant B chromosomes, Methods in MolecularBiology, 701, 97–111.Google Scholar
Houliston, G. J. & Chapman, H. M. (2004). Reproductive strategy and population variability in the facultative apomict Hieracium pilosella (Asteraceae). American Journal of Botany, 91, 37–44.CrossRefGoogle Scholar
House of Lords (2010). Report into systematics and taxonomy by the House of Lords Science and Technology Committee. London: HMSO.Google Scholar
Howe, C. J. (2007). Chloroplast genome. In Handbook of plant science, ed. Roberts, K., pp. 585–587. Chichester: Wiley.Google Scholar
Hoyle, B. (2008). Plight of the pines. Nature Reports Climate Change, 24 April 2008, DOI:10.1038/climate.2008.35.CrossRef
Hsiao, J.-Y. & Li, H.-L. (1973). Chromatographic studies on the Red Horsechestnut (Aesculus × carnea) and its putative parent species. Brittonia, 25, 57–63.CrossRefGoogle Scholar
Hu, S., Dilcher, D. L. & Taylor, D. W. (2012). Pollen evidence for the pollination biology of early flowering plants. In Evolution of the plant–pollinator relationships, ed. Patiny, S., pp.165–218. Cambridge: Cambridge University Press.Google Scholar
Hu, S.-J.et al. (2011). Hybridization and asymmetric introgression between Cypripedium tibeticum and C. yunnanense in Shangrila County, Yunnan Province, China. Nordic Journal of Botany, 29, 625–31.CrossRefGoogle Scholar
Hu, Y., Poh, H. M. & Chua, N. H. (2006). The Arabidopsis ARGOS-like gene regulates cell expansion during organ growth. Plant Journal, 47, 1–9.CrossRefGoogle ScholarPubMed
Huang, J.et al. (2010). Functional analysis of the Arabidopsis PAL gene family in plant growth, development, and response to environmental stress. Plant Physiology, 153, 1526–38.CrossRefGoogle ScholarPubMed
Huck, S.et al. (2009). Range-wide phylogeography of the European temperate-montane herbaceous plant Meum athamanticum Jacq.: evidence for periglacial persistence. Journal of Biogeography, 36, 1588–99.CrossRefGoogle Scholar
Hudson, R. R., Keitman, M. & Aguadé, M.et al. (1987). A test of neutral molecular evolution based on nucleotide data. Genetics, 116, 153–9.Google ScholarPubMed
Hughes, A. (1959). A history of cytology. London & New York: Abelard-Schuman.Google Scholar
Hughes, C. & Eastwood, R. (2006). Island radiation on a continental scale: exceptional rates of plant diversification after uplift of the Andes. Proceedings of the National Academy of Sciences, USA, 103, 10334–9.CrossRefGoogle ScholarPubMed
Hughes, C. E., Eastwood, R. J. & Bailey, C. D. (2006). From famine to feast? Selecting nuclear DNA sequence loci for plant species-level phylogeny reconstruction. Philosophical Transactions of the Royal Society, B, 361, 211–25.CrossRefGoogle Scholar
Hughes, M. A. (1991). The cyanogenic polymorphism in Trifolium repens L. (white clover). Heredity, 66, 105–15.CrossRefGoogle Scholar
Hughes, M. B. & Babcock, E. B. (1950). Self-incompatibility in Crepis foetida L. subsp. rhoeadifolia. Genetics, 35, 570–88.Google Scholar
Hughes, N. F. (1994). The enigma of angiosperm origins. Cambridge: Cambridge University Press.Google Scholar
Humborg, C.et al. (1997). Effect of Danube river dam on Black Sea biogeochemistry and ecosystem structure. Nature, 386, 385.CrossRefGoogle Scholar
Humeau, L., Pailler, T. & Thompson, J. D. (1999). Cryptic dioecy and leaky dioecy in endemic species of Dombeya (Sterculiaceae) on La Réunion. American Journal of Botany, 86, 1437–47.CrossRefGoogle ScholarPubMed
Hummel, S. (2003). Ancient DNA: typing methods, strategies and applications. Heidelberg: Springer.CrossRefGoogle Scholar
Humphries, C. J. & Funk, V. A. (1984). Cladistic methodology. In Current concepts in plant taxonomy, ed. Heywood, V. H. & Moore, D. M., pp. 323–62. London: Systematics Association Special Volume 25, Academic Press.Google Scholar
Husband, B. C. (2004). The role of triploid hybrids in the evolutionary dynamics of mixed-ploidy populations. Biological Journal of the Linnean Society, 82, 537–46.CrossRefGoogle Scholar
Husband, B. C.et al. (2008). Mating consequences of polyploid evolution in flowering plants: current trends and insights from synthetic polyploids. International Journal of Plant Science, 169, 195–206.CrossRefGoogle Scholar
Huskins, C. L. (1930). The origin of Spartina townsendii. Genetica, 12, 531–8.CrossRefGoogle Scholar
Hussey, M. A.et al. (1991). Incluence of photoperiod on the frequency of sexual embryo sacs in facultative apomictic Buffelgrass. Euphytica, 54, 141–5.Google Scholar
Hutchings, M. J. (1989). Population biology and conservation of Ophrys sphegodes. In Methods of orchid conservation, ed. Pritchard, H. W., pp. 101–15. Cambridge: Cambridge University Press.Google Scholar
Hutchinson, A. H. (1936). The polygonal representation of polyphase phenomena. Transactions of the Royal Society of Canada, Ser. 3, Sect. V, 30, 19–26.Google Scholar
Hutchinson, J. (1926). The families of flowering plants, vol. 1, Dicotyledons. London: Macmillan.Google Scholar
Hutchinson, J. (1959) The families of flowering plants, edn. Oxford: Oxford University Press.Google Scholar
Hutchinson, T. C. (1967). Ecotype differentiation in Teucrium scorodonia with respect to susceptibility to lime-induced chlorosis and to shade factors. New Phytologist, 66, 439–53.CrossRefGoogle Scholar
Huxley, A., Griffiths, M. & Levy, M. (1992). The new Royal Horticultural Society dictionary of gardening. London: Macmillan Press.Google Scholar
Huxley, C. R. (1991). Ants and plants: a diversity of interactions. In Ant–plant interactions, ed. Huxley, C. R. & Cutler, D. F., pp. 1–11. Oxford: Oxford University Press.Google Scholar
Huxley, C. R. & Cutler, D. F. (eds.) (1991). Ant–plant interactions. Oxford: Oxford University Press.Google Scholar
Huxley, J. S. (1938). Clines: an auxiliary taxonomic principle. Nature, 142, 219–20.CrossRefGoogle Scholar
Huxley, J. S. (ed.) (1940). The new systematics. Oxford: Clarendon Press.Google Scholar
Huxley, J. S. (1942). Evolution: the modern synthesis. London: Allen & Unwin.Google Scholar
Igic, B., Lande, R. & Kohn, J. R. (2008). Loss of self-incompatibility and its evolutionary consequences. International Journal of Plant Sciences, 169, 93–104.CrossRefGoogle Scholar
Iltis, H. (1932). Life of Mendel. London: Allen & Unwin. Reprinted 1966 New York: Hafner.Google Scholar
Imerson, A. (2012). Desertification, land degradation and sustainability. London: Wiley.Google Scholar
Ingrouille, M. J. & Stace, C. A. (1985). Pattern of variation of agamospermous Limonium (Plumbaginaceae) in the British Isles. Nordic Journal of Botany, 5, 113–25.Google Scholar
Ingvarsson, P. K. & Giles, B. E. (1999). Kin-structured colonization and small-scale genetic differentiation in Silene dioica. Evolution, 53, 605–11.CrossRefGoogle ScholarPubMed
IPCC, Intergovernmental Panel on Climate Change (2013). Summary for policy makers. 5th Assessment Report: In Climate change 2013: the physical science basis (AR5). (IPCC website www.ipcc.ch and the IPCC WGI AR5 website www.climatechange2013.org.)
IPCC, Intergovernmental Panel on Climate Change (2014). Summary for policy makers. In Climate change 2014: impacts, adaptation, and vulnerability. Part A: global and sectoral aspects. 5th Assessment Report of the IPCC, ed. Field, C. B.et al., pp.1–32. Cambridge & New York: Cambridge University Press.
Irwin, R. E., Adler, L. S. & Brody, A. K. (2004). The dual role of floral traits: pollinator attraction and plant defense. Ecology, 85, 1503–11.CrossRefGoogle Scholar
Jablonka, E. (2013). Epigenetic plasticity: the responsive germline. Progress in Biophysics and Molecular Biology, 111, 99–107.CrossRefGoogle ScholarPubMed
Jablonka, E. & Lamb, M. J. (1999). Epigenetic inheritance and evolution: the Lamarckian dimension. Oxford: Oxford University Press.Google Scholar
Jablonka, E. & Lamb, M. J (2010). Transgenerational epigenetic inheritance. In Evolution – the extended synthesis, ed. Pigliucci, M. & Müller, G. B., pp. 137–74. Cambridge, MA, & London: The MIT Press.Google Scholar
Jackson, R. C. (1962). Interspecific hybridization in Haplopappus and its bearing on chromosome evolution in the Blepharodon section. American Journal of Botany, 49, 119–32.CrossRefGoogle Scholar
Jackson, R. C. (1965). A cytogenetic study of a three-paired race of Haplopappus gracilis. American Journal of Botany, 52, 946–53.CrossRefGoogle Scholar
Jackson, S. T. & Hobbs, R. J. (2009). Ecological restoration in the light of ecological history. Science, 325, 567–8.CrossRefGoogle ScholarPubMed
Jacobson, M. Z. (2002). Atmospheric pollution: history, science, and regulation. New York: Cambridge University Press.CrossRefGoogle Scholar
Jaén-Molinia, R.et al. (2009). The molecular phylogeny of Matthiola R. Br. (Brassicaceae) inferred from ITS sequences, with special emphasis on the Macaronesian endemics. Molecular Phylogenetics and Evolution, 53, 972–81.Google Scholar
Jaenicke-Després, V.et al. (2003). Early allelic selection in maize as revealed by ancient DNA. Science, 302, 1206–8.CrossRefGoogle ScholarPubMed
Jain, S. K. & Martins, P. S. (1979). Ecological genetics of the colonizing ability of Rose Clover (Trifolium hirtum All.). American Journal of Botany, 66, 361–6.CrossRefGoogle Scholar
Jäkäläniemi, A., Tuomi, J., Siikamäki, P. & Kilpiä, A. (2005). Colonization-extinction and patch dynamic of the perennial riparian plant, Silene tatarica. Journal of Ecology, 93, 670–80.Google Scholar
Jakubowski, A. R., Casier, M. D. & Jackson, R. D. (2011). Has selection for improved agronomic traits made reed canarygrass invasive?PLoS ONE, 6, e25757.CrossRefGoogle ScholarPubMed
James, J. K. & Abbott, R. J. (2005). Recent, allopatric, homoploid hybrid speciation: the origin of Senecio squalidus (Asteraceae) in the British Isles from a hybrid zone on Mount Etna, Sicily. Evolution, 59, 2533–47.CrossRefGoogle ScholarPubMed
Jameson, D. L. (ed.) (1977). Evolutionary genetics. Stroudsburg, PA.: Dowden, Hutchinson & Ross.Google Scholar
Jang, T.-S.et al. (2013). Chromosomal diversification and karyotype evolution of diploids in the cytologically diverse genus Prospero (Hyacinthaceae). BMC Evolutionary Biology, 13, Article ARTN 136.CrossRefGoogle Scholar
Janovec, J. P., Clark, L. G. & Mori, S. A (2003). Is the Neotropical Flora ready for the phylocode?The Botanical Review, 69, 22–43.CrossRefGoogle Scholar
Jansen, P. A., Bongers, F. & Hemerik, L. (2004). Seed mass and mast seeding enhance dispersal by a neotropical scatter-hoarding rodent. Ecological Monographs, 74, 569–89.CrossRefGoogle Scholar
Jansen, R. K., Holsinger, K. E., Michaels, H. J. & Palmer, J. D. (1990). Phylogenetic analysis of restriction site data at higher taxonomic levels: an example from the Asteraceae. Evolution, 44, 2089–105.CrossRefGoogle ScholarPubMed
Jansen, R. K., Michaels, H. J. & Palmer, J. D. (1991). Phylogeny and character evolution in the Asteraceae based on chloroplast DNA restriction site mapping. Systematic Botany, 16, 98–115.CrossRefGoogle Scholar
Janzen, D. H. (1966). Coevolution of mutualism between ants and acacias in Central America. Evolution, 3, 249–75.Google Scholar
Janzen, D. J. (1977). What are dandelions and aphids? AmericanNaturalist, 111, 586–9.CrossRefGoogle Scholar
Janzen, D. H. (2001). Latent extinction: the living dead. In Encyclopedia of biodiversity, vol. 3, ed. Levin, S. A., pp. 689–99. London: Academic Press.Google Scholar
Jarvis, D. I. & Hodgkin, T. (1999). Wild relatives and crop cultivars: detecting natural introgression and farmer selection of new genetic combinations in agroecosystems. Molecular Ecology, 8, S159–S173.CrossRefGoogle Scholar
Jenczewski, E. & Alix, K. (2004). From diploids to allopolyploids: the emergence of efficient pairing control genes in plants. Critical Reviews in the Plant Sciences, 23, 21–45.CrossRefGoogle Scholar
Jenkin, F. (1867). Unsigned review of Darwin's ‘On the origin of species’. The North British Review, July, 277–318.
Jennersten, O. (1988). Pollination in Dianthus deltoides (Caryophyllaceae): effects of habitat fragmentation on visitation and seed set. Conservation Biology, 2, 359–66.CrossRefGoogle Scholar
Jennings, D. E. & Rohr, D. E. (2011). A review of conservation threats to carnivorous plants. Biological Conservation, 144, 1356–63.CrossRefGoogle Scholar
Jensen, I. & Bogh, H. (1941). On conditions influencing the danger of crossing in the case of wind-pollinated cultivated plants. Tidsskrift for Planteavl, 46, 238–66.Google Scholar
Jentschke, G. & Godbold, D. (2000). Metal toxicity and ectomycorrhizas. Physiologia Plantarum, 109, 107–16.CrossRefGoogle Scholar
Jersáková, J.et al. (2006). Mechanisms and evolution of deceptive pollination in orchids. Biological Reviews of the Cambridge Philosophical Society, 81, 219–35.CrossRefGoogle ScholarPubMed
Johannsen, W. (1909). Elemente der exakten Erblichkeitslehre. Jena: Fischer.Google Scholar
Johannsen, W. (1911). The genotype concept in heredity. American Naturalist, 45, 129–59.CrossRefGoogle Scholar
Johanson, U.et al. (2000). Molecular analysis of FRIGIDA, a major determinant of natural variation in Arabidopsis flowering time. Science, 290, 344–7.CrossRefGoogle Scholar
Johnson, B. L. (1972). Protein electrophoretic profiles and the origin of the B genome of wheat. Proceedings of the National Academy of Sciences, USA, 69, 1398–402.CrossRefGoogle Scholar
Johnson, H. (1945). Interspecific hybridization within the genus Betula. Hereditas, 31, 163–76.Google Scholar
Johnson, H. B. (1975). Plant pubescence: an ecological perspective. Botanical Review, 41, 233–58.CrossRefGoogle Scholar
Johnson, M. A. T., Kenton, A. Y., Bennett, M. D. & Bradham, P. E. (1989). Voanioala gerardii has the highest known chromosome number in the monocotyledons. Genome, 32, 328–33.CrossRefGoogle Scholar
Johnson, N. A. (2010). Hybrid incompatibility genes: remnants of a genomic battlefield? Trends inGenetics, 26, 317–25.Google Scholar
Johnson, P. E. (1993). Darwin on trial. Washington DC: Regnery Gateway.Google Scholar
Johnson, S. D. (1992). Plant–animal relationships. In The ecology of fynbos: nutrients, fire and diversity, ed. Cowling, R., pp. 195–205. Cape Town: Oxford University Press.Google Scholar
Jones, A. G.et al. (2010) A practical guide to methods of parentage analysis. Molecular Ecology Resources, 10, 6–30.CrossRefGoogle ScholarPubMed
Jones, C. J.et al. (1997). Reproducibility testing of RAPD, AFLP and SSR markers in plants by a network of European laboratories. Molecular Breeding, 3, 381–90.CrossRefGoogle Scholar
Jones, C. J.et al. (1998). Reproducibility testing of RAPDs by a network of European laboratories. In Molecular tools for screening biodiversity, ed. Karp, A., Isaac, P. G. & Ingram, D. S., pp. 176–80. London: Chapman & Hall.Google Scholar
Jones, D. A. (1962). Selective eating of the acyanogenic form of the plant Lotus corniculatus L. by various animals. Nature, 193, 1109–10.CrossRefGoogle Scholar
Jones, D. A. (1966). On the polymorphism of cyanogenesis in Lotus corniculatus. Selection by animals. Canadian Journal of Genetics and Cytology, 8, 556–67.CrossRefGoogle Scholar
Jones, D. A. (1972). Cyanogenic glycosides and their function. In Phytochemical ecology, ed. Harborne, J. B., pp. 103–24. London & New York: Academic Press.Google Scholar
Jones, D. A. (1973). Co-evolution and cyanogenesis. In Taxonomy and ecology, ed. Heywood, V. H., pp. 213–42. Systematics Association Special Volume No. 5. London & New York: Academic Press.Google Scholar
Jones, D. A., Keymer, R. J. & Ellis, W. M. (1978). Cyanogenesis in plants and animal feeding. In Biochemical aspects of plant and animal coevolution, ed. Harborne, J. B., pp. 21–34. London, New York & San Francisco: Academic Press.Google Scholar
Jones, D. F. (1924). The attainment of homozygosity in inbred strains of Maize. Genetics, 9, 405–18.Google ScholarPubMed
Jones, K. (1958). Cytotaxonomic studies in Holcus. I. The chromosome complex in Holcus mollis L. New Phytologist, 57, 191–210.
Jones, K. (1964). Chromosomes and the nature and origin of Anthoxanthum odoratum L. Chromosoma, 15, 248–74.
Jones, K. & Borrill, M. (1961). Chromosomal status, gene exchange and evolution in Dactylis. 3. The role of the inter-ploid hybrids. Genetica, 32, 296–322.Google Scholar
Jones, M. D. & Brooks, J. S. (1952). Effect of tree barriers on outcrossing in Corn. Oklahoma Agricultural Experiment Station Bulletin, No. T-45.
Jones, M. E. (1971a). The population genetics of Arabidopsis thaliana. I. The breeding system. Heredity, 27, 39–50.Google Scholar
Jones, M. E. (1971b). The population genetics of Arabidopsis thaliana. II. Population structure, Heredity, 27, 51–8.Google Scholar
Jones, M. E. (1971c). The population genetics of Arabidopsis thaliana. III. The effect of vernalisation. Heredity, 27, 59–72.Google Scholar
Jones, R.et al. (2012). The molecular life of plants. Chichester: Wiley-Blackwell.Google Scholar
Jones, R. N. (1995). B chromosomes in plants. New Phytologist, 131, 411–34.CrossRefGoogle Scholar
Jones, R. N., Viegas, W. & Houben, A. (2008). A century of B chromosomes in plants: so what?Annals of Botany, 101, 767–75.CrossRefGoogle Scholar
Jones, S. (1999). Almost like a whale. London: Doubleday.Google Scholar
Jones, S. (2009). Darwin's island: the Galápagos in the garden of England. London: Little, Brown.Google Scholar
Jonsell, B. (1978). Linnaeus's views on plant classification and evolution. Botaniska Notiser, 131, 523–30.Google Scholar
Jonsell, B. (1984). The biological species concept reexamined. In Plant biosystematics, ed. Grant, W. F., pp. 159–68. Toronto: Academic Press.Google Scholar
Jordan, A. (1864). Diagnoses d'espèces nouvelles ou méconnues pour servir de matériaux à une flore réformée de la France et des Contrées voisines. Paris: Savy.Google Scholar
Jordanova, L. J. (1984). Lamarck. Oxford: Oxford University Press.Google Scholar
Jorgensen, T. H. & Frydenberg, J. (1999). Diversification in insular plants: inferring the phylogenetic relationship in Aeonium (Crassulaceae) using ITS sequences of nuclear ribosomal DNA. Nordic Journal of Botany, 19, 613–21.CrossRefGoogle Scholar
Judd, W. S.et al. (2007). Plant systematics: a phylogenetic approach, edn. Sunderland, MA: Sinauer.Google Scholar
Kadereit, J. W. (1994). Molecules and morphology, phylogenetics and genetics. Botanica Acta, 107, 369–73.CrossRefGoogle Scholar
Kadereit, J. W. (2015). The geography of hybrid speciation in plants. Taxon, DOI:10.12705/644.1.CrossRefGoogle Scholar
Kadereit, J. W. & Baldwin, B. G. (2012). Western Eurasian–western North American disjunct plant taxa: the dry-adapted ends of formerly widespread north temperate mesic lineages – and examples of long-distance dispersal. Taxon, 61, 3–17.Google Scholar
Kadereit, J. W. & Briggs, D. (1985). Speed of development of radiate and non-radiate plants of Senecio vulgaris L. from habitats subject to different degrees of weeding pressure. New Phytologist, 99, 155–69.CrossRefGoogle Scholar
Kane, N. C. & Rieseberg, L. H. (2008). Genetics and evolution of weedy Helianthus annuus populations: adaptation of an agricultural weed. Molecular Ecology, 17, 384–94.CrossRefGoogle ScholarPubMed
Kaplan, J. M (2010). Phenotypic plasticity and reaction norms. In A companion to the philosophy of biology, ed. Sarker, S. & Plutyski, A., pp. 205–22. Malden, MA: Blackwell.Google Scholar
Karpechenko, G. D. (1927). Polyploid hybrids of Raphanus sativus L. × Brassica oleracea L. Bulletin of Applied Botany & Plant Breeding (Leningrad), 17, 305–410.
Karpechenko, G. D. (1928). Polyploid hybrids of Raphanus sativus L. × Brassica oleracea L. Zeitschrift für induktive Abstrammungs Vererbungsiehre, 39, 1–7.
Karrenberg, S. & Favre, A. (2008). Genetic and ecological differentiation in the hybridizing campions Silene dioica and S. latifolia. Evolution, 62, 763–73.CrossRefGoogle ScholarPubMed
Karron, J. D.et al. (2012) New perspectives on the evolution of plant mating systems. Annals of Botany, 109, 493–503.CrossRefGoogle ScholarPubMed
Katori, T.et al. (2010). Dissecting the genetic control of natural variation in salt tolerance of Arabidopsis thaliana accessions. Journal of Experimental Botany, 61, 1125–38.CrossRefGoogle ScholarPubMed
Kauffman, M. J.Brodie, J. F. & Jules, E. S. (2010). Are wolves saving Yellowstone's aspen? A landscape-level test of a behaviorally mediated trophic cascade. Ecology, 91, 2742–55.CrossRefGoogle Scholar
Kawata, M. I., Murakami, K. & Ishikawa, T. (2009). Dispersal and persistence of genetically modified oilseed rape around Japanese harbors. Environmental Science & Pollution Research International, 16, 120–6.CrossRefGoogle ScholarPubMed
Kay, Q. O. N. (1978). The role of preferential and assortative pollination in the maintenance of flower colour polymorphisms. In The pollination of flowers by insects, ed. Richards, A. J., pp. 175–90. Linnean Society Symposium Series 6. London: Academic Press.Google Scholar
Kearey, P., Klepeis, K. A. & Vine, F. J. (2009). Global tectonics, edn. Oxford: Wiley-Blackwell.Google Scholar
Keeling, P. J. & Palmer, J. D. (2008). Horizontal gene transfer in eukaryotic evolution. Nature Reviews Genetics, 9, 605–18.CrossRefGoogle ScholarPubMed
Keller, S. R. & Taylor, D. R. (2008). History, chance, and adaptation during biological invasion: separating stochastic phenotypic evolution from response to selection. Ecology Letters, 8, 852–6.Google Scholar
Kellogg, E. A. (2002) Are macroevolution and microevolution quantitatively different? Evidence from the Poaceae. In Developmental genetics and plant evolution, ed. Cronk, Q. C. B., Bateman, R. M. & Hawkin, J. A., pp. 70–84. London: Taylor & Francis.Google Scholar
Kelly, L. J. & Leitch, I. J. (2011). Exploring giant plant genomes with next-generation sequencing technology. Chromosome Research, 19, 939–53.CrossRefGoogle ScholarPubMed
Kelvin, Lord (Thompson, W.) (1871). On geological time. Transactions of the Geological Society of Glasgow, 3, 1–28.Google Scholar
Kemp, J., Milne, R. & Reay, D. S. (2010). Sceptics and deniers of climate change not to be confused. Nature, 464, 673.CrossRefGoogle Scholar
Kemp, T. S. (2005). The origin and evolution of mammals. New York: Oxford University Press.Google Scholar
Kemperman, J. A. & Barnes, B. V. (1976). Clone size in American aspens. Canadian Journal of Botany, 54, 2603–7.CrossRefGoogle Scholar
Kendall, M. G. & Plackett, R. L. (1977). Studies in the history of statistics and probability, vol. II. London: Griffin.Google Scholar
Kenrick, P. (2011). Timescales and timetrees. New Phytologist, 192, 3–6.CrossRefGoogle ScholarPubMed
Kerner, A. (1895). The natural history of plants, their forms, growth, reproduction and distribution. Translated and edited by Oliver, F. W.. London: Blackie.CrossRefGoogle Scholar
Kernick, M. D. (1961). Seed production of specific crops. In Agricultural and horticultural seeds, pp. 181–547. FAO Agriculture Studies No. 55.Google Scholar
Kettle, C. J.et al. (2012). Importance of demography and dispersal for the resilience and restoration of a critically endangered tropical conifer Araucaria nemorosa. Diversity and Distributions, 18, 248–59.CrossRefGoogle Scholar
Keymer, R. & Ellis, W. M. (1978). Experimental studies on plants of Lotus corniculatus L. from Anglesey polymorphic for cyanogenesis. Heredity, 40, 189–206.CrossRefGoogle Scholar
Keynes, M. (1993). Sir Francis Galton, FRS. The legacy of his ideas. London: Macmillan.CrossRefGoogle Scholar
Khandelwal, S. (1990). Chromosome evolution in the genus Ophioglossum L. Botanical Journal of the Linnean Society, 102, 205–17.
Khoury, C., Laliberté, B. & Guarino, L. (2010). Trends in ex situ conservation of plant genetic resources: a review of global crop and regional conservation strategies. Genetic Resources and Crop Evolution, 57, 625–39.CrossRefGoogle Scholar
Kihara, H. & Ono, T. (1926). Chromosomenzahlen und systematische Gruppierung der Rumex-Arten. Zeitschrift für Zellforschung und mikroskopische Anatomie, Berlin, Wien, 4, 475–81.Google Scholar
Kikuchi, R.et al. (2010). Disjunct distribution of chloroplast DNA haplotypes in the understory perennial Veratrum album ssp. oxysepalum (Melanthiaceae) in Japan as a result of ancient introgression. New Phytologist, 188, 879–91.CrossRefGoogle ScholarPubMed
Kilian, J.et al. (2007). The AtGenExpress global stress expression data set: protocols, evaluation and model data analysis of UV-B light, drought and cold stress responses. Plant Journal, 5, 347–63.Google Scholar
Kim, K. C. & Byrne, L. B. (2006). Biodiversity loss and the taxonomic bottleneck: emerging biodiversity science. Ecological Research, 21, 794–810.CrossRefGoogle Scholar
Kim, S.-C., Crawford, D. J. & Jansen, R. K. (1996). Phylogenetic relationships among the genera of the subtribe Sonchinae (Asteraceae): evidence from ITS sequences. Systematic Botany, 21, 417–32.CrossRefGoogle Scholar
Kim, S.-C.et al. (1996). A common origin for woody Sonchus and five related genera in the Macaronesian islands: molecular evidence for extensive radiation. Proceedings of the National Academy of Sciences, USA, 93, 7743–8.CrossRefGoogle ScholarPubMed
Kimber, G. & Athwal, R. S. (1972) A reassessment of the course of evolution of Wheat. Proceedings of the National Academy of Sciences, USA, 69, 912–15.CrossRefGoogle ScholarPubMed
Kimura, M. (1983). The neutral theory of molecular evolution. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Kimura, M. (1987). Molecular evolutionary clock and neutral theory. Journal of Molecular Evolution, 26, 24–33.CrossRefGoogle ScholarPubMed
King, M. (1993). Species evolution, the role of chromosome change. Cambridge: Cambridge University Press.Google Scholar
King, R. C., Stansfield, W. D. & Mulligan, P. K. (2006). A dictionary of genetics, edn. Oxford: Oxford University Press.Google Scholar
Kirk, J. T. O. & Tilney-Bassett, R. A. E. (1978). The plastids. Their chemistry, structure, growth and inheritance, edn. Amsterdam: Elsevier.Google Scholar
Kirschbaum, M. U. F. (1995). The temperature dependence of soil organic matter decomposition, and the effect of global warming on soil organic C storage. Soil Biology & Biochemistry, 27, 753–60.CrossRefGoogle Scholar
Kitajima, K.et al. (2006). Cultivar selection prior to introduction may increase invasiveness: evidence from Ardisia crenata. Biological Invasions, 8, 1471–82.CrossRefGoogle Scholar
Klaas, M.et al. (2011) Progress towards elucidating the mechanisms of self-incompatibility in the grasses: further insights from studies in Lolium. Annals of Botany, 108, 677–85.CrossRefGoogle ScholarPubMed
Klein, J. T. & Kadereit, J. W. (2015). Phylogeny, biogeography and evolution of edaphic association in the European oreophytes Sempervivum and Jovibarba (Crassulaceae). International Journal of Plant Science, 176, 44–71.CrossRefGoogle Scholar
Kliman, R., Sheehy, B. & Schultz, J. (2008). Genetic drift and effective population size. Nature Education, 1, 3.Google Scholar
Klug, A. (2004). The discovery of the DNA double helix. Journal of Molecular Biology, 335, 3–26.CrossRefGoogle ScholarPubMed
Knapp, S. (2008). Naming nature: the future of the Linnaean system. The Linnean Special Issue, 8, 161–7.Google Scholar
Knapp, S. (2010). Four new vining species of Solanum (Dulcamaroid clade) from montane habitats in Tropical America. PLoS ONE, 5(5), e10502.CrossRefGoogle ScholarPubMed
Knapp, S.et al. (2002) Taxonomy needs evolution, not revolution. Nature, 419, 559.CrossRefGoogle Scholar
Knapp, S., Polaszek, A. & Watson, M. (2007). Spreading the word. Nature, 446, 261–2.CrossRefGoogle Scholar
Knight, T. M.et al. (2005). Pollen limitation of plant reproduction. Pattern and process. Annual Review of Ecology, Evolution and Systematics, 36, 467–97.CrossRefGoogle Scholar
Knispel, A. L. & McLachlan, S. M. (2010). Landscape-scale distribution and persistence of genetically modified oilseed rape (Brassica napus) in Manitoba, Canada. Environmental Science & Pollution Research International, 17, 13–25.CrossRefGoogle Scholar
Knispel, A. L.et al. (2008). Gene flow and multiple herbicide resistance in escaped canola populations. Weed Science, 56, 72–80.CrossRefGoogle Scholar
Knobloch, I. W. (1971). Intergeneric hybridization in flowering plants. Taxon, 21, 97–103.Google Scholar
Knox, R. B. & Heslop-Harrison, J. (1963). Experimental control of aposporous apomixis in a grass of the Andropogoneae. Botaniska Notiser, 116, 127–41.Google Scholar
Koch, C. & Kollmann, J. (2012). Clonal reintroduction of endangered plant species: the case of the German False Tamarisk in pre-Alpine rivers. Environmental Management, 50, 217–25.CrossRefGoogle Scholar
Koerner, L. (1999). Linnaeus: nature and nation. Cambridge, MA, & London: Harvard University Press.Google Scholar
Kohn, D. (1981). On the origin of the principle of diversity. Science, 213, 1105–8.CrossRefGoogle ScholarPubMed
Kohn, D. (1985a). The Darwinian heritage. Princeton, NJ: Princeton University Press.Google Scholar
Kohn, D. (1985b). Darwin's principle of divergence as internal dialogue. In The Darwinian heritage, ed. Kohn, D., pp. 245–57. Princeton, NJ: Princeton University Press.Google Scholar
Kollmann, J. & Bañuelos, M. J. (2004). Latitudinal trends in growth and phenology of the invasive alien plant Impatiens glandulifera (Balsaminaceae). Diversity & Distributions, 10, 377–85.CrossRefGoogle Scholar
Koornneef, M. & Schere, B. (2007) Arabidopsis thaliana as an experimental organism. In Handbook of plant science, ed. Roberts, K., pp. 684–9. Chichester: Wiley.Google Scholar
Kooyers, N. J. & Olsen, K. M (2012). Rapid evolution of an adaptive cyanogenesis cline in introduced North American white clover (Trifolium repens L.). Molecular Ecology, 21, 2455–68.CrossRefGoogle Scholar
Kooyers, N. J.et al. (2014). Aridity shapes cyanogenesis cline evolution in white clover (Trifolium repens L.). Molecular Ecology, 23, 1053–70.
Koshy, T. K. (1968). Evolutionary origin of Poa annua L. in the light of karyotypic studies. Canadian Journal of Genetics and Cytology, 10, 112–18.CrossRefGoogle Scholar
Kovác, J. (1995). Micropropagation of Dianthus arenarius subsp. bohemicus – an endangered endemic from the Czech Republic. Botanic Gardens Micropropagation News, 1, 105–7.Google Scholar
Kramer, A. T. & Havens, K. (2009). Plant conservation genetics in a changing world. Trends in Plant Science, 14, 599–607.CrossRefGoogle Scholar
Krämer, U. (2010). Metal hyperaccumulation in plants. Annual Review of Plant Biology, 61, 517–34.CrossRefGoogle ScholarPubMed
Krauss, S. L., Dixon, B. & Dixon, K. W. (2002). Rapid genetic decline in a translocated population of the endangered plant Grevillea scapigera. Conservation Biology, 16, 986–94.CrossRefGoogle Scholar
Kremer, A.et al. (2012). Long-distance gene flow and adaptation of forest trees to rapid climate change. Ecological Letters, 15, 378–92.CrossRefGoogle ScholarPubMed
Krishnamurphy, P. K. & Francis, R. A. (2012). A critical review of the utility of DNA barcoding in biodiversity conservation. Biodiversity and Conservation, 21, 1901–19.Google Scholar
Kroymann, J.et al. (2003). Evolutionary dynamics of an insect resistance quantitative trait locus. Proceedings of the National Academy of Sciences, USA, 100, suppl. 2, 14587–92.CrossRefGoogle ScholarPubMed
Kruckeberg, A. R. (1951). Intraspecific variability in the response of certain native plant species to serpentine soil. American Journal of Botany, 38, 408–19.CrossRefGoogle Scholar
Kruckeberg, A. R. (1954). The ecology of serpentine soils. III. Plant species in relation to serpentine soils. Ecology, 35, 267–74.Google Scholar
Krushelnycky, P. D., Loope, L. L., Reimer, N. J. (2005). The ecology, policy, and management of ants in Hawaii. Proceedings of the Hawaiian Entomological Society, 37, 1–25.Google Scholar
Kühl, S. (2013). For the betterment of the race. Translated by Schofer, L.. New York: Palgrave Macmillan.CrossRefGoogle Scholar
Kuhn, T. S., Mooers, A. Ø. & Thomas, G. H. (2011). A simple polytomy resolver for dated phylogenies. Methods in Ecology and Evolution, 2, 427–36.CrossRefGoogle Scholar
Kujala, S. T. & Savolainen, O. (2012). Sequence variation patterns along a latitudinal cline in Scots Pine (Pinus sylvestris): signs of clinal adaptation?Tree Genetics & Genomes, 8, 1451–67.CrossRefGoogle Scholar
Kukkala, A. S. & Moilanen, A. (2013). Core concepts of spatial prioritization in systematic conservation planning. Biological Reviews, 88, 443–64.CrossRefGoogle Scholar
Kulpa, S. M. & Leger, E. A. (2012) Strong natural selection during plant restoration favors an unexpected suite of plant traits. Ecological Applications, 6, 510–23.Google Scholar
Kunz, W. (2012). Do species exist? Principles of taxonomic classification. Weinheim: Wiley-Blackwell.CrossRefGoogle Scholar
Küster, E. (1899). Über Stammverwachsungen. Jahrbücher für Wissenschaftliche Botanik, Berlin, 33, 487–512.Google Scholar
Kuta, E.et al. (2004). Chromosome and nuclear DNA study on Luzula – a genus with holokinetic chromosomes. Genome, 47, 246–56.CrossRefGoogle ScholarPubMed
Kyhos, D. W. (1965). The independent aneuploid origin of two species of Chaenactis (Compositae) from a common ancestor. Evolution, 19, 26–43.CrossRefGoogle Scholar
Ladizinsky, G. (1998). Plant evolution under domestication. Dordrecht: Kluwer Academic Publishers.CrossRefGoogle Scholar
Lai, Z.et al. (2005). Identification and mapping of SNPs from ESTs in sunflower. Theoretical and Applied Genetics, 111, 1532–44.CrossRefGoogle ScholarPubMed
Lake, J. A. & Rivera, M. C. (1994). The prokaryotic origin of eukaryotes. In Evolution of microbial life, ed. Roberts, A.et al. Cambridge: Cambridge University Press.Google Scholar
Laland, K. N. (2002). Niche construction. In Encyclopedia of evolution, ed. Patel, M., pp. 821–3. Oxford & New York: Oxford University Press.Google Scholar
Lamarck, J. B. (1809). Philosophie zoologique. English translation, Zoological philosophy, translated by Elliot, H., published 1914, London & New York: Macmillan.Google Scholar
Lamb, H. H. (1970). Our changing climate. In Flora of a changing Britain, ed. Perring, F. H., pp. 11–24. Hampton, Middlesex: Botanical Society of the British Isles.Google Scholar
Lambers, H., Chapin, F. S. & Pons, T. L. (2008). Plant physiological ecology. New York: Springer.CrossRefGoogle Scholar
Lamprecht, H. (1961). Die Genekarte von Pisum bei normaler Struktur der Chromosomen. Agri Hortique Genetica, 19, 360–401.Google Scholar
Lande, R. (1988). Genetics and demography in biological conservation. Science, 241, 1455–60.CrossRefGoogle ScholarPubMed
Lane, C. (1962). Notes on the Common Blue (Polyommatus icarus) egg laying and feeding on the cyanogenic strains of the Bird's-foot Trefoil (Lotus corniculatus). Entomologist's Gazette, 13, 112–16.Google Scholar
Langerhans, R. B. & Riesch, R. (2013). Speciation by selection: a framework for understanding ecology's role in speciation. Current Zoology, 59, 31–52.CrossRefGoogle Scholar
Langlet, O. (1934). Om variationen hos tallen Pinus sylvestris och dess samband med climatet. Meddelanden från Statens Skogsførsøksanstalt, 27, 87–93.Google Scholar
Langlet, O. (1971). Two hundred years genecology. Taxon, 20, 653–722.CrossRefGoogle Scholar
Lankau, R. A. (2012). Coevolution between invasive and native plants driven by chemical competition and soil biota. Proceedings of the National Academy of Sciences, USA, 109, 11240–5.CrossRefGoogle ScholarPubMed
Lankau, R. A.et al. (2009). Evolutionary limits ameliorate the negative impact of an invasive plant. Proceedings of the National Academy of Sciences, USA, 106, 15362–7.CrossRefGoogle ScholarPubMed
Lankester, E. (1848). The correspondence of John Ray. London: Ray Society, British Museum.Google Scholar
LAPG III (2009). The Linear Angiosperm Phylogeny Group (LAPG) III: a linear sequence of the families in APG III. Botanical Journal of the Linnean Society, 161, 128–31.
Larcher, W. (2003). Physiological plant ecology, edn. Berlin: Springer.CrossRefGoogle Scholar
Larsen, E. C. (1947). Photoperiodic responses of geographical strains of Andropogon scoparius. Botanical Gazette, 109, 132–50.CrossRefGoogle Scholar
Lasso, E. (2008). The importance of setting the right genetic distance threshold for identification of clones using amplified fragment length polymorphism: a case study with five species in the tropical plant genus Piper. Molecular Ecology Resources, 8, 74–82, DOI: 10.1111/j.1471-8286.2007.01910.xCrossRefGoogle ScholarPubMed
Laurie, D. A. & Bennett, M. D. (1985). Nuclear DNA content in the genera Zea and Sorghum: intergeneric, interspecific and intraspecific variation. Heredity, 55, 307–13.CrossRefGoogle Scholar
Lauterbach, D., Burkart, M. & Gemeinholzer, B. (2012). Rapid genetic differentiation between ex situ and their in situ source populations: an example of the endangered Silene otitis (Caryophyllaceae). Botanic Journal of the Linnean Society, 168, 64–75.Google Scholar
Lavergne, S. & Molofsky, J. (2007). Increased genetic variation and evolutionary potential drive the success of an invasive grass. Proceedings of the National Academy of Sciences, USA, 104, 3883–8.CrossRefGoogle ScholarPubMed
Lawrence, B. A. & Kaye, T. N. (2011). Reintroduction of Castilleja levisecta: effects of ecological similarity, source population genetics and habitat quality. Restoration Ecology, 19, 166–76.CrossRefGoogle Scholar
Lawrence, W. E. (1945). Some ecotypic relations of Deschampsia caespitosa. American Journal of Botany, 32, 298–314.CrossRefGoogle Scholar
Lawrence, W. J. C. (1950). Science and the glasshouse. Edinburgh & London: Oliver & Boyd.Google Scholar
Le Comber, S. C.et al. (2010). Making a functional diploid: from polysomic to disomic inheritance. New Phytologist, 186, 113–22.CrossRefGoogle ScholarPubMed
Le Page, M. (2012). A brief history of the genome. New Scientist, 2882, 30–5.Google Scholar
Lebaron, H. M. & Gressel, J. (eds.) (1982). Herbicide resistance in plants. New York: Wiley.Google Scholar
Ledig, F. T., Rehfeldt, G. E. & Jaquish, B. (2012). Projections of suitable habitat under climate change scenarios: implications for trans-boundary assisted colonization. American Journal of Botany, 99, 1217–30.CrossRefGoogle ScholarPubMed
Lee, A. (1902). Dr Ludwig on variation and correlation in plants. Biometrika, 1, 316–19.CrossRefGoogle Scholar
Lee, P. L. M.et al. (2004). Comparison of genetic diversities in native and alien populations of hoary mustard (Hirschfeldia incana [L.] Lagreze-Fossat). International Journal of Plant Sciences, 165, 833–43.CrossRefGoogle Scholar
Lefèbvre, C. (1973). Outbreeding and inbreeding in a zinc–lead mine population of Armeria maritima. Nature, 243, 96–7.CrossRefGoogle Scholar
Lemey, P., Salemi, M. & Vandamme, A.-M. (eds.) (2009). The phylogenetic handbook: a practical approach to phylogenetic analysis and hypothesis testing. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Lengyel, S.et al. (2010). Convergent evolution of seed dispersal by ants, and phylogeny and biogeography in flowering plants: a global survey. Perspectives in Plant Ecology, Evolution and Systematics, 12, 43–55.CrossRefGoogle Scholar
Lenz, J.et al. (2011). Seed-dispersal distributions by trumpeter hornbills in fragmented landscapes. Proceedings of the Royal Society, B, 278, 2257–64.CrossRefGoogle ScholarPubMed
Lenzen, M.et al. (2012). International trade drives biodiversity threats in developing nations. Nature, 486, 109–12.CrossRefGoogle ScholarPubMed
Leon-Lobos, P.et al. (2012). The role of ex situ seed banks in the conservation of plant diversity and in ecological restoration in Latin America. Plant Ecology & Diversity, 5, 245–58.CrossRefGoogle Scholar
Lev-Yadun, S. & Inbar, M. (2002). Defensive ant, aphid and caterpillar mimicry in plants?Biological Journal of the Linnean Society, 77, 393–8.CrossRefGoogle Scholar
Levan, A. (1938). The effect of colchicine on root mitosis in Allium. Hereditas, 24, 471–86.Google Scholar
Levey, D. J. & Sargent, S. (2000). A simple method for tracking vertebrate-dispersed seeds. Ecology, 81, 267–74.CrossRefGoogle Scholar
Levin, D. A. (1973). The role of trichomes in plant defense. Quarterly Review of Biology, 48, 3–15.CrossRefGoogle Scholar
Levin, D. A. (1975a). Minority cytotype exclusion in local plant populations. Taxon, 24, 35–43.CrossRefGoogle Scholar
Levin, D. A. (1975b). Pest pressure and recombination systems in plants. American Naturalist, 190, 437–51.Google Scholar
Levin, D. A. (1978a). Pollinator behaviour and the breeding structure of plant populations. In The pollination of flowers by insects, ed. Richards, A. J., pp. 133–50. Linnean Society Symposium Series 6. London: Academic Press.Google Scholar
Levin, D. A. (1978b). The origin of isolating mechanisms in flowering plants. Evolutionary Biology, 11, 185–317.Google Scholar
Levin, D. A. (1979). The nature of plant species. Science, 204, 381–4.CrossRefGoogle ScholarPubMed
Levin, D. A. (1984). Immigration in plants: an exercise in the subjunctive. In Perspectives on plant population ecology, ed Dirzo, R. & Sarukhán, J., pp. 242–60. Sunderland, MA: Sinauer.Google Scholar
Levin, D. A. (1985). Reproductive character displacement in Phlox. Evolution, 39, 1275–81.CrossRefGoogle ScholarPubMed
Levin, D. A. (1988). Local differentiation and the breeding structure of plant populations. In Plant evolutionary biology, ed. Gottlieb, L. D. & Jain, S. K., pp. 305–29. London: Chapman & Hall.Google Scholar
Levin, D. A. (1993). Local speciation in plants: the rule not the exception. Systematic Botany, 18, 197–208.CrossRefGoogle Scholar
Levin, D. A. (2001a) The recurrent origin of plant races and species. Systematic Botany, 26, 197–204.Google Scholar
Levin, D. A. (2001b). 50 years of plant speciation. Taxon, 50, 69–91.CrossRefGoogle Scholar
Levin, D. A. (2003). The ecological transition in speciation. New Phytologist, 161, 91–6.CrossRefGoogle Scholar
Levin, D. A. (2011). Mating system shifts on the trailing edge. Annals of Botany, 2011, 1–8, DOI:10.1093/aob/mcr159.CrossRef
Levin, D. A. & Kerster, H. W. (1967). An analysis of interspecific pollen exchange in Phlox. American Naturalist, 101, 387–400.CrossRefGoogle Scholar
Levin, D. A. & Kerster, H. W. (1974). Gene flow in seed plants. Evolutionary Biology, 7, 139–220.Google Scholar
Levin, S. A.et al. (2003). The ecology and evolution of dispersal: a theoretical perspective. Annual Review of Ecology, Evolution and Systematics, 34, 575–604.CrossRefGoogle Scholar
Lewis, D. (1979). Sexual incompatibility in plants. Studies in Biology, No. 110. London: Arnold.Google Scholar
Lewis, D. & Crowe, L. K. (1956). The genetics and evolution of gynodioecy. Evolution, 10, 115–25.CrossRefGoogle Scholar
Lewis, H. (1973). The origin of diploid neospecies in Clarkia. The American Naturalist, 107, 161–70.CrossRefGoogle Scholar
Lewis, H & Lewis, M. E. (1955). The genus Clarkia. Berkeley & Los Angeles: University of California Press.Google Scholar
Lewis, W. H. (1976). Temporal adaptation correlated with ploidy in Claytonia virginica. Systematic Botany, 1, 340–7.CrossRefGoogle Scholar
Lewis, W. H. (ed.) (1980a). Polyploidy. London & New York: Plenum Press.CrossRefGoogle Scholar
Lewis, W. H. (1980b). Polyploidy in species populations. In Polyploidy, ed. Lewis, W. H., pp.103–44. New York & London: Plenum Press.CrossRefGoogle Scholar
Lewis, W. H. (1980c). Polyploidy in angiosperms: dicotyledons. In Polyploidy, ed. Lewis, W. H., pp. 241–68. New York & London: Plenum Press.CrossRefGoogle Scholar
Li, D.-Z. & Pritchard, H. W. (2009). The science and economics of ex situ plant conservation. Trends in Plant Science, 14, 614–21.CrossRefGoogle ScholarPubMed
Li, H. J. & Zhang, Z. B. (2003). Effect of rodents on acorn dispersal and survival of the Liaodong oak (Quercus liaotungensis Koidz.). Forest Ecology & Management, 176, 387–96.CrossRefGoogle Scholar
Li, H. J. & Zhang, Z. B. (2007). Effects of mast seeding and rodent abundance on seed predation and dispersal by rodents in Prunus armeniaca (Rosaceae). Forest Ecology & Management, 242, 511–17.CrossRefGoogle Scholar
Li, H. L. (1956). The story of the cultivated Horse-Chestnuts. Morris Arboretum Bulletin, 7, 35–9.Google Scholar
Li, X.et al. (2015). Plant DNA barcoding: from gene to genome. Biological Reviews, 90, 157–66.CrossRefGoogle Scholar
Lieberman, B. S. & Kaesler, R. (2010). Prehistoric life. Oxford: Wiley-Blackwell.Google Scholar
Liepelt, S., Bialozyt, R. & Ziegenhagen, B. (2002). Wind-dispersed pollen mediates postglacial gene flow in refugia. Proceedings of the National Academy of Sciences, USA, 99, 14590–4.CrossRefGoogle ScholarPubMed
Lihová, J.et al. (2004). Origin of the disjunct tetraploid Cardamine amporitana (Brassicaceae) assessed with nuclear and chloroplast DNA sequence data. American Journal of Botany, 9, 1231–42.Google Scholar
Liljefors, A. (1953). Studies on propagation, embryology and pollination in Sorbus. Acta Horti Bergiani, 16, 227–329.Google Scholar
Liljefors, A. (1955). Cytological studies in Sorbus. Acta Horti Bergiani, 17, 47–113.Google Scholar
Lim, K. Y.et al. (2007). Parental origin and genome evolution in the allopolyploid Iris versicolor. Annals of Botany, 100, 219–24.CrossRefGoogle ScholarPubMed
Lindahl, B. D.et al. (2013). Fungal community analysis by high-throughput sequencing of amplified markers – a user guide. New Phytologist, 199, 288–99.CrossRefGoogle Scholar
Linde, M., Diel, S. & Neuffer, B. (2001). Flowering ecotypes of Capsella bursa-pastoris (L.) Medik. (Brassicaceae) analysed by a co-segregation of phenotypic characters (QTL) and molecular markers. Annals of Botany, 87, 91–9.CrossRefGoogle Scholar
Linder, C. R. & Rieseberg, L. H. (2004). Reconstructing patterns of reticulate evolution in plants. American Journal of Botany, 91, 1700–8.CrossRefGoogle ScholarPubMed
Linhart, V. B. & Baker, I. (1973). Intra-population differentiation of physiological response to flooding in a population of Veronica peregrina. Nature, 242, 275–6.CrossRefGoogle Scholar
Linhart, Y. B. (1988). Intra-population differentiation in annual plants III. The contrasting effects of intra- and inter-specific competition. Evolution, 42, 1047–64.CrossRefGoogle Scholar
Linhart, Y. B., Mitton, J. B., Sturgeon, K. B. & Davis, M. L. (1981). Genetic variation in space and time in a population of Ponderosa Pine. Heredity, 46, 407–26.CrossRefGoogle Scholar
Linington, S. H. & Pritchard, H. W. (2001). Gene banks. In Encyclopedia of biodiversity, vol. 3, ed. Levin, S. A., pp.165–81. San Diego, CA, & London: Academic Press.Google Scholar
Linnaeus, C. (1737, but not distributed until 1738). Hortus cliffortianus. Amsterdam.Google Scholar
Linnaeus, C. (1737). Critica botanica. (English translation by Hort, A., 1938, London: Ray Society, British Museum.)Google Scholar
Linnaeus, C. (1744). Peloria. In Amoenitates academicae (1749–90). (See Stearn, (1957) for details of the many editions.)
Linnaeus, C. (1749–90). Amoenitates academicae. (For details of the many editions see Stearn, (1957).)
Linnaeus, C. (1751). Linnaeus’ Philosophia botanica. Translated by Freer, S. (2003). Oxford: Oxford University Press.Google Scholar
Linnaeus, C. (1751). Philosophia botanica. Stockholm.Google Scholar
Linnaeus, C. (1753). Species plantarum. (Facsimile edn 1957, London: Ray Society, British Museum.)Google Scholar
Linnaeus, C. (1762–3). Species plantarum, edn. Stockholm.Google Scholar
Linroth, S. (1983).The two faces of Linnaeus. In Linnaeus. The man and his work, ed. Frangsmyr, T., pp. 1–62. Berkeley: University of California Press.Google Scholar
Lipman, M. J.et al. (2013). Natural hybrids between Tragopogon mirus and T. miscellus (Asteraceae): a new perspective on karyotype changes following hybridization at the polyploid level. American Journal of Botany, 100, 2016–22.CrossRefGoogle Scholar
Lipscomb, D. L., Platnick, N. and Wheeler, Q. (2003). The intellectual content of taxonomy: a comment on DNA taxonomy. Trends in Ecology and Evolution, 18, 65–6.CrossRefGoogle Scholar
Litardière, R. (1939). Sur les caractères chromosomiques et la systématique des Poa du group du P. annua L. Revue de Cytologie et de Cytophysiologie végétales, 4, 82–5.
Lloyd, D. G. (1965). Evolution of self-compatibility and racial differentiation in Leavenworthia (Cruciferae). Contributions from the Gray Herbarium, 195, 1–134.Google Scholar
Lloyd, D. G. (1975). The maintenance of gynodioecy and androdioecy in angiosperms. Genetica, 45, 325–39.CrossRefGoogle Scholar
Lloyd, D. G. (1979a). Some reproductive factors affecting the selection of self-fertilisation in plants. American Naturalist, 113, 67–79.CrossRefGoogle Scholar
Lloyd, D. G. (1979b). Evolution towards dioecy in heterostylous populations. Plant Systematics and Evolution, 131, 71–80.CrossRefGoogle Scholar
Lloyd, D. G. (1992). Self- and cross-fertilization in plants. II. The selection of self-fertilization. International Journal of Plant Science, 153, 370–80.Google Scholar
Lloyd, D. G. & Webb, C. J. (1992). The evolution of heterostyly. In Evolution and function of heterostyly, ed. Barrett, S. C. H., pp. 151–78. Heidelberg: Springer-Verlag.Google Scholar
Lo, E. Y. Y., Stefanović, S. & Dickinson, T. A. (2009). Population genetic structure of diploid sexual and polyploidy apomictic hawthorns (Crateagus; Rosaceae) in the Pacific Northwest. Molecular Ecology, 18, 1145–60.CrossRefGoogle Scholar
Lo, E. Y. Y., Stefanović, S. & Dickinson, T. A (2010). Reconstructing reticulation history in a phylogenetic framework and the potential of allopatric speciation driven by polyploidy in an agamic complex in Crataegus (Rosaceae). Evolution, 64, 3593–608.CrossRefGoogle Scholar
Lockwood, J. & McKinney, M. (eds.) (2001). Biotic homogenization. New York: Kluwer Academic/Plenum Publishers.CrossRefGoogle Scholar
Lofflin, D. L. & Kephart, S. R. (2005). Outbreeding, seedling establishment, and maladaptation in natural and reintroduced populations of rare and common Silene douglasii (Caryophyllaceae). American Journal of Botany, 92, 1691–700.CrossRefGoogle Scholar
Lomolino, M. V.et al. (2005). Biogeography, edn. Sutherland, MA: Sinauer; London: Hatchette.Google Scholar
Lonsdale, W. M. (1994). Inviting trouble: introduced pasture species in northern Australia. Australian Journal of Ecology, 19, 345–54.Google Scholar
Lookerman, D. J. & Jansen, R. K. (1995). The use of herbarium material for DNA studies. In Sampling the green world, ed. Stuessy, T. F., pp. 205–20. New York: Columbia University Press.Google Scholar
Löve, A. (1960). Biosystematics and the classification of apomicts. Feddes Repertorium, 62, 136–48.Google Scholar
Löve, A. (1962). The biosystematic species concept. Preslia, 34, 127–39.Google Scholar
Löve, A. & Löve, D. (1961). Chromosome numbers of Central and Northwest European plant species. Opera Botanica, 5, 1–581.Google Scholar
Löve, A. & Löve, D. (1974). Origin and evolution of the arctic and alpine floras. In Arctic and Alpine environments, ed. Ives, J. D. & Barry, R. G., pp. 571–603. London: Methuen.Google Scholar
Lovejoy, A. O. (1966). The Great Chain of Being. Cambridge, MA: Harvard University Press.Google Scholar
Lovett Doust, J. & Lovett Doust, L. (1988). Plant reproductive ecology, patterns and strategies. New York: Oxford University Press.Google Scholar
Lövkvist, B. (1956). The Cardamine pratensis complex. Symbolae Botanicae Upsalienses, 14(2), 1–131.Google Scholar
Lövkvist, B. (1962). Chromosome and differentiation studies in flowering plants of Skåne, South Sweden. 1. General aspects. Type species with coastal differentiation. Botaniska Notiser, 115, 261–87.Google Scholar
Lowe, A., Harris, S. & Ashton, P. (2009). Ecological genetics: design, analysis and application. London: Wiley.Google Scholar
Lowe, S.et al. (2000). 100 of the world's worst invasive alien species: a selection from the Global Invasive Species Database. Auckland: The Invasive Species Specialist Group (ISSG), a specialist group of the Species Survival Commission (SSC) of the World Conservation Union (IUCN).Google Scholar
Lowry, D. B. (2012). Ecotypes and the controversy over stages in the formation of new species. Biological Journal of the Linnean Society, 106, 241–57.CrossRefGoogle Scholar
Lowry, D. B. (2012). Local adaptation in the model plant. New Phytologist, 194, 888–90.CrossRefGoogle ScholarPubMed
Lowry, D. B., Rockwood, R. C. & Willis, J. H. (2008). Ecological reproductive isolation of coast and inland races of Mimulus guttatus. Evolution, 62, 2196–214.CrossRefGoogle ScholarPubMed
Luceño, M. & Castroviejo, S. (1991). Agamatoploidy in Carex laevigata (Cyperaceae). Fusion and fission of chromosomes as the mechanism of cytogenetic evolution in Iberian populations. Plant Systematics & Evolution, 177, 149–59.CrossRefGoogle Scholar
Luckow, M. (1995). Species concepts : assumptions, methods and applications. Systematic Botany, 20, 589–605.CrossRefGoogle Scholar
Ludwig, F. (1895). Uber Variationskurven und Variationsflächen der Pflanzen. Botanisches Zentralblatt, 64, 1–8, 33–41, 65–72, 97–105.Google Scholar
Ludwig, F. (1901). Variationsstatistische Probleme und Materialen. Biometrika, 1, 11–29.CrossRefGoogle Scholar
Lui, B. & Wendel, J. F. (2003). Epigenetic phenomena and the evolution of plant allopolyploids. Molecular Phylogenetics and Evolution, 29, 365–79.Google Scholar
Lui, B.et al. (2009). Rapid genomic changes in polyploid wheat and related species: implications for genome evolution and genetic improvement. Journal of Genetics & Genomics, 36, 519–28.Google Scholar
Lui, G.et al. (2006). Comparison of genetic variability on populations of wild rice, Oryza rufipogon, plants and their soil seed banks. Conservation Genetics, 7, 909–17.Google Scholar
Lui, L.et al. (2009). Coalescent methods for estimating phylogenetic trees. Molecular Phylogenetics and Evolution, 53, 320–8.Google Scholar
Lumaret, R. (1984). The role of polyploidy in the adaptive significance of polymorphism at the GOT I locus in the Dactylis glomerata complex. Heredity, 52, 153–69.CrossRefGoogle Scholar
Lumaret, R. (1988). Cytology, genetics and evolution in the genus Dactylis. Critical Reviews in Plant Sciences, 7, 55–91.CrossRefGoogle Scholar
Lumaret, R. & Ouazzani, N. (2001). Ancient wild olives in Mediterranean forests. Nature, 413, 700.CrossRefGoogle ScholarPubMed
Lynch, M. & Walsh, B. (1997). Genetics and analysis of quantitative traits. Sunderland, MA: Sinauer.Google Scholar
Lynch, R. L. (1900). Hybrid Cinerarias. Journal of the Royal Horticultural Society, 24, 269–74.Google Scholar
Lyons, E. & Freeling, M. (2008). How to usefully compare homologous plant genes and chromosomes as DNA sequences. The Plant Journal, 53, 661–73.CrossRefGoogle ScholarPubMed
Lysak, M. A. (2014). Live and let die: centromere loss during evolution of plant chromosomes. New Phytologist, 203, 1082–9.CrossRefGoogle Scholar
Lysak, M. A.et al. (2009). The dynamic ups and downs of genome size evolution in Brassicaceae. Molecular Biology & Evolution, 26, 85–98.Google ScholarPubMed
Mabey, R. (1980). The common ground. a place for nature in Britain's future. London: Hutchinson.Google Scholar
MacArthur, R. H. & Wilson, E. O. (1967). The theory of island biogeography. Princeton, NJ: Princeton University Press.Google Scholar
MacDonald, G. M. (2003). Biogeography: space, time and life. New York: Wiley.Google Scholar
Mace, G. M. (2005). Biodiversity: an index of intactness. Nature, 434, 32–3.CrossRefGoogle ScholarPubMed
Mace, G. M. & Purvis, A. (2008). Evolutionary biology and practical conservation: bridging a widening gap. Molecular Ecology, 17, 9–19.CrossRefGoogle ScholarPubMed
Mace, G. M., Possingham, H. P. & Leader-Williams, N. (2007). Prioritizing choices in conservation. In Key topics in conservation biology, ed. Macdonald, D. W. & Service, K., pp. 17–34. Oxford: Blackwell.Google Scholar
Maceira, N. O., De Haan, A. A., Lumaret, R., Billon, M. & Delay, J. (1992). Production of 2n gametes in diploid subspecies of Dactylis glomerata L. I. Occurrence and frequency of 2n pollen. Annals of Botany, 69, 335–43.CrossRefGoogle Scholar
MacIsaac, H. J.et al. (2004). Backcasting and forecasting biological invasions of inland lakes. Ecological Applications, 14, 773–83.CrossRefGoogle Scholar
Mack, R. N. (1985). Invading plants: their potential contribution to population biology. InStudies in plant demography, ed. White, J., pp. 127–42. London: Academic Press.Google Scholar
Mack, R. N. (1991). The commercial seed trade: an early disperser of weeds in the United States. Economic Botany, 45, 257–73.CrossRefGoogle Scholar
Mack, R. N. & Lonsdale, W. M. (2001). Humans as global dispersers: getting more than we bargained for. BioScience, 51, 95–102.CrossRefGoogle Scholar
Mack, R. N.et al. (2000). Biotic invasions: causes, epidemiology, global consequences and control. Ecological Applications, 10, 689–710.CrossRefGoogle Scholar
Macnair, M. R. (1983). The genetic control of copper tolerance in the Yellow Monkey Flower Mimulus guttatus. Heredity, 50, 283–93.CrossRefGoogle Scholar
Macnair, M. R. (1987). Heavy metal tolerance in plants: a model evolutionary system. Trends in Ecology & Evolution, 2, 354–9.CrossRefGoogle ScholarPubMed
Macnair, M. R. (1993). The genetics of metal tolerance in vascular plants. New Phytologist, 124, 541–59.CrossRefGoogle Scholar
Macnair, M. R. & Christie, P. (1983). Reproductive isolation as a pleiotropic effect of copper tolerance in Mimulus guttatus. Heredity, 50, 295–302.CrossRefGoogle Scholar
Macnair, M. R., Macnair, V. E. & Martin, B. E. (1989). Adaptive speciation in Mimulus: an ecological comparison of M. cupriphilus with its presumed progenitor, M. guttatus. New Phytologist, 112, 269–79.CrossRefGoogle Scholar
Macnair, M. R., Cumbes, Q. J. & Meharg, A. A. (1992). The genetics of arsenate tolerance in Yorkshire fog, Holcus lanatus L. Heredity, 69, 325–35.
Maddox, B. (2002). Rosalind Franklin: the dark lady of DNA. London: Harper Collins.Google Scholar
Maddox, D. G. (1989). Clone structure in Solidago altissima populations: rhizome connections within genotypes. American Journal of Botany, 76, 318–21.CrossRefGoogle Scholar
Madlung, A. & Wendel, J. F. (2013). Genetic and epigenetic aspects of polyploid evolution in plants. Cytogenetic and Genome Research, 140, 270–85.CrossRefGoogle ScholarPubMed
Madlung, A.et al. (2005). Genomic changes in synthetic Arabidopsis polyploids. Plant Journal, 41, 221–30.Google ScholarPubMed
Magallón, S. (2010). Using fossils to break long branches in molecular dating: a comparison of relaxed clocks applied to the origin of angiosperms. Systematic Biology, 59, 384–99.CrossRefGoogle ScholarPubMed
Magallón, S. & Castillo, A. (2009). Angiosperm diversification through time. American Journal of Botany, 96, 349–65.CrossRefGoogle ScholarPubMed
Magallón, S., Hilu, K. W. & Quandt, D. (2013). Land plant evolutionary timeline: gene effects are secondary to fossil constraints in relaxed clock estimation of age and substitution rates. American Journal of Botany, 100, 556–73.CrossRefGoogle ScholarPubMed
Magee, B. (1973). Popper. Glasgow: Collins.Google Scholar
Magri, D.et al. (2006). A new scenario for the Quaternary history of European beech populations: palaeobotanical evidence and genetic consequences. New Phytologist, 171, 199–221.CrossRefGoogle ScholarPubMed
Magri, D.et al. (2007), The distribution of Quercus suber chloroplast haplotypes matches the palaeogeographical history of the western Mediterranean. Molecular Ecology, 16, 5259–66.CrossRefGoogle ScholarPubMed
Majeský, L.et al. (2012). The pattern of genetic variability in apomictic clones of Taraxacum officinale indicates the alternation of asexual and sexual histories of apomicts. PLOS ONE, 1 August, DOI: 10.1371/journal.pone.0041868CrossRef
Major, J. (1988). Endemism; a botanical perspective. In Analytical biogeography, ed. Myers, A. A. & Giller, P. S., pp. 117–46. London: Chapman & Hall.Google Scholar
Malcolm, J. R.et al. (2002). Estimated migration rates under scenarios of global climate change. Journal of Biogeography, 29, 835–49.CrossRefGoogle Scholar
Malhi, Y.et al. (2008). Climate change, deforestation, and the fate of the Amazon. Science, 319, 169–72.CrossRefGoogle ScholarPubMed
Malinska, H.et al. (2011). Ribosomal RNA genes evolution in Tragopogon: a story of New and Old World allotetraploids and the synthetic lines. Taxon, 60, 348–54.Google Scholar
Mallet, J. (2005). Hybridization as an invasion of the genome. Trends in Ecology and Evolution, 20, 229–37.CrossRefGoogle ScholarPubMed
Mallet, J. & Willmott, K. (2003). Taxonomy: renaissance or Tower of Babel?Trends in Ecology & Evolution, 18, 57–9.CrossRefGoogle Scholar
Mann, C. C. (1991). Extinction: are ecologists crying wolf?Science, 235, 736–8.Google Scholar
Mansion, G.et al. (2009). Origin of Mediterranean endemics in the Boraginales: integrative evidence from molecular dating and ancestral area reconstruction. Journal of Biogeography, 36, 1282–96.CrossRefGoogle Scholar
Manton, I. (1950). Problems of cytology and evolution in the Pteridophyta. London & New York: Cambridge University Press.CrossRefGoogle Scholar
Mao, Q. & Huff, D. R. (2012). The evolutionary origin of Poa annua L. Crop Science, 52, 1910–22.
Marble, B. K. (2004). Polyploidy and self-compatibility: is there an association?New Phytologist, 162, 803–11.Google Scholar
Marchant, C. J. (1963). Corrected chromosome numbers for Spartina ×townsendii and its parent species. Nature, 199, 299.CrossRefGoogle Scholar
Marchant, C. J. (1967). Evolution in Spartina (Gramineae). 1. The history and morphology of the genus in Britain. Journal of the Linnean Society (Botany), 60, 1–24.Google Scholar
Marchant, C. J. (1968). Evolution in Spartina (Gramineae). 2. Chromosomes, basic relationships and the problem of S. ×townsendii agg. Journal of the Linnean Society (Botany), 60, 381–409.Google Scholar
Marchi, P., Illuminati, O., Macioce, A., Capineri, R. & D'Amato, G. (1983). Genome evolution and polyploidy in Leucathemum vulgare Lam. aggr. (Compositae). Karyotype analysis and DNA microdensitometry. Caryologia, 36, 1–18.CrossRefGoogle Scholar
Margulis, L. (1971). Symbiosis and evolution. Scientific American, 225, 48–57.CrossRefGoogle ScholarPubMed
Marion, G. M.et al. (1997). Open-top designs for manipulating field temperature in high-latitude ecosystems. Global Change Biology, 3 (Suppl. 1), 20–32.CrossRefGoogle Scholar
Marks, G. E. (1966). The origin and significance of intraspecific polyploidy: experimental evidence from Solanum chaeoense. Evolution, 20, 552–7.CrossRefGoogle Scholar
Marsden-Jones, E. (1930). The genetics of Geum intermedium Willd. haud Ehrh. and its back-crosses. Journal of Genetics, 23, 377–95.CrossRefGoogle Scholar
Marsden-Jones, E. M. & Turrill, W. B. (1945). Report of the transplant experiments of the British Ecological Society. Journal of Ecology, 33, 59–81. [See also earlier reports in the Journal of Ecology: 18, 352; 21, 268; 23, 443; 25, 189; 26, 359 & 380.]CrossRefGoogle Scholar
Marshall, D. R. & Brown, A. H. D. (1981). The evolution of apomixis. Heredity, 47, 1–15.CrossRefGoogle Scholar
Marshall, D. R. & Weiss, P. W. (1982). Isozyme variation within and among Australian populations of Emex spinosa (L.). Campd. Australian Journal of Biological Sciences, 35, 327–32.Google Scholar
Martin, N. H., Bouck, C. & Arnold, M. L. (2006). Detecting adaptive trait introgression between Iris fulva and I.brevicaulis in highly selective field conditions. Genetics, 172, 2481–9.Google ScholarPubMed
Martin, S. L. & Husband, B. C. (2009). Influence of phylogeny and ploidy on species ranges of North American angiosperms. Journal of Ecology, 97, 913–22.CrossRefGoogle Scholar
Maschinski, J. (2001). Extinction risk of Ipomopsis sancti-spiritus in the Holy Ghost Canyon with and without management intervention. In Southwestern rare and endangered plants, ed. Maschinski, J. and Holter, L., pp. 206–12. Fort Collins, CO: US Department of Agriculture, Forest Service, Rocky Mountain Research Station.Google Scholar
Maschinski, J. & Haskins, K. E. (eds.) (2012). Plant reintroduction in a changing climate: promises and perils. Washington DC: Island Press.CrossRefGoogle Scholar
Maschinski, J., Baggs, J. E. & Sacchi, C. F. (2004). Seedling recruitment and survival of an endangered limestone endemic in its natural habitat and experimental reintroduction sites. American Journal of Botany, 91, 689–98.CrossRefGoogle ScholarPubMed
Maschinski, J.et al. (2011). Sinking ships: conservation options for endemic taxa threatened by sea level rise. Climate Change, 107, 147–67.CrossRefGoogle Scholar
Mason, B. J. (1992). Acid rain: its causes and its effects on island waters. Oxford: Clarendon Press.Google Scholar
Mason-Gamer, R. J. (2004). Reticulate evolution, introgression, and intertribal gene capture in an allohexaploid grass. Systematic Biology, 53, 25–37.CrossRefGoogle Scholar
Massart, J. (1902). L'accomodation individuelle chez le Polygonum amphibium. Bulletin de Jardin Botanique de l'Etat à Bruxelles, 1, 73–95.Google Scholar
Masterson, J. (1981). Stomatal size in fossil plants: evidence for polyploidy in majority of angiosperms. Science, 264, 421–4.Google Scholar
Mather, K. (1943). Polygenic inheritance and natural selection. Biological Reviews, 18, 32–64.CrossRefGoogle Scholar
Mather, K. (1966). Breeding systems and response to selection. In Reproductive biology and taxonomy of vascular plants, ed. Hawkes, J. G., pp. 13–19. Conference report of Botanical Society of the British Isles. Oxford: Pergamon Press.Google Scholar
Matsushita, S. C.et al. (2012). Allopolyploidization lays the foundation for evolution of distinct populations: evidence from analysis of synthetic Arabidopsis allohexaploids. Genetics, DOI:10.1534/genetics.112.139295.CrossRef
Matthew, P. (1831). Ideas on evolution, published in an Appendix to On naval timber and arboriculture. London.Google Scholar
Matthews, R. E. F. (1991). Plant virology, edn. New York: Academic Press.Google Scholar
Matz, M. V. & Nielsen, R. (2005). A likelihood ratio test for species membership based on DNA sequence data. Philosophical Transactions of the Royal Society of London, B, 359, 1969–74.Google Scholar
Maunder, M. & Ramsay, M. (1994). The reintroduction of plants into the wild: an integrated approach to the conservation of native plants. In The common ground of wild and cultivated plants, ed. Perry, A. R. & Ellis, R. G., pp. 81–8. Cardiff: National Museum of Wales.Google Scholar
Maunder, M.et al. (2004). Hybridization in ex situ plant collections; conservation concerns, liabilities and opportunities. In Ex situ conservation: supporting species survival in the wild, ed. Guerrant, E. O., Havens, K. & Maunder, M., pp. 325–64. Washington DC, Covelo, CA, & London: Island Press.Google Scholar
Maunder, M., Higgens, S. & Culham, A. (2001). The effectiveness of botanic garden collections in supporting plant conservation: a European case history. Biodiversity and Conservation, 10, 383–401.CrossRefGoogle Scholar
Maurice, S.et al. (1993). The evolution of gender in hermaphrodites of gynodioecious populations with nucleo-cytoplasmic male-sterility. Proceedings of the Royal Society of London, B. 251, 253–61.CrossRefGoogle Scholar
Maxwell, B. D. & Mortimer, A. M. (1994). Selection for herbicide resistance. In Herbicide resistance in plants: biology and biochemistry, ed. Powles, S. B. & Holtum, J. A. M., pp. 1–26. Boca Raton, FL, & London: Lewis Publishers.Google Scholar
May, R. M. (2011). Why should we be concerned about loss of biodiversity?Comptes Rendus Biologies, 334, 346–50.CrossRefGoogle ScholarPubMed
Mayer, M. S., Soltis, P. S. & Soltis, D. E. (1994). The evolution of the Streptanthus glandulosus complex (Cruciferae) – genetic divergence and gene flow in serpentine endemics. American Journal of Botany, 81, 1288–99.CrossRefGoogle Scholar
Maynard Smith, J. (1966). Sympatric speciation. American Naturalist, 100, 637–50.Google Scholar
Maynard Smith, J. (1978). The evolution of sex. Cambridge: Cambridge University Press.Google Scholar
Maynard Smith, J. (1983). The genetics of stasis and punctuation. Annual Review of Genetics, 17, 11–25.Google Scholar
Maynard Smith, J. (1989). Evolutionary genetics. New York: Oxford University Press.Google Scholar
Maynard Smith, J. & Haigh, J. (1974). The hitch-hiking effect of a favourable gene. Genetic Research, 23, 23–35.Google Scholar
Maynard-Smith, J. & Szathmáry, E. (1995). The major transitions in evolution. Oxford: Oxford University Press.Google Scholar
Mayr, E. (1942). Systematics and the origin of species. New York: Columbia University Press.Google Scholar
Mayr, E. (1963). Animal species and evolution. London: Oxford University Press.CrossRefGoogle Scholar
Mayr, E. (1969). Principles of systematic zoology. New York: McGraw-Hill.Google Scholar
Mayr, E. (1982). The growth of biological thought: diversity, evolution and inheritance. Cambridge, MA: Harvard University Press.Google Scholar
Mayr, E. & Provine, W. B. (eds.) (1980). The evolutionary synthesis: perspectives on the unification of biology. Cambridge, MA: Harvard University Press.CrossRefGoogle Scholar
McBreen, K. & Lockhart, P. J. (2006). Reconstructing reticulate evolutionary histories of plants. Trends in Plant Science, 11, 398–404.CrossRefGoogle Scholar
McCarthy, J. J.et al. (2001) Climate change 2001: impacts, adaptation, and vulnerability. Contribution of Working Group II to the Third Assessment Report of the Intergovernmental Panel on Climate Change. New York: Cambridge University Press.Google Scholar
McCauley, D. E. (1994). Contrasting the distribution of chloroplast DNA and allozyme polymorphism among local populations of Silene alba: implications for studies of gene flow in plants. Proceedings of the National Academy of Sciences, USA, 91, 8127–31.CrossRefGoogle ScholarPubMed
McCauley, D. E. & Bailey, M. F. (2009). Recent advances in the study of gynodioecy: the interface of theory and empiricism. Annals of Botany, 104, 611–20.CrossRefGoogle Scholar
McCauley, D. J. (2006). Selling out to nature. Nature, 443, 27–8.CrossRefGoogle Scholar
McClintock, B. (1984). The significance of responses of the genome to challenge. Science, 226, 792–801.CrossRefGoogle ScholarPubMed
McCubbin, A. G. (2008). Heteromorphic self-incompatibility in Primula: twenty-first century tools promise to unravel a classic nineteenth century model system. In Self-incompatibility in flowering plants, evolution, diversity and mechanisms, ed. Franklin-Tong, V., pp. 286–308. Berlin & Heidelberg: Springer.Google Scholar
McDonald, J. H. & Kreitman, M. (1991). Adaptive protein evolution at the Adh locus in Drosophila. Nature, 351, 652–4.CrossRefGoogle ScholarPubMed
McDonald, R. I. & Boucher, T. M. (2011). Global development and the future of the protected area strategy. Biological Conservation, 144, 383–92.CrossRefGoogle Scholar
McElwain, J. C. & Punyasena, S. W. (2007). Mass extinction events and the plant fossil record. Trends in Ecology and Evolution, 22, 548–57.CrossRefGoogle ScholarPubMed
McFadden, E. S. & Sears, E. R. (1946). The origin of Triticum spelta and its free-threshing hexaploid relatives. Hybrids of synthetic T. spelta with cultivated hexaploids. Journal of Heredity, 37, 81–9, 107–16.Google Scholar
McFadden, G. I. & van Dooren, G. G. (2004). Evolution: red algal genome affirms a common origin of all plastids. Current Biology, 14, R514–R516.CrossRefGoogle ScholarPubMed
Macfarlane, A. (1916). Lectures on ten British mathematicians of the nineteenth century. New York: Wiley; London: Chapman & Hall.Google Scholar
McGrath, C. L. & Lynch, M. (2012). Evolutionary significance of whole-genome duplication. In Polyploidy and genome evolution, ed. Soltis, P. S. & Soltis, D. E. pp.1–20. Heidelberg, New York, Dordrecht & London: Springer.Google Scholar
McInerney, J. O. & Wilkinson, M. (2005). New methods ring changes for the tree of life. Trends in Ecology & Evolution, 20, 105–7.CrossRefGoogle ScholarPubMed
McKay, J. K.et al. (2005) ‘How local is local?’ – A review of practical and conceptual issues in the genetics of restoration. Restoration Ecology, 13, 432–40.CrossRefGoogle Scholar
McKenney, D. W.et al. (2007). Potential impacts of climate change on the distribution of North American trees. Bioscience, 57, 939–48.CrossRefGoogle Scholar
McKenney, D. W.et al. (2011). Revisiting projected shifts in the climate envelopes of North American trees using updated general circulation models. Global Change Biology, 17, 2720–30.CrossRefGoogle Scholar
McLachlan, J. S., Hellmann, J. J.& Schwartz, M. W. (2007). A framework for debate of assisted migration in an era of climate change. Conservation Biology, 21, 297–302.CrossRefGoogle Scholar
McLean, R. C. & Ivimey-Cook, W. R. (1956). Textbook of theoretical botany. London, New York & Toronto: Longmans.Google Scholar
McLeish, J. & Snoad, B. (1962). Looking at chromosomes. London & New York: Macmillan. [edn, 1972.]Google Scholar
McMillan, C. (1970). Photoperiod in Xanthium populations from Texas and Mexico. American Journal of Botany, 57, 881–8.CrossRefGoogle Scholar
McMillan, C. (1971). Photoperiod evidence in the introduction of Xanthium (Cocklebur) to Australia. Science, 171, 1029–31.CrossRefGoogle Scholar
McMullen, C. K. (1987). Breeding systems of selected Galápagos Islands angiosperms. American Journal of Botany, 74, 1694–1705.CrossRefGoogle Scholar
McNaughton, I. H. & Harper, J. L. (1960). The comparative biology of closely related species living in the same area. New Phytologist, 59, 27–41.Google Scholar
McNeill, C. I. & Jain, S. K. (1983). Genetic differentiation studies and phylogenetic inference in the plant genus Limnanthes (Section Inflexae). Theoretical and Applied Genetics, 66, 257–69.Google Scholar
McNeill, J. (1976). The taxonomy and evolution of weeds. Weed Research, 16, 399–413.CrossRefGoogle Scholar
McNeill, J.et al. (eds.) (2006). International Code of Botanical Nomenclature (Vienna Code): adopted by the Seventeenth International Botanical Congress Vienna, Austria, July 2005. Regnum Vegetabile, 146, 1–568.Google Scholar
McNeill, J. R. (2000). Something new under the sun. London & New York: Allen Lane, The Penguin Press.Google Scholar
McNeilly, T. (1968). Evolution in closely adjacent plant populations. III. Agrostis tenuis on a small copper mine. Heredity, 23, 99–108.CrossRefGoogle Scholar
McNeilly, T. & Antonovics, J. (1968). Evolution in closely adjacent plant populations. IV. Barriers to gene flow. Heredity, 23, 205–18.CrossRefGoogle Scholar
McPherson, J. D. (2009). Next-generation gap. Nature Methods, 6, S2–S5.CrossRefGoogle ScholarPubMed
Meacher, T. R. (1986). Analysis of paternity within a natural population of Chamaelirium luteum. I. Identification of most-likely parents. American Naturalist, 127, 199–215.Google Scholar
Meacher, T. R. & Thompson, E. (1987). Analysis of parentage for naturally established seedlings of Chamaelirium luteum (Liliacae). Ecology, 68, 803–12.Google Scholar
Mead, R. (1988). The design of experiments. Cambridge: Cambridge University Press.Google Scholar
Médail, F. & Diadema, K. (2009). Glacial refugia influence plant diversity patterns in the Mediterranean Basin. Journal of Biogeography, 36, 1333–45.CrossRefGoogle Scholar
Medawar, P. (1984). Pluto's Republic. Oxford, New York: Oxford University Press.Google Scholar
Medawar, P. (1991). The threat and the glory. Oxford: Oxford University Press.Google Scholar
Meffe, G. K. & Carroll, C. R. (1994). Principles of conservation biology. Sunderland, MA: Sinauer.Google Scholar
Meharg, A. A. (2003). The mechanistic basis of interactions between mycorrhizal associations and toxic metal cations. Mycological Research, 107, 1253–65.CrossRefGoogle ScholarPubMed
Meharg, P. A., Cumbes, Q. J. & Macnair, M. R. (1993). Pre-adaptation of Yorkshire Fog Holcus lanatus L. (Poaceae) to arsenate tolerance. Evolution, 47, 313–16.CrossRefGoogle Scholar
Meirmans, P. G., Den Nijs, H. (J.) C. M. & Van Tienderen, P. H. (2006). Male sterility in triploid dandelions: asexual females vs asexual hermaphrodites. Heredity, 96, 45–62.CrossRefGoogle ScholarPubMed
Memon, A. R. & Schröder, P. (2009). Implications of metal accumulation mechanisms to phytoremediation. Environmental Science and Pollution Research International, 16, 162–75.CrossRefGoogle ScholarPubMed
Mena-Ali, J. I. & Stephenson, A. G. (2007). Segregation analyses of partial self-incompatibility in self and cross progeny of Solanum carolinense reveal a leaky S-allele. Genetics, 177, 501–10.CrossRefGoogle ScholarPubMed
Menchari, Y.et al. (2006). Weed response to herbicides: regional-scale distribution of herbicide resistance alleles in the grass weed Alopecurus myosuroides. New Phytologist, 171, 861–74.CrossRefGoogle ScholarPubMed
Mendel, G. (1866). Versuche über Planzenhybriden. Verhandlungen des Naturforschenden Vereins in Brünn, 4, 344. [English translation in Bateson, W. (1909). Mendel's principles of heredity. London: Cambridge University Press; also Bennett, J. H. (ed.) (1965). Experiments in plant hybridisation. Edinburgh & London: Oliver & Boyd.]Google Scholar
Mendel, G. (1869). Uber einige aus künstlicher Befruchtung gewonnenen Hieracium-Bastarde. Verhandlungen des naturforschen den Vereines in Brünn, 8.Google Scholar
Méndez-Vigo, B.et al. (2011). The flowering repressor SVP underlies a novel Arabidopsis thaliana QTL interacting with the genetic background. PLoS, Biology, 31 January 2013, DOI:10.1371/journal.pgen.1003289.CrossRef
Menges, E. S. (1990). Population viability analysis for a rare plant. Conservation Biology, 5, 158–64.Google Scholar
Menges, E. S. (1991). The application of minimum viable population theory to plants. In Genetics and conservation of rare plants, ed. Falk, D. A. & Holsinger, K. E., pp. 45–61. Oxford: Oxford University Press.Google Scholar
Menges, E. S. (2008). Restoration demography and genetics of plants: when is translocation successful?Australian Journal of Botany, 56, 187–96.CrossRefGoogle Scholar
Mengoni, A.et al. (2001). Characterization of nickel-resistant bacteria isolated from serpentine soil. Environmental Microbiology, 3, 691–8.CrossRefGoogle ScholarPubMed
Mengoni, A.et al. (2003). Chloroplast genetic diversity and biogeography in the serpentine endemic Ni-hyperaccumulator Alyssum bertolonii. New Phytologist, 157, 349–56.CrossRefGoogle Scholar
Mercer, K. L. & Wainwright, J. D. (2008). Gene flow from transgenic maize to landraces in Mexico: an analysis. Agriculture, Ecosystems and Environment, 126, 109–15.CrossRefGoogle Scholar
Mereschkowsky, C. (1905). Über Natur und Ursprung der Chromatophoren im Planzenreich. Biologisches Zentralblatt, 25, 593–604.Google Scholar
Mergen, F. (1963). Ecotypic variation in Pinus strobus. Ecology, 44, 716–27.CrossRefGoogle Scholar
Merrell, D. J. (1962). Evolution and genetics: the modern theory of evolution. New York: Holt, Rinehart & Winston.Google Scholar
Merxmüller, H. (1970). Biosystematics: still alive? Provocation of biosystematics. Taxon, 19, 140–5.CrossRefGoogle Scholar
Mes, T. H. M. (1998). Character compatibility of molecular markers to distinguish asexual and sexual reproduction. Molecular Ecology, 7, 1719–27.CrossRefGoogle Scholar
Mes, T. H. M., van Brederode, J. & Hart, H. (1996). Origin of the Woody Macaronesian Sempervivoideae and the phylogenetic position of the East African species of Aeonium. Botanica Acta, 109, 477–91.CrossRefGoogle Scholar
Meyer, P. (2007). Gene silencing. In Handbook of plant science, ed. Roberts, K., pp. 627–34. Chichester: Wiley.Google Scholar
Michaelis, P. (1954). Cytoplasmic inheritance in Epilobium and its theoretical significance. Advances in Genetics, 6, 288–402.Google ScholarPubMed
Millener, L. H. (1961). Day length as related to vegetative development in Ulex europaeus. 1. The experimental approach. New Phytologist, 60, 339–54.CrossRefGoogle Scholar
Miller, T. E. (1987). Systematics and evolution. In Wheat breeding; its scientific basis, ed. Lupton, F. G. H., pp.1–30. London: Chapman & Hall.Google Scholar
Mills, A. D. (1993). English place names. Oxford: Oxford University Press.Google Scholar
Milne, R. I. (2006). Northern hemisphere plant disjunctions: a window on Tertiary land bridges and climate change?Annals of Botany, 98, 465–72.Google Scholar
Milton, S. J.et al. (1999) A protocol for plant conservation by translocation in threatened lowland fynbos. Conservation Biology, 13, 735–43.CrossRefGoogle Scholar
Minder, A. M, Rothenbuehler, C. & Widmer, A. (2007). Genetic structure of hybrid zones between Silene latifolia and Silene dioica (Caryophyllaceae): evidence for introgressive hybridization. Molecular Ecology, 16, 2504–16.CrossRefGoogle ScholarPubMed
Minguzzi, C. & Vergnano, O. (1948). 11 contenuto di niche1 nelle ceneri di Alyssum bertolonii. Atti della Società Toscana de Scienze Naturali di Pisa, 55, 49–74.Google Scholar
Mishler, B. D. (2010). Species are not uniquely real biological entities. In Contemporary debates in philosophy of biology, ed. Ayala, F. & Arp, R., pp. 110–22. Singapore: Wiley-Blackwell.Google Scholar
Mitchell, R. S. (1968). Variation in the Polygonum amphibium complex and its taxonomic significance. University of California Publications in Botany, 45, 1–54.Google Scholar
Mitchell-Olds, T. & Schmidt, J. (2006). Genetic mechanisms and evolutionary significance of natural variation in Arabidopsis. Nature, 41, 947–52.Google Scholar
Mivart, St. G. (1871). The genesis of species, edn. London: Macmillan.CrossRefGoogle Scholar
Modliszewski, J. L. & Willis, J. H. (2012). Allotetraploid Mimulus sookensis are highly interfertile despite independent origins. Molecular Ecology, 21, 5280–98.CrossRefGoogle ScholarPubMed
Mogie, M. (1992). The evolution of asexual reproduction in plants. London: Chapman & Hall.Google Scholar
Mølgaard, P. (1976). Plantago major ssp. major and ssp. pleiosperma. Morphology, biology and ecology in Denmark. Botanik Tidsskrift, 71, 31–56.Google Scholar
Montesinos-Navarro, A.et al. (2011). Arabidopsis thaliana populations show clinal variation in a climatic gradient associated with altitude. New Phytologist, 189, 282–94.CrossRefGoogle Scholar
Mooney, H. A. & Billings, W. D. (1961). Comparative physiological ecology of Arctic and Alpine populations of Oxyria digyna. Ecological Monographs, 31, 1–29.CrossRefGoogle Scholar
Mooney, H. A. & Cleland, E. E. (2001). The evolutionary impact of invasive species. Proceedings of the National Academy of Sciences, USA, 98, 5446–51.CrossRefGoogle ScholarPubMed
Moore, D. M. (1959). Population studies on Viola lactea Sm. and its wild hybrids. Evolution, 13, 318–32.CrossRefGoogle Scholar
Moore, D. M. (1976). Plant cytogenetics. London: Chapman & Hall; New York: Wiley & Sons.Google Scholar
Moore, D. M. (1982). Flora Europaea check-list and chromosome index. London: Cambridge University Press.CrossRefGoogle Scholar
Moore, D. M. & Harvey, M. J. (1961). Cytogenetic relationships of Viola lactea Sm. and other West European arosulate violets. New Phytologist, 60, 85–95.CrossRefGoogle Scholar
Moore, R. J. & Mulligan, G. A. (1956). Natural hybridization between Carduus acanthoides and Carduus nutans in Ontario. Canadian Journal of Botany, 34, 71–85.CrossRefGoogle Scholar
Moore, R. J. & Mulligan, G. A. (1964). Further studies on natural selection among hybrids of Carduus acanthoides and Carduus nutans. Canadian Journal of Botany, 42, 1605–13.CrossRefGoogle Scholar
Morisset, P. & Boutin, C. (1984) The biosystematic importance of phenotypic plasticity. In Plant systematics, ed. Grant, W. F., pp. 293–306. Toronto: Academic Press.Google Scholar
Moritz, D. M. L. & Kadereit, J. W. (2001).The genetics of evolutionary change in Senecio vulgaris L.: a QTL mapping approach. Plant Biology, 3, 544–52.CrossRefGoogle Scholar
Moro, C., Rollo, A. & Tittensor, D. P. (2013). Comment on ‘Can we name Earth's species before they go extinct?Science, 341, 237.CrossRefGoogle Scholar
Morris, M. G. & Perring, F. H. (1974). The British oak: its history and natural history. Faringdon: Classey.Google Scholar
Morris, S. (2012). National Trust secret £700,000 complex keeps rare plants safe. The Guardian, 21 June.
Morrone, J. J. (2009). Evolutionary biogeography: an integrative approach with case studies. New York: Columbia University Press.Google Scholar
Morton, J. K. (1966). The role of polyploidy in the evolution of a tropical flora. In Chromosomes today, vol. 1, ed. Darlington, C. D. & Lewis, K. R., pp.73–6, Edinburgh: Oliver & Boyd.Google Scholar
Motley, T. J., Zerega, N. & Cross, H. (2006). Darwin's harvest. New York: Columbia University Press.CrossRefGoogle Scholar
Mueller, J. M. & Hellman, J. J. (2008). An assessment of invasion risk from assisted migration. Conservation Biology, 22, 562–7.CrossRefGoogle ScholarPubMed
Muguia-Rosas, M. A.et al. (2011). Meta-analysis of phenotypic selection on flowering phenology suggests that early flowering plants are favoured. Ecology Letters, 14, 511–21.Google Scholar
Muller, G. (1977). Cross-fertilization in a conifer stand inferred from enzyme gene-markers in seeds. Silvae Genetica, 26, 223–6.Google Scholar
Müller, G. B. (2008). Evo-devo as a discipline. In Evolving pathways: key themes in evolutionary developmental biology, ed. Minelli, A. and Fusco, G., pp. 3–29. Cambridge: Cambridge University Press.Google Scholar
Müntzing, A. (1930a). Uber Chromosomenvermehrung in Galeopsis-kreuzungen und ihre phylogenetische Bedeutung. Hereditas, 14, 153–72.Google Scholar
Müntzing, A. (1930b). Outlines to a genetic monograph of the genus Galeopsis with special reference to the nature and inheritance of partial sterility. Hereditas, 13, 185–341.Google Scholar
Müntzing, A. (1932). Cyto-genetic investigations on synthetic Galeopsis tetrahit. Hereditas, 16, 105–54.Google Scholar
Müntzing, A. (1936). The evolutionary significance of autopolyploidy. Hereditas, 21, 363–78.Google Scholar
Müntzing, A. (1961). Genetic research. Stockholm: L. T. Førlag.Google Scholar
Müntzing, A., Tedin, O. & Turesson, G. (1931). Field studies and experimental methods in taxonomy. Hereditas, 15, 1–12.Google Scholar
Murphy, C. E. & Lemerle, D. (2006). Continuous cropping systems and weed selection. Euphytica, 148, 61–73.CrossRefGoogle Scholar
Murphy, S. D.et al. (2006). Promotion of weed species diversity and reduction of weed seedbanks with conservation tillage and crop rotation. Weed Science, 54, 69–77.CrossRefGoogle Scholar
Myers, K. (1986). Introduced vertebrates in Australia, with emphasis on the mammals. In Ecology of biological invasions, ed. Groves, R. H. & Burdon, J. J., pp. 120–36. Cambridge: Cambridge University Press.Google Scholar
Myers, N. (1979). The sinking ark: a new look at the problem of disappearing species. Oxford: Pergamon Press.Google Scholar
Myers, N. & Knoll, A. H. (2001). The biotic crisis and the future of evolution. Proceedings of the National Academy of Sciences, USA, 98, 5389–91.CrossRefGoogle Scholar
Nägeli, C. (1865). Die Bastardbindung im Pflanzenreiche. Sitzungsbericht der Königlich-Bayerischen Akademie der Wissenschaften zu München Botanische Mitteilungen, 2, 159–87.Google Scholar
Naidoo, R.et al. (2008). Integrating economic costs into conservation planning. Trends in Ecology and Evolution, 21, 681–7.Google Scholar
Naiki, A. (2012). Heterostyly and the possibility of its breakdown by polyploidization. Plant Species Biology, 27, 3–29.CrossRefGoogle Scholar
Nanney, D. L. (1982). Genes and phenes in Tetrahymena. BioScience, 32, 783–8.CrossRefGoogle Scholar
Nannfeldt, J. A. (1937). The chromosome numbers of Poa, Sect. Ochlopoa A. and Gr. and their taxonomical significance. Botaniska Notiser, 1937, 238–57.Google Scholar
Nascimento, C. W. A. & Xing, B. (2006). Phytoextraction: a review on enhanced metal availability and plant accumulation. Scientia Agriccola, (Piracicaba, Brazil), 63, 299–311.Google Scholar
Nasrallah, M. E.et al. (2004). Natural variation in expression of self-incompatibility in Arabidopsis thaliana: implications for the evolution of selfing. Proceedings of the National Academy of Sciences, USA, 101, 16070–4.CrossRefGoogle ScholarPubMed
Nathan, R.et al. (2008). An emerging movement ecology paradigm. Proceedings of the National Academy of Sciences, USA, 105, 19050–1.Google ScholarPubMed
Naumova, T. N. (1993). Apomixis in Angiosperms; nucellar and integumental embryology. Boca Raton, FL: CRC Press.Google Scholar
Navarro, L. M. & Pereira, H. M. (2012). Rewilding abandoned landscape in Europe. Ecosystems, 15, 900–12.CrossRefGoogle Scholar
Navashin, M. (1926). Variabilität des Zellkerns bei Crepis-Arten in Bezug auf die Artbildung. Zeitschrift für Zellforschung und mikroskopische Anatomie, 4, 171–215.Google Scholar
Nei, M. (1972). Genetic distance between populations. American Naturalist, 106, 283–93.CrossRefGoogle Scholar
Nelson, A. P. (1967). Racial diversity in Californian Prunella vulgaris. New Phytologist, 66, 707–46.CrossRefGoogle Scholar
Nelson, G. (1979) From Candolle to Croizat: comments on the history of biogeography. Journal of the History of Biology, 11, 269–305.Google Scholar
Nelson-Jones, E. B., Briggs, D. & Smith, A. C. (2002). The origin of intermediate species of the genus Sorbus. Theoretical and Applied Genetics, 105, 953–63.Google ScholarPubMed
Neuffer, B. & Hurka, H. (1999). Colonization history and introduction dynamics of Capsella bursa-pastoris (Brassicaceae) in North America: isozymes and quantitative traits. Molecular Ecology, 8, 1667–81.CrossRefGoogle ScholarPubMed
Neuffer, B. & Linde, M. (1999). Capsella bursa-pastoris: colonisation and adaptation; a globetrotter conquers the world. In Plant evolution in man-made habitats, ed. van Raamsdonk, L. W. D. & Nijs, J. C. M. den, pp. 49–72. Amsterdam: Hugo de Vries Laboratory.Google Scholar
Neuhaus, D., Kühl, H., Kohl, J. G., Dörfel, P. & Börner, T. (1993). Investigations on the genetic diversity of Phragmites stands using genomic fingerprinting. Aquatic Botany, 45, 357–64.CrossRefGoogle Scholar
Nevill, P. G.et al. (2013). DNA barcoding for conservation, seed banking and ecological restoration of Acacia in the Midwest of Western Australia. Molecular Ecology, 13, 1033–42.Google ScholarPubMed
New, J. K. (1958). A population study of Spergula arvensis 1. Annals of Botany, New Series, 22, 457–77.
New, J. K. (1959). A population study of Spergula arvensis 2. Annals of Botany, New Series, 23, 23–33.
New, J. K. (1978). Change and stability of clines in Spergula arvensis L. (Corn Spurrey) after 20 years. Watsonia, 12(2), 137–43.Google Scholar
New, J. K. & Herriott, J. C. (1981). Moisture for germination as a factor affecting the distribution of the seedcoat morphs of Spergula arvensis L. Watsonia, 13(4), 323–4.
New, T. R (2006). Conservation biology in Australia: an introduction. Oxford: Oxford University Press.Google Scholar
Newmaster, S. G.et al. (2013). DNA barcoding detects contamination in North American herbal products. BMC Medicine, 11, 1741–7015.CrossRefGoogle ScholarPubMed
Newton, W. C. F. & Pellew, C. (1929). Primula kewensis and its derivatives. Journal of Genetics, 20, 405–66.CrossRefGoogle Scholar
Nichols, D. J. & Johnson, K. R. (2008). Plants and the K-T boundary. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Nicholson, M. (1970). The environmental revolution: a guide for the new masters of the world. London: Hodder & Stoughton.Google Scholar
Nickrent, D. L.et al. (1998). Molecular phylogenetic and evolutionary studies of parasitic plants. In Plant molecular systematics II, ed. Soltis, D. E., Soltis, P. S. & Doyle, J. J., pp.211–41. Boston: Kluwer.Google Scholar
Nicolia, A.et al. (2013). An overview of the last 10 years of genetically engineered crop safety research. Critical Reviews in Biotechnology, 34, 77–88.Google ScholarPubMed
Niemela, P. & Tuomi, J. (1987). Does the leaf morphology of some plants mimic caterpillar damage?Oikos, 50, 256–7.CrossRefGoogle Scholar
Nilsson-Ehle, E. (1909). Kreuzungsuntersuchungen an Hafer und Weizen. Acta Universitatis Lundensis, Ser. 2, 5(2), 1–122.Google Scholar
Njoku, E. (1956). Studies on the morphogenesis of leaves. II. The effect of light intensity on leaf shape in Ipomoea caerulea. New Phytologist, 55, 91–110.Google Scholar
Noble, I. R. & Dirzo, R. (1997). Forests as human-dominated ecosystems. Science, 277, 522–5.CrossRefGoogle Scholar
Nogales, M.et al. (2012). Evidence for overlooked mechanisms of long-distance seed dispersal to and between oceanic islands. New Phytologist, 194, 313–17.CrossRefGoogle ScholarPubMed
Nogler, G. A. (1984). Gametophytic apomixis. In Embryology of angiosperms, ed. Johri, B. M., pp.475–518. Berlin: Springer Verlag.Google Scholar
Nogués-Bravo, D. (2009). Predicting the past distribution of species climatic niches. Global Ecology and Biogeography, 18, 521–31.CrossRefGoogle Scholar
Noirot, M., Couvet, D. & Hamon, S. (1997). Main role of self-pollination rate on reproductive allocations in pseudogamous apomicts. Theoretical and Applied Genetics, 95, 479–83.CrossRefGoogle Scholar
Nordenskiøld, H. (1949). The somatic chromosomes of some Luzula species. Botaniska Notiser, 1949, 81–92.Google Scholar
Nordenskiøld, H. (1951). Cyto-taxonomical studies in the genus Luzula 1. Somatic chromosomes and chromosome numbers. Hereditas, 37, 325–55.Google Scholar
Nordenskiøld, H. (1956). Cyto-taxonomical studies in the genus Luzula. 2. Hybridization experiments in the campestris-multiflora complex. Hereditas, 42, 7–73.Google Scholar
Nordenskiøld, H. (1961). Tetrad analysis and the course of meiosis in three hybrids of Luzula campestris. Hereditas, 47, 203–38.Google Scholar
Noret, N.et al. (2005). Palatability of Thlaspi caerulescens for snails: influence of zinc on glucosinolates. New Phytologist, 165, 763–72.Google ScholarPubMed
North, H.et al. (2009). Arabidopsis seed secrets unravelled after a decade of genetic and omics-driven research. The Plant Journal, 61, 971–81.Google Scholar
Norton, B. J. (1983). Fisher's entrance into evolutionary science: the role of eugenics. In Dimensions of Darwinism, ed. Grene, M., pp. 19–30. Cambridge: Cambridge University Press.Google Scholar
Norton, D. A. (2009). Species invasions and the limits to restoration: learning from the New Zealand experience. Nature, 325, 569–73.Google ScholarPubMed
Nosil, P. & Feder, J. L. (2012). Genomic divergence during speciation: causes and consequences. Philosophical Transactions of the Royal Society, B, 367, 332–42.CrossRefGoogle ScholarPubMed
Nosil, P., Funk, D. J. & Ortiz-Barrientos, D. (2009). Divergent selection and heterogeneous genomic divergence. Molecular Ecology, 18, 375–402.CrossRefGoogle ScholarPubMed
Novak, J. & Mack, R. N. (1993). Genetic variation in Bromus tectorum (Poaceae): comparison between native and introduced populations. Heredity, 71, 167–76.CrossRefGoogle Scholar
Novak, S. J. & Mack, R. N. (1995). Allozyme diversity in the apomictic vine Bryonia alba (Cucurbitaceae): potential consequences of multiple introductions. American Journal of Botany, 82, 1153–62.CrossRefGoogle Scholar
Novak, S. J. & Mack, R. N. (2001). Tracing plant introduction and spread: genetic evidence from Bromus tectorum (cheatgrass). Bioscience, 51, 114–22.CrossRefGoogle Scholar
Novak, S. J., Soltis, D. E. & Soltis, P. S. (1991). Ownbey's Tragopogons – 40 years later. American Journal of Botany, 78, 1586–1600.CrossRefGoogle Scholar
Novak, S. J., Mack, R. N. & Soltis, P. S. (1993). Genetic variation in Bromus tectorum (Poaceae): introduction dynamics in North America. Canadian Journal of Botany, 71, 1441–8.CrossRefGoogle Scholar
Novy, A.Flory, S. L. & Hartman, J. M. (2013). Evidence for rapid evolution of phenology in an invasive grass. Journal of Evolutionary Biology, 26, 443–50.CrossRefGoogle Scholar
Núñez-Farfán, J. & Schlichting, C. D. (2005). Natural selection in Potentilla glandulosa revisited. Evolutionary Ecology Research, 7, 105–19.Google Scholar
Nyberg Berglund, A. B., Dalgren, S. & Westerbergh, A. (2003). Evidence for parallel evolution and site-specific selection of serpentine tolerance in Cerastium alpinum during the colonization of Scandinavia. New Phytologist, 161, 199–209.Google Scholar
Oelschlaeger, M. (1991). The idea of wilderness: from prehistory to the age of ecology. New Haven, CT: Yale University Press.Google Scholar
Ohno, S. (1970). Evolution by gene duplication. London: Allen & Unwin.CrossRefGoogle Scholar
Okaura, T. & Harada, K. (2002). Phylogeographical structure revealed by chloroplast DNA variation in Japanese beech (Fagus crenata Blume). Heredity, 88, 322–9.CrossRefGoogle Scholar
Olby, R. C. (1979). Mendel no mendelian?History of Science, 17, 53–72.CrossRefGoogle Scholar
Olby, R. C. (1985). Origins of Mendelism, edn. Chicago: University of Chicago Press.Google Scholar
Olby, R. C. (2009). Francis Crick: hunter of life's secrets. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory Press.Google Scholar
Olby, R. C. & Gautrey, P. (1968). Eleven references to Mendel before 1900. Annals of Science, 24, 7–20.CrossRefGoogle Scholar
Olivieri, L. (2001) The evolution of dispersal and other traits in metapopulation. In Integrating ecology and evolution in a spatial context, ed. Antonovics, J. & Silvertown, J., pp. 245–68. Oxford: Blackwell Science.Google Scholar
Olmstead, R. G. & Palmer, J. D. (1994). Chloroplast DNA systematics: a review of methods and data analysis. American Journal of Botany, 81, 1205–24.CrossRefGoogle Scholar
Olsen, K. M. & Ungerer, M. C. (2008). Freezing tolerance and cyanogenesis in White Clover (Trifolium repens L. Fabaceae). International Journal of Plant Sciences, 169, 1141–7.CrossRefGoogle Scholar
Olsen, K. M. & Wendel, J. F. (2013). A bountiful harvest: genomic insights into crop domestication phenotypes. Annual Review of Plant Biology, 64, 47–70.CrossRefGoogle ScholarPubMed
Olsen, K. M., Sutherland, B. L. & Small, L. L. (2007). Molecular evolution of the Li/li chemical defence polymorphism in white clover (Trifolium repens L.). Molecular Ecology, 16, 4180–93.CrossRefGoogle Scholar
Olsen, K. M., Hsu, S.-C. & Small, L. L. (2008). Evidence on the molecular basis of the Ac/ac adaptive cyanogenesis polymorphism in white clover (Trifolium repens L.). Genetics, 179, 517–26.CrossRefGoogle Scholar
Olsen, K. M., Kooyers, N. J. & Small, L. L. (2013). Recurrent gene deletions and the evolution of adaptive cyanogenesis polymorphisms in white clover (Trifolium repens L.). Molecular Ecology, 22, 724–38.CrossRefGoogle Scholar
Olsen, K. M., Kooyers, N. J. & Small, L. L. (2014). Adaptive gains through repeated gene loss: parallel evolution of cyanogenesis polymorphisms in the genus Trifolium (Fabaceae). Philosophical Transactions of the Royal Society, B, 369, 20130347, DOI:10.1098/rstb.2013.0347.CrossRefGoogle Scholar
Ooi, M. K. J., Auld, T. D. & Denham, A. J. (2012). Projected soil temperature increase and seed dormancy response along an altitudinal gradient: implications for seed bank persistence under climate change. Plant and Soil, 353, 289–303.CrossRefGoogle Scholar
Oostermeijer, J. G. B., Den Nijs, J. C. M., Raijmann, L. E. L. & Menken, S. B. J. (1992). Population biology and management of the Marsh Gentian (Gentiana pneumonanthe L.), a rare species in the Netherlands. Botanical Journal of the Linnean Society, 108, 117–30.CrossRefGoogle Scholar
Orel, V. (1984). Mendel. Past Masters Series. Oxford: Oxford University Press.Google Scholar
Orel, V. (1996). Gregor Mendel: the first geneticist. New York: Oxford University Press.Google Scholar
Orel, V. & Matalová, A. (1983). Gregor Mendel and the foundation of genetics. Brno: Mendelianum of the Moravian Museum.Google Scholar
Oreskes, N. (2004). The scientific consensus on climate change. Science, 306, 1686.CrossRefGoogle ScholarPubMed
Ornduff, R. (1966). The origin of dioecism from heterostyly in Nymphoides (Menyanthaceae). Evolution, 20, 309–14.CrossRefGoogle Scholar
Ornduff, R. (1969). Reproductive biology in relation to systematics. Taxon, 18, 121–33.CrossRefGoogle Scholar
Ornduff, R. (1970). Pathways and patterns of evolution – a discussion. Taxon, 19, 202–4.CrossRefGoogle Scholar
Ortiz-Garcia, S.et al. (2005). Absence of detectable transgenes in local landraces of maize in Oaxaca, Mexico (2003–2004). Proceedings of the National Academy of Sciences, USA, 102, 12338–43.CrossRefGoogle Scholar
Osborn, H. F. (1894). From the Greeks to Darwin: an outline of the development of the evolution idea. London & New York: Macmillan.Google Scholar
Osevik, K. L.et al. (2012). Parallel ecological speciation in plants?Hindawi Publishing Corporation International Journal of Ecology, 2012, Article ID 939862, DOI:10.1155/2012/939862.Google Scholar
Ossowski, S.et al. (2010). The rate and molecular spectrum of spontaneous mutations in Arabidopsis thaliana. Science, 327, 92–4.CrossRefGoogle ScholarPubMed
Oswald, P. H. & Preston, C. D. (eds.) (2011). John Ray's Cambridge catalogue (1660) translated by Oswald, P. H., & Preston, C. D.. London: Ray Society.Google Scholar
Ouborg, N. J. & Eriksson, O. (2004). Towards a metapopulation concept for plants. In Ecology, genetics, and evolution of metapopulations, ed. Hanski, I. & Gaggiotti, O. E., pp. 447–69. Amsterdam: Elsevier.Google Scholar
Ouborg, N. J. & van Treuren, R. (1995). Variation in fitness-related characters among small and large populations of Salvia pratensis. Journal of Ecology, 83, 369–80.CrossRefGoogle Scholar
Overpeck, J. T., & Weiss, J. L. (2009) Projections of future sea level becoming more dire. Proceedings of the National Academy of Sciences, USA, 106, 21461–2.CrossRefGoogle ScholarPubMed
Ownbey, M. (1950). Natural hybridisation and amphidiploidy in the genus Tragopogon. American Journal of Botany, 37, 487–99.CrossRefGoogle Scholar
Ownbey, M. & McCollum, G. D. (1953). Cytoplasmic inheritance and reciprocal amphiploidy in Tragopogon. American Journal of Botany, 40, 788–96.CrossRefGoogle Scholar
Ownbey, M. & McCollum, G. D. (1954). The chromosome of Tragopogon. Rhodora, 56, 7–21.Google Scholar
Ozias-Atkins, P. (2006). Apomixis: developmental characteristics and genetics. Critical Review of Plant Sciences, 25, 199–214.Google Scholar
Pagel, M. (2002). Phylogenetic inference: methods. In Oxford encyclopaedia of evolution, ed. Pagel, M., pp. 895–904. Oxford: Oxford University Press.Google Scholar
Paine, D. P. & Kiser, J. D. (2012). Aerial photography and image interpretation, edn. Hoboken: Wiley.CrossRefGoogle Scholar
Palmer, J. D. (1988). Intraspecific variation and multicircularity in Brassica mitochondrial DNAs. Genetics, 118, 341–51.Google ScholarPubMed
Palmer, M. A. & Filoso, S. (2009). Restoration of ecosystems services for environmental markets. Science, 325, 575–6.CrossRefGoogle ScholarPubMed
Palumbi, S. R. (2001). The evolution explosion: how humans cause rapid evolutionary change. New York & London: Norton.Google Scholar
Pankhurst, R. J. (1991). Practical taxonomic computing. Cambridge: Cambridge University Press.Google Scholar
Pannell, J. (2008). Widespread functional androdioecy in Mercurialis annua L. (Euphorbiaceae). Biological Journal of the Linnean Society, 61, 95–116.Google Scholar
Parisod, C. (2008). Postglacial recolonisation of plants in the western Alps of Switzerland. Botanica Helvetica, 118, 1–12.CrossRefGoogle Scholar
Parisod, C. & Bescard, G. (2007). Glacial in situ survival in the Western Alps and polytopic autopolyploidy in Biscutella laevigata L. (Brassicaceae). Molecular Ecology, 16, 2755–67.CrossRefGoogle Scholar
Parisod, C., Trippi, C. & Galland, N. (2005). Genetic variability and founder effect in the Pitcher Plant Sarracenia purpurea (Sarraceniaceae) in populations introduced into Switzerland: from inbreeding to invasion. Annals of Botany, 95, 277–86.CrossRefGoogle ScholarPubMed
Parisod, C.et al. (2009). Rapid structural and epigenetic reorganization near transposable elements in hybrid and allopolyploid genomes in Spartina. New Phytologist, 184, 1003–15.CrossRefGoogle ScholarPubMed
Parisod, C., Holderegger, R. & Brochmann, C. (2010). Evolutionary consequences of autopolyploidy. New Phytologist, 186, 5–17.CrossRefGoogle ScholarPubMed
Parker, D. M. (1982). The conservation, by restocking, of Saxifraga cespitosa in North Wales. Watsonia, 14, 104–5.Google Scholar
Parker, I. M., Rodriguez, J. & Loik, M. E. (2002). An evolutionary approach to understanding the biology of invasions: local adaptation and general-purpose genotypes in the weed Verbascum thapsus. Conservation Biology, 17, 59–72.Google Scholar
Parker, R. E. (1973). Introductory statistics for biology. London: Arnold.Google Scholar
Parks, J. C. & Werth, C. R. (1993). A study of spatial features of clones in a population of bracken fern, Pteridium aquilinum (Dennstaeditiaceae). American Journal of Botany, 80, 537–44.CrossRefGoogle Scholar
Parmentier, I.et al. (2013). How effective are DNA barcodes in the identification of African rainforest trees?PLOS ONE, 8, 1–10.CrossRefGoogle ScholarPubMed
Parmesan, C. & Yohe, G. (2003). A globally coherent fingerprint of climate change impacts across natural systems. Nature, 421, 37–42.CrossRefGoogle ScholarPubMed
Parmesan, C.et al. (1999). Poleward shifts in geographical ranges of butterfly species associated with regional warming. Nature, 399, 579–83.CrossRefGoogle Scholar
Parokonny, A. S., Kenton, A., Gleba, Y. Y. & Bennett, M. D. (1994). The fate of recombinant chromosomes and genome interaction in Nicotiana hybrids and their sexual progeny. Theoretical and Applied Genetics, 89, 488–97.CrossRefGoogle ScholarPubMed
Parr, C. Set al. (2011). Evolutionary informatics: unifying knowledge about the diversity of Life. Trends in Ecology and Evolution, 27, 94–103.Google ScholarPubMed
Parsons, J. J. (1970). The Africanization of the New World tropical grasslands. Tübinger Geographische Studien, 34, 141–53.Google Scholar
Parsons, P. A. (1959). Some problems in inbreeding and random mating in tetrasomics. Agronomy Journal, 51, 465–7.CrossRefGoogle Scholar
Paterniani, E. (1969). Selection for reproductive isolation between two populations of Maize, Zea mays L. Evolution, 23, 534–47.
Paterson, A. H.et al. (2006). Many gene and domain families have convergent fates following independent whole-genome duplication events in Arabidopsis, Oryza, Saccharomyces and Tetraodon. Trends in Genetics, 22, 597–602.CrossRefGoogle ScholarPubMed
Patterson, C. (ed.) (1987). Molecules and morphology in evolution: conflict or compromise?Cambridge: Cambridge University Press.Google Scholar
Paule, J., Shorbel, A. & Dobes, C. (2011). Implications of hybridisation and cytotypic differentiation in speciation assessed by AFLP and plastid haplotypes – a case study of Potentilla alpicola La Soie. BMC Evolutionary Biology, 12, 132.CrossRefGoogle Scholar
Paun, O., Stuessy, T. F. & Hörandl, E. (2006). The role of hybridization, polyploidization and glaciation in the origin and evolution of the apomictic Ranunculus cassubicus complex. New Phytologist, 171, 223–36.CrossRefGoogle ScholarPubMed
Paun, O.et al. (2007). Genetic and epigenetic alterations after hybridization and genome doubling. Taxon, 56, 649–56.CrossRefGoogle ScholarPubMed
Pauwels, M.et al. (2005). Multiple origin of metallicolous populations of the pseudometallophyte Arabidopsis halleri (Brassicaceae) in central Europe: the cpDNA testimony. Molecular Ecology, 14, 4403–14.CrossRefGoogle ScholarPubMed
Pauwels, M.et al. (2008). Merging methods in molecular and ecological genetics to study the adaptation of plants to anthropogenic metal-polluted sites: implications for phytoremediation. Molecular Ecology, 17, 108–19.CrossRefGoogle Scholar
Pazy, B. & Zohary, D. (1965). The process of introgression between Aegilops polyploids: natural hybridization between A. variabilis, A. ovata and A. biuncialis. Evolution, 19, 385–94.Google Scholar
Pearson, E. S. & Kendall, M. G. (1970). Studies in the history of statistics and probability. London: Griffin.Google Scholar
Pearson, K. (1900). The grammar of science, edn. London: Black.Google Scholar
Pearson, K. (1924). The life, letters and labours of Francis Galton, vol. II. Cambridge: Cambridge University Press.Google Scholar
Pearson, K.et al. (1903). Cooperative investigation on plants. 2. Variation and correlation in Lesser Celandine from diverse localities. Biometrika 2, 145–64.Google Scholar
Pearson, R. G. (2006). Climate change and the migration capacity of species. Trends in Ecology and Evlution, 21, 111–13.Google ScholarPubMed
Pecinka, A.et al. (2011). Polyploidization increases meiotic recombination frequency in Arabidopsis. BMC Biology, 9, 24.CrossRefGoogle ScholarPubMed
Peckham, M. (1959). The origin of species by Charles Darwin. A variorum text. London: Oxford University Press; Philadelphia: University of Pennsylvania Press.Google Scholar
Pelaz, S.et al. (2000). B and C floral organ identity functions require SEPALLATA MADS-box genes. Nature, 405, 200–3.CrossRefGoogle Scholar
Pellew, C. (1913). Note on gametic reduplication in Pisum. Journal of Genetics, 3, 105–6.CrossRefGoogle Scholar
Pellino, M.et al. (2013). Asexual genome evolution in the apomictic Ranunculus auricomus complex: examining the effects of hybridization and mutation accumulation. Molecular Ecology, DOI:10.1111/mec.12533CrossRef
Pennington, W. (1974). The history of British vegetation, edn. London: English Universities Press.Google Scholar
Peñuelas, J.et al. (2007) Migration, invasion and decline: changes in recruitment and forest structure in a warming-linked shift of European beech forest in Catalonia (NE Spain). Ecography, 30, 830–8.CrossRefGoogle Scholar
Percy, D. M. & Cronk, Q. C. B. (1997). Conservation in relation to mating system in Nesohedyotis arborea (Rubiaceae), a rare endemic tree from St. Helena. Biological Conservation, 80, 135–46.CrossRefGoogle Scholar
Perrie, L. R.et al. (2010). Parallel polyploid speciation: distinct sympatric gene-pools of recurrently derived allo-octoploid Asplenium ferns. Molecular Ecology, 19, 2916–32.CrossRefGoogle ScholarPubMed
Perring, F. H. & Farrell, L. (1983). British red data books. Vol. 1, Vascular plants, edn. Lincoln: The Society for the Promotion of Nature Conservation with the financial support of the World Wildlife Fund.Google Scholar
Perring, F. H. & Walters, S. M. (1976). Atlas of the British flora, edn. Wakefield: EP Publishing.Google Scholar
Perrow, M. R. & Davy, A. J. (eds.) (2002). Handbook of ecological restoration. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Peterken, G. F. (1981). Woodland conservation and management. London: Chapman & Hall.CrossRefGoogle Scholar
Petit, C. & Thompson, J. D. (1999). Species diversity and ecological range in relation to ploidy level in the flora of the Pyrenees. Evolutionary Ecology, 13, 45–66.CrossRefGoogle Scholar
Petit, R. J.et al. (1997). Chloroplast DNA footprints of postglacial recolonization by oaks. Proceedings of the National Academy of Sciences, USA, 94, 9996–10001.CrossRefGoogle ScholarPubMed
Petit, R. J.et al. (2003). Glacial refugia: hotspots but not melting pots of genetic diversity. Science, 300, 1563–5.CrossRefGoogle Scholar
Peuke, A. D & Rennenberg, H. (2005). Phytoremediation. Embo Reports, 6, 497–501.CrossRefGoogle ScholarPubMed
Pharis, R. P. & Ferrell, W. K. (1966). Differences in drought resistance between coastal and inland sources of Douglas Fir. Canadian Journal of Botany, 44, 1651–9.CrossRefGoogle Scholar
Phy-Olsen, A. (2010). Evolution, creationism, and intelligent design. Westport, CT: Greenwood Press.Google Scholar
Pickett, S. T. A., Parker, V. T. & Fiedler, P. L. (1992). The new paradigm in ecology: implications for conservation above the species level. In Conservation biology: the theory and practice of nature conservation, preservation and management, ed. Fiedler, P. L. & Jain, S. K., pp. 65–88. New York: Chapman & Hall.Google Scholar
Pietsch, T. W. (2012). Trees of life: a visual history of evolution. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Pigliucci, M. (2001). Phenotypic plasticity: beyond nature and nurture. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Pigliucci, M. (2002). Denying evolution: creationism, scientism and the nature of science. Sunderland, MA: Sinauer.Google Scholar
Pigliucci, M. (2010). Phenotypic plasticity. In Evolution – the extended synthesis, ed. Pigliucci, M. & Müller, G. B., pp. 355–78. Cambridge, MA: MIT Press.CrossRefGoogle Scholar
Pijul, P. M.et al. (2012). In vitro propagation of tropical hardwood tree species – a review 2001–2011. Propagation of Ornamental Plants, 12, 25–51.Google Scholar
Piñeyro-Nelson, A.et al. (2009). Transgenes in Mexican maize: molecular evidence and methodological considerations for GMO detection in landrace populations. Molecular Ecology, 18, 750–61.CrossRefGoogle ScholarPubMed
Pires, C. J.et al. (2004a). Molecular cytogenetic analysis of recently evolved Tragopogon (Asteraceae) allopolyploids reveal a karyotype that is additive of the diploid progenitors. American Journal of Botany, 91, 1022–35.CrossRefGoogle ScholarPubMed
Pires, J. C.et al. (2004b). Flowering time divergence and genomic rearrangements in resynthesized Brassica polyploids (Brassicaceae). Biological Journal of the Linnean Society, 82, 675–88.CrossRefGoogle Scholar
Pires, N. D. & Dolan, L. (2012). Morphological evolution in land plants: new designs with old genes. Philosophical Transactions of the Royal Society, B, 367, 508–18.CrossRefGoogle ScholarPubMed
Pitman, N. C. A. & Jørgensen, P. M. (2002). Estimating the size of the threatened world flora. Science, 298, 989.CrossRefGoogle Scholar
Pivard, S.et al. (2008). Where do the feral oilseed rape populations come from? A large-scale study of their possible origin in a farmland area. Journal of Applied Ecology, 45, 476–85.CrossRefGoogle Scholar
Pleijel, F. (1999). Phylogenetic taxonomy, a farewell to species, and a revision of Heteropodarke (Annelida, Polychaeta, Hesionidae). Systematic Biology, 48, 755–89.CrossRefGoogle Scholar
Podani, J. (2010a). Monophyly and paraphyly: a discourse without end?Taxon, 59, 1011–15.Google Scholar
Podani, J. (2010b). Taxonomy in evolutionary perspective. An essay on the relationships between taxonomy and evolutionary theory. Synbiologia Hungarica, 6, 1–42.Google Scholar
Pollard, A. J. (1980). Diversity of metal tolerances in Plantago lanceolata L. from the southeastern United States. New Phytologist, 86, 109–17.CrossRefGoogle Scholar
Pollard, A. J. & Baker, A. J. M. (1997). Deterrence of herbivory by zinc hyperaccumulation in Thlaspi caerulescens (Brassicaceae). New Phytologist, 135, 655–8.CrossRefGoogle Scholar
Pond, W. G. & Pond, K. R. (2002). Introduction to animal science. New York & Chichester, UK: Wiley.Google Scholar
Pope, O. A., Simpson, D. M. & Duncan, E. N. (1944). Effect of Corn barriers on natural crossing in Cotton. Journal of Agricultural Research, 68, 347–61.Google Scholar
Popp, M., Mirré, V. & Brochmann, C. (2011). A single mid-Pleistocene long-distance dispersal by a bird can explain the extreme bipolar disjunction in crowberries (Empetrum). Proceedings of the National Academy of Sciences, USA, 108, 6520–5.CrossRefGoogle Scholar
Popper, K. (1963). Conjectures and refutations: the growth of scientific knowledge. London: Routledge.Google Scholar
Porter, D. M. (1983). Vascular plants of the Galápagos: origins and dispersal. In Patterns of evolution in Galápagos organisms, ed. Bowman, R. I., Berson, M. & Levitan, A. E., pp. 33–96. San Francisco: American Association for the Advancement of Science.Google Scholar
Porter, T. M. (1986). The rise of statistical thinking, 1820–1900. Princeton, NJ: Princeton University Press.Google Scholar
Porter, T. M. (2004). Karl Pearson: the scientific life in a statistical age. Princeton, NJ: Princeton University Press.Google Scholar
Potato Genome Sequencing Consortium (2011). Genome sequence and analysis of the tuber crop potato. Nature, 475, 189–95.
Potts, S. G.et al. (2010). Global pollinator declines: trends, impacts and drivers. Trends in Ecology & Evolution, 25, 345–53.CrossRefGoogle ScholarPubMed
Potvin, C. (1986). Biomass allocation and phenological differences among southern and northern populations of the C4 grass Echinochloa crus-galli. Journal of Ecology, 74, 915–23.CrossRefGoogle Scholar
Powles, S. B. (2008). Evolved glyphosate-resistant weeds around the world: lessons to be learnt. Pest Management Science, 64, 360–5.CrossRefGoogle ScholarPubMed
Powles, S. B. & Yu, Q. (2010). Evolution in action: plants resistant to herbicides. Annual Review of Plant Biology, 61, 317–47.CrossRefGoogle ScholarPubMed
Prance, G. T. (2004). Introduction. In Ex situ plant conservation, ed. Guerrant, E. O., Havens, K. & Maunder, M., pp. xxiii–xxix. Washington DC, Covelo, CA, & London: Island Press.Google Scholar
Prendini, L. (2005). Comment on ‘Identifying spiders through DNA barcodes’. Canadian Journal of Zoology, 83, 498–504.CrossRefGoogle Scholar
Prentice, H. C. (1986). Climate and clinal variation in seed morphology of the White Campion, Silene latifolia (Caryophyllaceae). Biological Journal of the Linnean Society, 27, 179–89.CrossRefGoogle Scholar
Presgraves, D. C. (2010). The molecular evolutionary basis of species formation. Nature Reviews Genetics, 11, 175–80.CrossRefGoogle ScholarPubMed
Presgraves, D. C. (2013). Primer: hitchhiking to speciation. PLoS Biology, 11: e1001498.CrossRefGoogle Scholar
Pridgeon, A. M.et al. (1997). Phylogenetics of subtribe Orchidineae (Orchidoideae, Orchidaceae) based on nuclear ITS sequences. 1. Intergeneric relationships and polyphyly of Orchis sensu lato. Lindleyana, 12, 89–109.Google Scholar
Primack, R. B. (1993). Essentials of conservation biology. Sunderland, MA: Sinauer.Google Scholar
Primack, R. B. (2010). Essentials of conservation biology, edn. Sunderland, MA: Sinauer Associates.Google Scholar
Primack, R. & Dayton, B. (1997). The experimental ecology of reintroduction. Plant Talk, 97, 25–8.Google Scholar
Primack, R. B. & Miao, S. L. (1992). Dispersal can limit local plant distribution. Conservation Biology, 6, 513–19.CrossRefGoogle Scholar
Prime, C. T. (1960). Lords and ladies. London: Collins.Google Scholar
Prober, S. M. & Brown, A. H. D. (1994). Conservation of the Grassy White Box woodlands: population genetics and fragmentation of Eucalyptus albens. Conservation Biology, 8, 1003–13.CrossRefGoogle Scholar
Proctor, J. (1971a). The plant ecology of serpentine. II. Plant response to serpentine soils. Journal of Ecology, 59, 397–410.CrossRefGoogle Scholar
Proctor, J. (1971b). The plant ecology of serpentine. III. The influence of a high magnesium/calcium ratio and high nickel and chromium levels in some British and Swedish serpentine soils. Journal of Ecology, 59, 827–42.CrossRefGoogle Scholar
Proctor, M. C. F. & Yeo, P. F. (1973). The pollination of flowers. London: Collins.Google Scholar
Proctor, M. C. F., Proctor, M. E. & Groenhof, A. C. (1989). Evidence from peroxidase polymorphism on the taxonomy and reproduction of some Sorbus populations in south-west England. New Phytologist, 112, 569–75.CrossRefGoogle ScholarPubMed
Proctor, M. C. F., Yeo, P. F. & Lack, A. J. (1996). The natural history of pollination. London: Harper Collins.Google Scholar
Provine, W. B. (1971). The origins of theoretical population genetics. Chicago & London: University of Chicago Press.Google Scholar
Provine, W. B. (1986). Sewall Wright and evolutionary biology. Chicago: University of Chicago Press.Google Scholar
Provine, W. B. (1987). The origins of theoretical population genetics, edn. Chicago: University of Chicago Press.Google Scholar
Putnam, A. R. & Tang, C.-S. (1986). The science of allelopathy. New York: Wiley.Google Scholar
Qiu, Y.-L.et al. (1999). The earliest angiosperms: evidence from mitochondrial, plastid and nuclear genomes. Nature, 402, 404–7.CrossRefGoogle ScholarPubMed
Quarin, C. L. (1986). Seasonal changes in the incidence of apomixis of diploid, triploid, and tetraploid plants of Paspalum cromyorrizon. Euphytica, 35, 515–22.CrossRefGoogle Scholar
Quarin, C. L. & Hanna, W. W. (1980). Effect of three ploidy levels on meiosis and mode of reproduction in Paspalum hexastachyum. Crop Science, 20, 69–75.CrossRefGoogle Scholar
Quattrocchio, F.et al. (1999). Molecular analysis of the anthocyanin2 gene of Petunia and its role in the evolution of flower color. The Plant Cell, 11, 1433–44.CrossRefGoogle ScholarPubMed
Queller, D. C. (1987). Sexual selection in flowering plants. In Sexual selection: testing the alternatives, ed. Bradbury, J. W. & Andersson, M. B., pp. 165–79. Chichester: Wiley.Google Scholar
Quetelet, M. A. (1846). Lettres à S.A.R. Ie Duc Régnant de Saxe-Coburg et Gotha, sur la théorie des probabilités, appliquée aux sciences morales et politiques. Brussels. Translation by Downes, O. G. (1849): Letters addressed to H.R.H. the Grand Duke of Saxe-Coburg and Gotha on the theory of probabilities as applied to the moral and political sciences. London: Charles & Edwin Layton.Google Scholar
Quinn, J. A. (1978). Plant ecotypes: ecological or evolutionary units. Bulletin of the Torrey Botanical Club, 105, 58–64.CrossRefGoogle Scholar
Quintana-Ascencio, P. F. & Menges, E. S. (1996). Inferring metapopulation dynamics from patch-level incidence of Florida scrub plants. Conservation Biology, 10, 1210–19.CrossRefGoogle Scholar
Quist, D. & Chapela, I. H. (2001). Transgenic DNA introgressed into traditional maize landraces in Oaxaca, Mexico. Nature, 414, 541–3.CrossRefGoogle ScholarPubMed
Rabinowitz, D., Cairns, S. & Dillon, T. (1986). Seven forms of rarity and their frequency in the flora of the British Isles. In Conservation biology, ed. Soulé, M. E., pp.182–204. Sunderland, MA: Sinauer.Google Scholar
Rackham, O. (1975). Hayley Wood. Its history and ecology. Cambridge: Cambridgeshire and Isle of Ely Naturalists Trust.Google Scholar
Rackham, O. (1980). Ancient woodland: its history, vegetation and uses in England. London: Arnold.Google Scholar
Rackham, O. (2008). Ancient woodlands: modern threats. New Phytologist, 180, 571–86.CrossRefGoogle ScholarPubMed
Radford, A. E., Dickison, W. C., Massey, J. R. & Bell, R. (1974). Vascular plant systems. New York: Harper & Row.Google Scholar
Rafferty, N. E. & Ives, A. R. (2011). Effects of experimental shifts in flowering phenology on plant–pollinator interactions. Ecology Letters, 14, 69–74.CrossRefGoogle ScholarPubMed
Rafinski, J. N. (1979). Geographic variability of fiower colour in Crocus scepusiensis (Iridaceae). Plant Systematics and Evolution, 131, 107–25.CrossRefGoogle Scholar
Raijmann, L. E. L., van Leeuwen, N. C., Kersten, R., Oostermeijer, J. G. B., Den Nijs, H. C. N. & Menken, S. B. J. (1994). Genetic variation and outcrossing rate in relation to population size in Gentiana pneumonathe L. Conservation Biology, 8, 1014–26.
Ramsbottom, J. (1938). Linnaeus and the species concept. Proceedings of the Linnean Society of London, 150, 192–219.CrossRefGoogle Scholar
Ramsey, J. & Schemske, D. W. (1998). Pathways, mechanisms, and rates of polyploid formation in flowering plants. Annual Review of Ecology and Systematics, 29, 467–501.CrossRefGoogle Scholar
Ramsey, J. & Schemske, D. W. (2002). Neopolyploidy in flowering plants. Annual Review of Ecology & Systematics, 33, 589–639.CrossRefGoogle Scholar
Ramsey, M. W., Cairns, S. C. & Vaughton, G. V. (1994). Geographic variation in morphological and reproductive characters of coastal and tableland populations of Blandfordia grandiflora. Plant Systematics & Evolution, 192, 215–30.CrossRefGoogle Scholar
Randolph, L. F., Nelson, I. S. & Plaisted, R. L. (1967). Negative evidence of introgression affecting the stability of Louisiana Iris species. Cornell University Agriculture Experimental Station Memoir, No. 398.
Ranker, T. A., Floyd, S. K. & Trapp, P. G. (1994). Multiple colonizations of Asplenium adiantum-nigrum onto the Hawaiian archipelago. Evolution, 48, 1364–7.CrossRefGoogle ScholarPubMed
Rasmussen, R. S.et al. (2009). DNA barcoding of commercially important salmon and trout species (Oncorhynchus and Salmo) from North America. Journal of Agricultural and Food Chemistry, 57, 8379–85.CrossRefGoogle ScholarPubMed
Raubeson, L. A. & Jansen, R. K. (2005). Chloroplast genomes of plants. In Plant diversity and evolution: genotypic and phenotypic variation in higher plants, ed. Henry, R. J., pp.45–68. Wallingford, UK; CABI Publishing.Google Scholar
Raup, D. M. (1991). Extinction. Oxford: Oxford University Press.Google ScholarPubMed
Raven, C. E. (1950). John Ray: naturalist, edn; reissued 1986. Cambridge: Cambridge University Press.Google Scholar
Raven, P. H. (1976). Systematics and plant population biology. Systematic Botany, 1, 284–316.CrossRefGoogle Scholar
Raven, P. H. & Thompson, H. J. (1964). Haploidy and angiosperm evolution. The American Naturalist, 98, 251–2.CrossRefGoogle Scholar
Raven, P. H.et al. (1960). Chromosome numbers in Compositae. I. Astereae. American Journal of Botany, 47, 124–32.CrossRefGoogle Scholar
Ravi, M. & Chan, S. W. L. (2010). Haploid plants produced by centromere-mediated genome elimination. Nature, 464, 615–18.CrossRefGoogle ScholarPubMed
Ray, J. (1691). The Wisdom of God Manifested in the Works of Creation, Facsimile edition of the 1826 edition published in 2005. London: Scion Publishing for The Ray Society to mark the 300th anniversary of John Ray's death in 1705.Google Scholar
Ray, M. F. (1995). Systematics of Lavatera and Malva (Malvaceae, Malveae) – a new perspective. Plant Systematics and Evolution, 198, 25–53.CrossRefGoogle Scholar
Raybould, A. F. (1995). Wild crops. In Encyclopedia of environmental biology, vol. 3, ed. Nierenberg, W. A., pp.551–65. San Diego, CA, & New York: Academic Press.Google Scholar
Raybould, A. F., Gray, A. J., Lawrence, M. J. & Marshall, D. F. (1990). The origin and taxonomy of Spartina × neyrautii Foucaud. Watsonia, 18, 207–9.Google Scholar
Raybould, A. F.et al. (1991a). The evolution of Spartina anglica. C. E. Hubbard (Gramineae): origin and genetic variation. Biological Journal of the Linnean Society, 43, 111–26.CrossRefGoogle Scholar
Raybould, A. F.et al. (1991b). The evolution of Spartina anglica C. E. Hubbard (Gramineae): genetic variation and status of the parental species in Britain. Biological Journal of the Linnean Society, 44, 369–80.CrossRefGoogle Scholar
Rayner, A. A. (1969). A first course in biometry for agricultural students. Pietermaritzburg: University of Natal Press.Google Scholar
Redford, K. H., Jensen, D. B. & Breheny, J. J. (2012). Integrating the captive and the wild. Science, 338, 1157–8.CrossRefGoogle Scholar
Ree, R. H. & Smith, S. A. (2008). Maximum likelihood inference of geographic range evolution by dispersal, local extinction and cladogenesis. Systematic Biology, 57, 4–14.CrossRefGoogle ScholarPubMed
Reed, D. H. & Frankham, R. (2001). How closely correlated are molecular and quantitative measures of genetic variation? A meta-analysis. Evolution, 55, 1095–1103.CrossRefGoogle ScholarPubMed
Rees, H. & Hutchinson, J. (1973). Nuclear DNA variation due to B chromosomes. Cold Spring Harbor in Quantitative Biology, 38, 175–82.Google Scholar
Rees, H. & Jones, R. N. (1977). Chromosome genetics. London: Arnold.Google Scholar
Régnière, J. & Bentz, B. J. (2007). Modeling cold tolerance in the mountain pine beetle, Dendroctonus ponderosae. Journal of Insect Physiology, 53, 559–72.CrossRefGoogle ScholarPubMed
Rehfeld, G. E.et al. (2002). Intraspecific responses to climate in Pinus sylvestris. Global Change Biology, 8, 912–29.Google Scholar
Reichman, J. Ret al. (2006). Establishment of transgenic herbicide-resistant creeping bentgrass (Agrostis stolonifera L.) in nonagronomic habitats. Molecular Ecology, 15, 4243–55.CrossRefGoogle Scholar
Reigosa, M. J., Pedrol, N. & González, L. (eds.) (2010). Allelopathy: a physiological process with ecological implications. Dordrecht: Springer.Google Scholar
Reiling, K. & Davison, A. W. (1992). Spatial variation in ozone resistance of British populations of Plantago major L. New Phytologist, 122, 699–708.
Reinartz, J. A. & Les, D. H. (1994). Bottleneck-induced dissolution of self-incompatibility and breeding system consequences in Aster furcatus. American Journal of Botany, 81, 446–55.CrossRefGoogle Scholar
Renfrew, J. M. (1973). Paleoethnobotany: the prehistoric food plants of the Near East and Europe. London: Methuen.Google Scholar
Renner, S. S. (2004). Tropical trans-Atlantic disjunctions, sea surface currents, and wind patterns. International Journal of Plant Sciences, 165, S23–S33.CrossRefGoogle Scholar
Renner, S. S. (2014). The relative and absolute frequencies of angiosperm sexual systems: dioecy, monoecy, gynodioecy and an updated online database. American Journal of Botany, 101, 1588–96.CrossRefGoogle ScholarPubMed
Renner, S. S. & Ricklefs, R. E. (1995). Dioecy and its correlates in the flowering plants. American Journal of Botany, 82, 596–606.CrossRefGoogle Scholar
Reuther, W., Batchelor, L. D. & Webber, H. J. (1968). The citrus industry: California. Vol. 1. Berkeley: University of California, Division of Agricultural Sciences.Google Scholar
Reyer, C.et al. (2013). A plant's perspective of extremes: terrestrial plant responses to climatic variability. Global Change Biology, 19, 79–89.CrossRefGoogle ScholarPubMed
Rhoné, B.et al. (2010). Evolution of flowering time in experimental wheat populations: a comprehensive approach to detect genetic signatures of natural selection. Evolution, 64, 2110–25.Google ScholarPubMed
Ricciari, A. & Simberloff, D. (2008). Assisted colonization is not a viable conservation strategy. Trends in Ecology and Evolution, 24, 248–53.Google Scholar
Rice, D. W.et al. (2013). Horizontal transfer of entire genomes via mitochondrial fusion in the angiosperm Amborella. Science, 342, 1468–73.CrossRefGoogle ScholarPubMed
Rice, E. L. (1984). Allelopathy, edn. London: Academic Press.Google Scholar
Rice, W. R. & Hostert, E. E. (1993). Laboratory experiments on speciation. Evolution 47, 1637–53.CrossRefGoogle ScholarPubMed
Richards, A. J. (1979). Reproduction in flowering plants. Nature, 278, 306.CrossRefGoogle Scholar
Richards, A. J. (1986). Plant breeding systems. London: George Allen & Unwin [edn 1997].Google Scholar
Richards, A. J. (1993). Primula. London: Batsford.Google Scholar
Richards, A. J. (2003). Apomixis in flowering plants: an overview. Philosophical Transactions of the Royal Society of London, B Biological Science, 358, 1085–93.CrossRefGoogle ScholarPubMed
Richards, A. J. & Ibrahim, H. (1978). Estimation of neighbourhood size in two populations of Primula veris. In The pollination of flowers by insects, ed. Richards, A. J., pp. 165–74. Linnean Society Symposium Series 6. London: Academic Press.Google Scholar
Richardson, A. O. & Palmer, J. D. (2007). Horizontal gene transfer in plants. Journal of Experimental Botany, 58, 1–9.Google ScholarPubMed
Richardson, D. M. (ed.) (2011). Fifty years of invasion ecology: the legacy of Charles Elton. Chichester: Wiley-Blackwell.Google Scholar
Richardson, D. M. & Higgins, S. I. (1999). Pines as invaders in the southern hemisphere. In Ecology and biogeography of Pinus, ed. Richardson, D. M., pp. 450–73. Cambridge: Cambridge University Press.Google Scholar
Ridgman, W. J. (1975). Experimentation in biology. Glasgow: Blackie.Google Scholar
Ridley, H. N. (1930). The dispersal of plants throughout the world. Ashford: L. Reeve & Co., Ltd.Google Scholar
Ridley, M. (1996). Evolution, edn. Cambridge, MA, & Oxford: Blackwell Scientific.Google Scholar
Rieger, R., Michaelis, A. & Green, M. M. (1976). Glossary of genetics and cytogenetics, edn. Berlin, Heidelberg & New York: Springer-Verlag.CrossRefGoogle Scholar
Rieseberg, L. H. (1995). The role of hybridization in evolution: old wine in new skins. American Journal of Botany, 82, 944–53.CrossRefGoogle Scholar
Rieseberg, L. H. (1997). Hybrid origins of plant species. Annual Review of Ecology and Systematics, 28, 359–89.CrossRefGoogle Scholar
Rieseberg, L. H. & Blackman, B. K. (2010). Speciation genes in plants. Annals of Botany, 106, 439–55.CrossRefGoogle ScholarPubMed
Rieseberg, L. H. & Ellstrand, N. C. (1993). What can molecular and morphological markers tell us about plant hybridization?Critical Reviews in Plant Sciences, 12, 213–41.Google Scholar
Rieseberg, L. H. & Soltis, D. E. (1991). Phylogenetic consequences of cytoplasmic gene flow in plants. Evolutionary Trends in Plants, 5, 65–84.Google Scholar
Rieseberg, L. H. & Wendel, J. F. (1993). Introgression and its consequences in plants. In Hybrid zones and the evolutionary process, ed. Harrison, R. G., pp. 70–109. New York: Oxford University Press.Google Scholar
Rieseberg, L. H. & Wendel, J. F. (2004). Plant speciation – rise of the poor cousins. New Phytologist, 161, 3–8.Google Scholar
Rieseberg, L. H. & Willis, J. H. (2007). Plant speciation. Science, 317, 910–14.CrossRefGoogle ScholarPubMed
Rieseberg, L. H., Van Fossen, C. & Desrochers, M. (1995). Hybrid speciation accompanied by genomic reorganization in wild sunflowers. Nature, 375, 313–16.CrossRefGoogle Scholar
Rieseberg, L. H., Church, S. A. & Morjan, C. L. (2003). Integration of populations and differentiation of species. New Phytologist, 161, 59–69.CrossRefGoogle Scholar
Rieseberg, L. H., Wood, T. E. & Baack, E. J. (2006). The nature of plant species. Nature, 440, 524–7.CrossRefGoogle ScholarPubMed
Rieseberg, L. H.et al. (2007). Hybridization and the colonization of novel habitats by annual sunflowers. Genetica, 129, 149–65.CrossRefGoogle ScholarPubMed
Riley, H. P. (1938). A character analysis of colonies of Iris fulva and Iris hexagona var. giganticaerulea and natural hybrids. American Journal of Botany, 25, 727–38.CrossRefGoogle Scholar
Riley, R. (1965). Cytogenetics and the evolution of Wheat. In Essays on crop plant evoluion, ed. Hutchinson, J., pp. 103–22. London: Cambridge University Press.Google Scholar
Riley, R. & Chapman, V. (1958). Genetic control of the cytologically diploid behaviour of hexaploid Wheat. Nature, 183, 713–15.Google Scholar
Riley, R., Unrau, J. & Chapman, V. (1958). Evidence on the origin of the B genome of Wheat. Journal of Heredity, 49, 91–8.CrossRefGoogle Scholar
Ritchie, J. C. (1955a). A natural hybrid in Vaccinium. 1. The structure, performance and chorology of the cross Vaccinium intermedium Ruthe. New Phytologist, 54, 49–67.Google Scholar
Ritchie, J. C. (1955b). A natural hybrid in Vaccinium. 2. Genetic studies in Vaccinium intermedium Ruthe. New Phytologist, 54, 320–35.Google Scholar
Ritchie, J. C. (1987). Postglacial vegetation of Canada. Cambridge: Cambridge University Press.Google Scholar
Rizvi, S. J. H. & Rizvi, V. (1992). Allelopathy: basic and applied aspects. London: Chapman & Hall.CrossRefGoogle Scholar
Roach, D. A. & Wulff, R. D. (1987). Material effects in plants. Annual Review of Ecology & Systematics, 18, 209–35.CrossRefGoogle Scholar
Roberts, H. F. (1929). Plant hybridisation before Mendel. Princeton, NJ: Princeton University Press; London: Oxford University Press.CrossRefGoogle Scholar
Roberts, N. (1989). The Holocene: an environmental history. Oxford: Blackwell.Google Scholar
Robledo-Aruncio, J. J. & Gil, L. (2005). Patterns of pollen dispersal in a small population of Pinus sylvestris L. revealed by total-exclusion paternity analysis. Heredity, 94, 13–22.Google Scholar
Robson, G. C. & Richards, O. W. (1936). The variation of animals in nature. London, New York & Toronto: Longmans, Green and Co.Google Scholar
Rodrigues, A. S. L.et al. (2004). Global gap analysis: priority regions for expanding the global protected-area network. BioScience, 54, 1092–1100.CrossRefGoogle Scholar
Rodriguez-Trelles, F., Tarrio, R. & Ayala, F. J. (2004). Molecular clocks: whence and whither? In Telling the evolutionary time: molecular clocks and the fossil record, ed. Donoghue, P. C. J. and Smith, M. P., pp. 5–26. London: Systematic Association Special Volume; CRC Press.Google Scholar
Rogstad, S. H., Nybom, H. & Schaal, B. A. (1991). The tetrapod DNA fingerprinting M13 repeat probe reveals genetic diversity and clonal growth in Quaking Aspen (Populus tremuloides, Salicaceae). Plant Systematics & Evolution, 175, 115–23.CrossRefGoogle Scholar
Roles, S. J. (1960). Illustrations (Part II) to Flora of the British Isles, Clapham, A. R., Tutin, T. G. & Warburg, E. F.Cambridge: Cambridge University Press.Google Scholar
Romeiras, M. M.et al. (2011). Origin and diversification of the genus Echium (Boraginaceae) in the Cape Verde archipelago. Taxon, 60, 1375–85.Google Scholar
Romme, W. H.et al. (2005). Establishment, persistence, and growth of Aspen (Populus tremuloides) seedlings in Yellowstone National Park. Ecology, 86, 404–18.CrossRefGoogle Scholar
Ronquist, F. & Sanmartin, I. (2011): Phylogenetic methods in biogeography. Annual Review of Ecology, Evolution and Systematics, 42, 441–64.CrossRefGoogle Scholar
Roose, M. L. & Gottlieb, L. D. (1976). Genetic and biochemical consequences of polyploidy in Tragopogon. Evolution, 30, 818–30.CrossRefGoogle ScholarPubMed
Rose, M. R. & Oakley, T. H. (2007). The new biology: beyond the Modern Synthesis. Biology Direct, DOI:10.1186/1745-6150-2-30CrossRef
Rosen, F. (1889). Systematische und biologische Beobachtungen über Erophila verna. Botanische Zeitung, 47, 565–80, 581–91, 597–608, 613–20.Google Scholar
Rosenberg, A. (1994). Instrumental biology or the disunity of science. Chicago: Chicago University Press.Google Scholar
Ross-Craig, S. (1948–73). Drawings of British plants. London: Bell & Sons Ltd.Google Scholar
Rosser, E. M. (1953). A new British species of Senecio. Watsonia, 3, 228–32.Google Scholar
Rotherham, I. D. & Lambert, R. A. (2011). Balancing species history, human culture and scientific insight: introduction and overview. In Invasive and introduced plants and animals, ed. Rotherham, I. D. & Lambert, R. A., pp. 3–18. London: Earthscan Publishing.Google Scholar
Rothrock, P. E. & Reznicek, A. A. (1998). Chromosome numbers in Carex section Ovales (Cyperaceae): additions, variations, and corrections. Sida, 18, 587–92.Google Scholar
Roux, C.et al. (2012). Recent and ancient signature of balancing selection around the S-locus in Arabidopsis halleri and A. lyrata. Molecular Biology and Evolution, 30, 435–47.Google ScholarPubMed
Rowell, T. A. (1984). Further discoveries of the Fen Violet (Viola persicifolia Schreber) at Wicken Fen, Cambridgeshire. Watsonia, 15, 122–3.Google Scholar
Rowell, T. A., Walters, S. M. and Harvey, H. J. (1982). The rediscovery of the Fen Violet, Viola persicifolia Schreber, at Wicken Fen, Cambridgeshire. Watsonia, 14, 183–4.Google Scholar
Rubledo-Arununcio, J. J. & Garcia, C. (2007). Estimation of the seed dispersal kernel from exact identification of source plants. Molecular Ecology, 16, 5098–109.Google Scholar
Rucinska, A. & Puchalski, J. (2011). Comparative molecular studies on the genetic diversity of an ex situ garden collection and its source population of the critically endangered Polish endemic plant Cochlearia polonica E. Fröhlich. Biodiversity & Conservation, 20, 401–13.CrossRefGoogle Scholar
Ruckert, J. (1892). Zur Entwicklungs Geschichte des Ovarioleies bei Selachiern. Anatomischer Anzeiger, 7, 107.Google Scholar
Rudall, P. J. & Bateman, R. M. (2004). Evolution of zygomorphy in monocot flowers: iterative patterns and developmental constraints. New Phytologist, 162, 25–44.CrossRefGoogle Scholar
Ruddiman, W. F. (2013) The Anthropocene. Annual Review of Earth and Planetary Sciences, 41, 45–68.CrossRefGoogle Scholar
Ruhsam, M.et al. (2010). Significant differences in outcrossing rate, self-incompatibility, and inbreeding depression between two widely hybridizing species of Geum. Biological Journal of the Linnean Society, 101, 977–90.CrossRefGoogle Scholar
Ruhsam, M., Hollingsworth, P. M. & Ennos, R. A. (2011) Early evolution in a hybrid swarm between outcrossing and selfing lineages in Geum. Heredity, 107, 246–55, DOI:10.1038/hdy.2011.9.CrossRefGoogle Scholar
Runte, A. (1997). National Parks: the American experience, edn. Lincoln, NE, & London: Nebraska Press.Google Scholar
Ruppel, C. D. (2011). Methane hydrates and contemporary climate change. Nature Education Knowledge, 3, 29.Google Scholar
Ruse, M. (1987). Biological species: natural kinds, individuals or what?British Journal of the Philosophy of Science, 38, 225–42.CrossRefGoogle Scholar
Ruse, M. (2003). Darwin and design: does evolution have a purpose?Cambridge, MA, & London: Harvard University Press.Google Scholar
Ruse, M. (2008). Charles Darwin. Oxford: Blackwell.CrossRefGoogle Scholar
Ruse, M. (2013). The Cambridge encyclopedia of Darwin and evolutionary thought. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Rushton, B. S. (1978). Quercus robur L. and Quercus petraea (Matt.) Liebl: a multivariate approach to the hybrid problem. 1. Data acquisition, analysis and interpretation. Watsonia, 12, 81–101.Google Scholar
Rushton, B. S. (1979). Quercus robur L. and Quercus petraea (Matt.) Liebl.: a multivariate approach to the hybrid problem. 2. The geographical distribution of population types. Watsonia, 12, 209–24.Google Scholar
Russell, B. (1931). The scientific outlook. London: Allen & Unwin.Google Scholar
Rustad, L. E.et al. (2001). A meta-analysis of the response of soil respiration, net nitrogen mineralization, and above ground growth to experimental ecosystem warming. Oecologia, 126, 543–62.CrossRefGoogle Scholar
Rutschman, F. (2006). Molecular dating of phylogenetic trees: a brief review of current methods that estimate divergence times. Diversity and Distributions, 12, 35–48.CrossRefGoogle Scholar
Saarela, J. M.et al. (2007). Hydatellaceae identified as a new branch near the base of the angiosperm phylogenetic tree. Nature, 446, 312–15.CrossRefGoogle ScholarPubMed
Sachs, J. (1865). Handbuch der Experimental-Physiologie der Pflanzen. Leipzig: Verlag von Wilhelm Engelmann.Google Scholar
Sage, R. F., Christin, P. A. & Edwards, E. J. (2011). The C4 plant lineages of planet Earth. Journal of Experimental Botany, 62, 3155–69.CrossRefGoogle Scholar
Salisbury, B. A. & Kim, J. (2001). Ancestral state estimation and taxon sampling density. Systematic Biology, 50, 557–64.CrossRefGoogle ScholarPubMed
Salmon, A. & Ainouche, M. L. (2010). Polyploidy and DNA methylation: new tools available. Molecular Ecology, 19, 213–15.CrossRefGoogle ScholarPubMed
Salmon, A., Ainouche, M. L. & Wendel, J. F. (2005). Genetic and epigenetic consequences of recent hybridization and polyploidy in Spartina (Poaceae). Molecular Ecology, 14, 1163–75.CrossRefGoogle Scholar
Salmon, A.et al. (2010). Homoeologous non-reciprocal recombination in polyploid cotton. New Phytologist, 186, 123–34.CrossRefGoogle Scholar
Salmon, S. C. & Hanson, A. A. (1964). The principles and practice of agricultural research. London: Leonard Hill.Google Scholar
Salse, J.et al. (2008). Identification and characterization of shared duplications between Rice and Wheat provide new insight into grass genome evolution. Plant and Cell, 20, 11–24.CrossRefGoogle ScholarPubMed
Samis, K. E.et al. (2012). Latitudinal trends in climate drive flowering time clines in North AmericanArabidopsis thaliana. Ecology and Evolution, 2, 1162–80.Google Scholar
Sanderson, M. J. (2002). Estimating absolute rates of molecular evolution and divergence times: a penalized likelihood approach. Molecular Ecology, 19, 101–9.Google ScholarPubMed
Sanderson, M. J. (2007). Construction and annotation of large phylogenetic trees. Australian Systematic Botany, 20, 287–301.CrossRefGoogle Scholar
Sandler, R. (2009). The value of species and the ethical foundations of assisted colonization. Conservation Biology, 24, 424–31.Google ScholarPubMed
Sang, T. & Ge, S. (2007). The puzzle of Rice domestication. Journal of Integrative Biology, 49, 760–8.Google Scholar
Sanmartín, I., Wanntorp, L. & Winkworth, R. C. (2007). West wind drift revisited: testing for directional dispersal in the southern hemisphere using event-based tree fitting. Journal of Biogeography, 34, 398–416.CrossRefGoogle Scholar
Sapp, J. (2003). Genesis: the evolution of biology. Oxford & New York: Oxford University Press.CrossRefGoogle Scholar
Sarasan, V.et al. (2006). Conservation in vitro of threatened plants – progress in the past decade. In Vitro Cellular & Developmental Biology –Plant, 42, 206–14.CrossRefGoogle Scholar
Sarhanova, P.et al. (2012). New insights into the variability of reproduction modes in European populations of Rubus subgen. Rubus: (how sexual are polyploid brambles?Sex Plant Reproduction, 25, 319–35.CrossRefGoogle ScholarPubMed
Sarkar, P. & Stebbins, G. L. (1956). Morphological evidence concerning the origin of the B genome in Wheat. American Journal of Botany, 43, 297–304.CrossRefGoogle Scholar
Särkinen, T.et al. (2012). How to open the treasure chest? Optimising DNA extraction from herbarium specimens. PLOS ONE, 7, 1–9.CrossRefGoogle ScholarPubMed
Saucy, F.et al. (1999). Preferences for acyanogenic white clover (Trifolium repens) in the vole Arvicola terrestris. I. Preliminary results with two varieties (Ladino and Aran) and complementary tests with the slugs Arion ater and A. subfuscus. Journal of Chemical Ecology, 25, 1441–54.CrossRefGoogle Scholar
Sauer, J. D. (1988). Plant migration. Berkeley: University of California Press.Google Scholar
Saunders, E. R. (1897). On discontinuous variation occurring in Biscutella laevigata. Proceedings of the Royal Society, B, 62, 11–26.Google Scholar
Savolainen, V.et al. (2005). Towards writing the Encyclopedia of Life: an introduction to DNA barcoding. Philosophical Transactions of the Royal Society of London, B, 359, 1805–11.Google Scholar
Sax, D. F., Stachowicz, J. J. & Gaines, S. D. (2006). Species invasions: insights into ecology, evolution and biogeography. Sunderland, MA: Sinauer.Google Scholar
Sayre, A. (1975). Rosalind Franklin and DNA. New York & London: Norton & Co.Google Scholar
Scascitelli, M.et al. (2010). Genome scan of hybridizing sunflowers from Texas (Helianthus annuus and H. debilis) reveals asymmetric patterns of introgression and small islands of genomic differentiation. Molecular Ecology, 19, 521–41.CrossRefGoogle ScholarPubMed
Scascitelli, M., Cognet, M. & Adams, K. L. (2010). An interspecific plant hybrid shows novel changes in parental splice forms of genes for splicing factors. Genetics, 184, 975–83.CrossRefGoogle ScholarPubMed
Schaal, B. A. (1980). Measurement of gene flow in Lupinus texensis. Nature, 284, 450–1.CrossRefGoogle Scholar
Schaal, B. A. (1988). Somatic variation and genetic structure in plant populations. In Plant population ecology, ed. Davy, A. J., Hutchings, M. J. & Watkinson, A. R., pp. 47–58. Oxford: Blackwell.Google Scholar
Schaal, B. A., Leverich, W. J. & Rogstad, S. H. (1991). A comparison of methods for assessing genetic variation in plant conservation biology. In Genetics and conservation of rare plants, ed. Falk, D. A. & Holsinger, K. E., pp. 123–34. New York: Oxford University Press.Google Scholar
Schaal, B. A., O'Kane, S. L. & Rogstad, S. H. (1991). DNA in plant populations. Trends in Ecology and Evolution, 6, 329–33.CrossRefGoogle ScholarPubMed
Schat, H., Vooijs, R. & Kuiper, E. (1996). Identical major gene loci for heavy metal tolerances that have independently evolved in different local populations and subspecies of Silene vulgaris. Evolution, 50, 1888–95.CrossRefGoogle ScholarPubMed
Schatlowski, N. & Köhler, C. (2012). Tearing down barriers: understanding the molecular mechanisms of interploidy hybridizations. Journal of Experimental Botany, 63, 6059–67.CrossRefGoogle ScholarPubMed
Schatz, G. E. (2009). Plants on the IUCN Red List: setting priorities to inform conservation. Trends in Plant Science, 14, 638–42.CrossRefGoogle ScholarPubMed
Schiffers, K.et al. (2013). Limited evolutionary rescue of locally adapted populations facing climate change. Philosophical Transactions of the Royal Society, B, 368, 20120083.CrossRefGoogle ScholarPubMed
Schlichting, C. D. (1986). The evolution of phenotypic plasticity in plants. Annual Review of Ecology and Systematics, 17, 667–93.CrossRefGoogle Scholar
Schlichting, C. D. & Levin, D. A. (1984). Phenotypic plasticity of annual Phlox: tests of some hypotheses. American Journal of Botany, 71, 252–60.CrossRefGoogle Scholar
Schlising, R. A. & Turpin, R. A. (1971). Hummingbird dispersal of Delphinium cardinale pollen treated with radioactive iodine. American Journal of Botany, 58, 401–6.CrossRefGoogle Scholar
Schmalhausen, I. I. (1949) Factors of evolution, trans. Dordick, I., ed. Dobzhansky, T.. Philadelphia: Blakiston.Google Scholar
Schmidt, J. (1899). Om ydre faktorers indflydelse paa løvbladets anatomiske bygning hos en af vore strandplanter. Botanisk Tidsskrift, 22, 145–65.Google Scholar
Schmitz, R. J.et al. (2011). Transgenerational epigenetic instability is a source of novel methylation variants. Science, 334, 369–73.CrossRefGoogle ScholarPubMed
Schnable, P. S. & Springer, N. M. (2013). Progress toward understanding heterosis in crop plants. Annual Review of Plant Biology, 64, 71–88.CrossRefGoogle ScholarPubMed
Schoen, D. J. & Lloyd, D. G. (1984). The selection of cleistogamy and heteromorphic diaspores. Biological Journal of the Linnean Society, 23, 303–22.CrossRefGoogle Scholar
Schön, I., Martens, K. & van Dijk, P. (eds.) (2009). Lost sex: the evolutionary biology of parthenogenesis. Dordrecht, Heidelberg, London & New York: Springer.CrossRefGoogle Scholar
Schönswetter, P.et al. (2005). Molecular evidence for glacial refugia of mountain plants in the European Alps. Molecular Ecology, 14, 3547–55.CrossRefGoogle ScholarPubMed
Schorr, G.et al. (2012). Integrating species distribution models (SDMs) and phylogeography for two species of Alpine Primula. Ecology and Evolution, 2, 1260–77.CrossRefGoogle ScholarPubMed
Schrödinger, E. (1944). What is life?London: Cambridge University Press; New York: Macmillan.Google Scholar
Schulte, P.et al. (2010). The Chicxulub Asteroid impact and mass extinction at the Cretaceous–Paleogene boundary. Science, 327, 1214–18.CrossRefGoogle ScholarPubMed
Schupp, E. W., Jordano, P. & Gomaz, J. M. (2010). Seed dispersal effectiveness revisited: a conceptual review. New Phytologist, 188, 333–53.CrossRefGoogle ScholarPubMed
Schwaegerle, K. E. & Schaal, B. A. (1979). Genetic variability and founder effect in the Pitcher Plant Sarracenia purpurea L. Evolution, 33, 1210–18.
Schwanitz, F. (1966). The origin of cultivated plants. Cambridge, MA.: Harvard University Press.Google Scholar
Schwartz, M. W.et al. (2012a). Managed relocation: integrating the scientific, regulatory, and ethical challenges. Bioscience, 62, 732–43.CrossRefGoogle Scholar
Schwartz, M. W.et al. (2012b). Predicting extinctions as a result of climate change. Ecology, 87, 1611–15.Google Scholar
Schweber, S. S. (1977). The origin of the origin revisited. Journal of the History of Biology, 10, 229–316.CrossRefGoogle ScholarPubMed
Scotese, C. R. (2004). Cenozoic and Mesozoic paleogeography: changing terrestrial biogeographic pathways. In Frontiers of biogeography: new directions in the geography of nature, ed. Lomolino, M. V. & Heany, L. R., pp. 9–26. Sunderland, MA: Sinauer Associates.Google Scholar
Scott, J. M. & Schipper, J. (2006). Gap analysis: a spatial tool for conservation planning. In Principles of conservation biology, ed. Groom, M. J., Meffe, G. K. & Carroll, C. R., edn, pp. 518–19. Sunderland, MA: Sinauer.Google Scholar
Seavey, S. R. & Bawa, K. S. (1986). Late-acting self-incompatibility in Angiosperms. Botanical Review, 52, 195–214.CrossRefGoogle Scholar
Seberg, O.et al. (2003). Short cuts in systematics? A commentary on DNA-based taxonomy. Trends in Taxonomy and Evolution, 18, 63–5.Google Scholar
Seddon, P. J. (2010). From reintroduction to assisted colonization: moving along the conservation translocation spectrum. Restoration Ecology, 18, 796–802.CrossRefGoogle Scholar
Seidel, C. F. (1879). Ueber Verwachsungen von Stämmen und Zweigen von Holzgewächsen und ihren Einfluss auf das Dickenwachsthum der betreffenden Theile. Naturwissenschaftliche Gesellschaft Isis, Dresden, Sitzber, 161–8.
Sellars, R. W. (1997). Preserving nature in the National Parks: a history. New Haven, CT, & London: Yale University Press.Google Scholar
Sepkoski, D. (2009). The origin and early reception of punctuated equilibrium. In The paleobiological revolution: essays on the history of recent paleontology, ed. Sepkoski, D. and Ruse, M., pp. 301–25. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Sepkoski, D. & Ruse, M. (eds.) (2009). The paleobiological revolution: essays on the history of recent paleontology. Chicago: University of Chicago Press.CrossRefGoogle Scholar
Sepp, S.et al. (2000). Genetic polymorphism detected with RAPA analysis and morphological variability in some microspecies of apomictic Alchemilla. Anales Botanici Fennici, 37, 105–23.Google Scholar
Sexton, J. P., Strauss, S. Y. & Rice, K. G. (2011). Gene flow increases fitness at the warm edge of a species’ range. Proceedings of the National Academy of Sciences, USA, 108, 11704–9.CrossRefGoogle Scholar
Shafer, A. B. A.et al. (2010). Of glaciers and refugia: a decade of study sheds new light on the phylogeography of northwestern North America. Molecular Ecology, 19, 4589–4621.CrossRefGoogle ScholarPubMed
Shafer, C. L. (1990). Nature reserves. Washington DC: Smithsonian Institution Press.Google Scholar
Shaffer, M. L. (1981). Minimum population sizes for species conservation. Bioscience, 31, 131–4.CrossRefGoogle Scholar
Shaked, H.et al. (2001). Sequence elimination and cytosine methylation are rapid and reproducible responses of the genome to wide hybridization and allopolyploidy in Wheat. Plant and Cell, 13, 1749–59.CrossRefGoogle ScholarPubMed
Shapcott, A. (1998). The genetics of Ptychosperma bleeseri, a rare palm from the Northern Territory, Australia. Biological Conservation, 85, 203–9.CrossRefGoogle Scholar
Shapiro, B. & Hofreiter, M. (eds.) (2012). Ancient DNA: methods and protocols. New York: Humana Press.CrossRefGoogle Scholar
Sharbel, T. F.et al. (2005). Biogeographic distribution of polyploidy and B chromosomes in the apomictic Boechera holboellii complex. Cytogenetic and Genome Research, 109, 283–92.CrossRefGoogle Scholar
Sharrock, S. & Jones, M.. (2011). Saving Europe's threatened flora: progress towards GSPC Target 8 in Europe. Biological Conservation, 20, 325–33.Google Scholar
Sheail, J., Treweek, J. R. & Mountford, J. O. (1997). The UK transition from nature preservation to ‘creative conservation’. Environmental Conservation, 24, 224–35.CrossRefGoogle Scholar
Sheffield, E., Wolf, P. G., Rumsey, F. J., Robson, D. J., Ranker, T. A. & Challiner, S. M. (1993). Spatial distribution and reproductive behaviour of a triploid bracken (Pteridium aquilinum) clone in Britain. Annals of Botany, 72, 231–7.CrossRefGoogle Scholar
Shepherd, L. D., De Lange, P. J. & Perrie, L. R. (2009). Multiple colonizations of a remote oceanic archipelago by one species: how common is long-distance dispersal?Journal of Biogeography, 36, 1972–7.CrossRefGoogle Scholar
Sherman, M. (1946). Karyotype evolution: a cytogenetic study of seven species and six interspecific hybrids of Crepis. University of California Publications in Botany, 18, 369–408.Google Scholar
Shindo, C., Bernasconi, G. & Hardtke, C. S. (2007). Natural genetic variation in Arabidopsis: tools, traits and prospects. Annals of Botany, 99, 1043–54.CrossRefGoogle ScholarPubMed
Shirley, P. D. & Lamberti, G. A. (2010). Assisted colonization under the US Endangered Species Act. Conservation Letters, 3, 45–52.Google Scholar
Shivas, M. G. (1961a). Contributions to the cytology and taxonomy of species of Polypodium in Europe and America. 1. Cytology. Journal of the Linnean Society, 58, 13–25.Google Scholar
Shivas, M. G. (1961b). Contributions to the cytology and taxonomy of species of Polypodium in Europe and America. 2. Taxonomy. Journal of the Linnean Society, 58, 27–38.Google Scholar
Shore, J. S. & Barrett, S. C. H. (1985). The genetics of distyly and homostyly in Turnera ulmifolia L. (Turneraceae). Heredity, 55, 167–74.CrossRefGoogle Scholar
Sicard, A. & Lenhard, M. (2011). The selfing syndrome: a model for studying the genetic and evolutionary basis of morphological adaptation in plants. Annals of Botany, 107, 1433–43.CrossRefGoogle ScholarPubMed
Silvertown, J. (1984). Phenotypic variety in seed germination behaviour: the ontogeny and evolution of somatic polymorphism in seeds. American Naturalist, 124, 1–16.CrossRefGoogle Scholar
Silvertown, J. (1992). An experimental test of frequency-dependent fitness in mixtures of the two seed morphs of Spergula arvensis. Acta Oecologica, 13, 627–34.Google Scholar
Silvertown, J. (2001). Plants stand still, but their genes don't: non-trivial consequences of the obvious. In Integrating ecology and evolution in a spatial context, ed. Silvertown, J. & Antonovics, J., pp. 3–20. Oxford: Blackwell.Google Scholar
Silvertown, J. & Charlesworth, D. (2001). Introduction to plant population biology. Malden, MA, & Oxford: Blackwell Publishing.Google Scholar
Silvertown, J. W. & Lovett Doust, J. (1993). Introduction to plant population biology. Oxford: Blackwell.Google Scholar
Silvertown, J.et al. (2005). Reinforcement of reproductive isolation between adjacent populations in the Park Grass Experiment. Heredity, 95, 198–205.CrossRefGoogle ScholarPubMed
Silvertown, J.et al. (2010). Environmental myopia: a diagnosis and a remedy. Trends in Ecology & Evolution, 25, 556–61.CrossRefGoogle Scholar
Simberloff, D. (2001). Introduced species, effects and distribution of. In Encyclopedia of biodiversity, vol. 3, ed. Levin, S. A., pp. 517–29. San Diego, CA: Academic Press.Google Scholar
Simberloff, D. (2013). Invasive species: what everyone needs to know. Oxford: Oxford University Press.Google Scholar
Simberloff, D. & Rejmánek, M. (eds.) (2011). Encyclopedia of biological invasions, Encyclopedias of the Natural World, No. 3. Berkeley, CA: University of California Press.Google Scholar
Simberloff, D. S., Farr, J. A., Cox, J. & Mehlman, D. W. (1992). Movement corridors: conservation bargains or poor investments?Conservation Biology, 6, 493–504.CrossRefGoogle Scholar
Simmonds, N. W. (ed.) (1976). Evolution of crop plants. London: Longmans.Google Scholar
Simpson, D. M. (1954). Natural cross-pollination in Cotton. U.S. Department of Agriculture Technical Bulletin, No. 1094.
Simpson, G. G. (1944). Tempo and mode in evolution. New York: Columbia University Press.Google Scholar
Simpson, G. G. (1961). Principles of animal taxonomy. The species and lower categories. New York: Columbia University Press.Google Scholar
Simpson, G. G. (1980). Splendid isolation: the curious history of South American mammals. New Haven, CT: Yale University Press.Google Scholar
Simpson, M. G. (1986). Phylogeny and structural evolution of plants. In Fundamentals of plant systematics, ed. Radford, A. E., pp. 217–48. New York: Harper & Row.Google Scholar
Simunek, M., Hossfeld, U. & Breidbach, O. (2012). ‘Further development’ of Mendel's legacy? Erich von Tschermak-Seysenegg in the context of the Mendelian-biometry controversy, 1901–1906. Theory in Biosciences, 131, 243–52.CrossRefGoogle Scholar
Sindu, A. S. & Singh, S. (1961). Studies on the agents of cross pollination of Cotton. Indian Cotton Growing Review, 15, 341–53.Google Scholar
Siol, M., Wright, S. I. & Barrett, S. C. H. (2010). The population genomics of plant adaptation. New Phytologist, 188, 313–32.CrossRefGoogle ScholarPubMed
Sletvold, N.et al. (2010). Cost of trichome production and resistance to a specialist insect herbivore in Arabidopsis lyrata. Evolutionary Ecology, 24, 1307–19.CrossRefGoogle Scholar
Sluys, R. (2013). The unappreciated, fundamentally analytic nature of taxonomy and implications for the inventory of biodiversity. Biodiversity and Conservation, 22, 1095–105.CrossRefGoogle Scholar
Smartt, J. O. & Simmonds, N. W. (1995). Evolution of crop plants, edn. Harlow: Longmans.Google Scholar
Smith, A. (1965). The assessment of patterns of variation in Festuca rubra L. in relation to environmental gradients. Scottish Plant Breeding Station Record, 1965, 163–95.Google Scholar
Smith, A. (1972). The pattern of distribution of Agrostis and Festuca plants of various genotypes in a sward. New Phytologist, 71, 937–45.CrossRefGoogle Scholar
Smith, A. C. (1957). Fifty years of botanical nomenclature. Brittonia, 9, 2–8.CrossRefGoogle Scholar
Smith, A. M.et al. (2010). Plant biology. New York & Abingdon, UK: Garland Science, Taylor & Francis.Google Scholar
Smith, C. H. & Beccaloni, G. (eds.) (2008). Natural selection and beyond: the intellectual legacy of Alfred Russel Wallace. Oxford: Oxford University Press.Google Scholar
Smith, D. C., Nielsen, E. L. & Ahlgren, H. L. (1946). Variation in ecotypes of Poa pratensis. The Botanical Gazette, 108, 143–66.CrossRefGoogle Scholar
Smith, D. M. & Levin, D. A. (1963). A chromatographic study of reticulate evolution in the Appalachian Asplenium complex. American Journal of Botany, 50, 952–8.CrossRefGoogle Scholar
Smith, F. D. M.et al. (1993). How much do we know about the current extinction rate?Trends in Ecology and Evolution, 8, 375–8.CrossRefGoogle ScholarPubMed
Smith, G. L. (1963a). Studies in Potentilla L. 1. Embryological investigations into the mechanism of agamospermy in British P. tabernaemontani Aschers. New Phytologist, 62, 264–82.CrossRefGoogle Scholar
Smith, G. L. (1963b). Studies in Potentilla L. 2. Cytological aspects of apomixis in P. crantzii (Cr.) Beck ex Fritsch. New Phytologist, 62, 283–300.CrossRefGoogle Scholar
Smith, G. L. (1971). Studies in Potentilla L. 3. Variation in British P. tabernaemontani Aschers. and P. crantzii (Cr.) Beck ex Fritsch. New Phytologist, 70, 607–18.CrossRefGoogle Scholar
Smith, J. (1841). Notice of a plant which produces perfect seeds without any apparent action of pollen. Transactions of the Linnean Society of London, 18, 509–12.CrossRefGoogle Scholar
Smith, P. M. (1976). The chemotaxonomy of plants. London: Arnold.Google Scholar
Smith, S. (1960). The origin of the Origin. Advancement of Science, 16, 391–401.Google Scholar
Smith, S. A. & Donoghue, M. J. (2010). Combining historical biogeography with niche modeling in the Caprifolium Clade of Lonicera (Caprifoliaceae, Dipsacales). Systematic Biology, 590, 322–41.Google Scholar
Smith, S. A., Beaulieu, J. M. & Donoghue, M. J. (2010). An uncorrelated relaxed-clock analysis suggests an earlier origin for flowering plants. Proceedings of the National Academy of Sciences, USA, 107, 5897–902.CrossRefGoogle ScholarPubMed
Smocovitis, V. B. (1996). Unifying biology: the evolutionary synthesis and evolutionary biology. Princeton, NJ: Princeton University Press.Google Scholar
Smocovitis, V. B. (2006). Keeping up with Dobzhansky: G. Ledyard Stebbins, Jr., plant evolution and the evolutionary synthesis. History & Philosophy of the Life Sciences, 28, 9–48.Google Scholar
Smyth, C. A. & Hamrick, J. L. (1987). Realised gene flow via pollen in artificial populations of Musk Thistle, Carduus nutans. Evolution, 39, 53–65.Google Scholar
Snaydon, R. W. (1970). Rapid population differentiation in a mosaic environment. 1. The response of Anthoxanthum odoratum populations to soils. Evolution, 24, 257–69.CrossRefGoogle Scholar
Snaydon, R. W. (1976). Genetic change within species. In The Park Grass experiment on the effect of fertilisers and liming on the botanical composition of permanent grassland and on the yield of hay, Thurston, J. M., Dyke, G. V. & Williams, E. D., Appendix. Harpenden: Rothamsted Experimental Station.Google Scholar
Snaydon, R. W. (1978). Genetic changes in pasture populations. In Plant relations in pastures, ed. Wilson, J. R., pp. 253–69. Melbourne: CSIRO.Google Scholar
Snaydon, R. W. & Davies, M. S. (1972). Rapid population differentiation in a mosaic environment. II. Morphological variation in Anthoxanthum odoratum. Evolution, 26, 390–405.CrossRefGoogle Scholar
Sneath, P. H. A. (1962). The construction of taxonomic groups. In Microbial classification, ed. Ainsworth, G. C. & Sneath, P. H. A.. pp. 289–332. Cambridge: Cambridge University Press.Google Scholar
Sneath, P. H. A. (1988). The phenetic and cladistic approaches. In Prospects in systematics. The Systematic Association Special Volume No. 36, ed. Hawksworth, D. L., pp. 252–73. Oxford: Clarendon Press.Google Scholar
Sneath, P. H. A. (1995). Thirty years of numerical taxonomy. Systematic Biology, 44, 281–98.CrossRefGoogle Scholar
Sneath, P. H. A. & Sokal, R. R. (1973). Numerical taxonomy. San Francisco: Freeman.Google Scholar
Snedecor, G. W. & Cochran, W. G. (1980). Statistical methods, edn. Ames, IA: lowa State University Press.Google Scholar
Snow, A. (2012). Illegal gene flow from transgenic creeping bent grass: the saga continues. Molecular Ecology, 21, 4663–4.CrossRefGoogle Scholar
Snow, A. A.et al. (2003). A Bt transgene reduces herbivory and enhances fecundity in wild sunflowers. Ecological Applications, 13, 279–86.CrossRefGoogle Scholar
Snyder, L. A. (1950). Morphological variability and hybrid development in Elymus glaucus. American Journal of Botany, 37, 628–35.CrossRefGoogle Scholar
Snyder, L. A. (1951). Cytology of inter-strain hybrids and the probable origin of variability in Elymus glaucus. American Journal of Botany, 38, 195–202.CrossRefGoogle Scholar
Sokal, R. R. & Crovello, T. J. (1970). The biological species concept: a critical evaluation. American Naturalist, 104, 127–53.CrossRefGoogle Scholar
Sokal, R. R. & Rohlf, F. J. (1969). Biometry. The principles and practice of statistics in biological research. [edn 1981.] San Francisco: Freeman.Google Scholar
Sokal, R. R. & Sneath, P. H. A. (1963). Principles of numerical taxonomy. London & San Francisco: Freeman.Google Scholar
Sokhi, R. S. (2011). World atlas of atmospheric pollution. London, New York & Delhi: Anthem Press.Google Scholar
Solbrig, O. T. (1994). Biodiversity: an introduction. In Biodiversity and global change, ed. Solbrig, O. T., Emden, H. M. van & Oordt, P. G. W. J. van, pp. 13–20. Wallingford, UK: CAB International.Google Scholar
Solbrig, O. T. & Solbrig, D. J. (1979). Introduction to population biology and evolution. London: Addison-Wesley Publishing Company.Google Scholar
Soltis, D. E. & Burleigh, J. G. (2009). Surviving the K-T mass extinction: new perspectives of polyploidization in angiosperms. Proceedings of the National Academy of Sciences, USA, 106, 5455–6.CrossRefGoogle ScholarPubMed
Soltis, D. E. & Soltis, P. S. (1989). Allopolyploid speciation in Tragopogon: insights from chloroplast DNA. American Journal of Botany, 76, 1119–24.CrossRefGoogle Scholar
Soltis, D. E. & Soltis, P. S. (1990). Isozymes in plant biology. London: Chapman & Hall.CrossRefGoogle Scholar
Soltis, D. E. & Soltis, P. S. (1993). Molecular data and the dynamic nature of polyploidy. Critical Reviews in Plant Science, 12, 243–73.CrossRefGoogle Scholar
Soltis, D. E. & Soltis, P. S. (1995). The dynamic nature of polyploid genomes. Proceedings of the National Academy of Sciences, USA, 92, 8089–91.CrossRefGoogle ScholarPubMed
Soltis, D. E. & Soltis, P. S. (1999). Polyploidy: recurrent formation and genome evolution. Trends in Ecology and Evolution, 14, 348–52.CrossRefGoogle ScholarPubMed
Soltis, P. S. & Soltis, D. E. (2004). The origin and diversification of angiosperms. American Journal of Botany, 91, 1614–26.CrossRefGoogle ScholarPubMed
Soltis, D. E., Soltis, P. S. & Tate, J. A. (2004). Advances in the study of polyploidy since Plant speciation. New Phytologist, 161, 173–91.Google Scholar
Soltis, D. S.et al. (2005). Phylogeny and evolution of angiosperms. Sunderland, MA: Sinauer.Google Scholar
Soltis, D. E.et al. (2005). Phylogeny, evolution, and classification of flowering plants. Sunderland, MA: Sinauer.Google Scholar
Soltis, D. E.et al. (2007). Autopolyploidy in angiosperms: have we grossly underestimated the number of species?Taxon, 56, 13–30.Google Scholar
Soltis, D. E.et al. (2008). Origin and early evolution of Angiosperms. Annals of the New York Academy of Sciences, 1133, 3–25.CrossRefGoogle ScholarPubMed
Soltis, D. E.et al. (2009). Polyploidy and angiosperm diversification. American Journal of Botany, 96, 336–48.CrossRefGoogle ScholarPubMed
Soltis, D. E.et al. (2010). What we still don't know about polyploidy. Taxon, 60, 324–32.Google Scholar
Soltis, D. E.et al. (2011). Angiosperm phylogeny: 17 genes, 640 taxa. American Journal of Botany, 98, 704–30.CrossRefGoogle ScholarPubMed
Soltis, D. E.et al. (2012). The early stages of polyploidy: rapid and repeated evolution in Tragopogon. In Polyploidy and genome evolution, ed. Soltis, P. S. & Soltis, D. E., pp. 271–92. Heidelberg: Springer.CrossRefGoogle Scholar
Soltis, D. E., Visger, C. J. & Soltis, P. S. (2014). The polyploidy revolution then … and now: Stebbins revisited. American Journal of Botany, 101, 1057–78.CrossRefGoogle Scholar
Soltis, P. S. & Soltis, D. E (2009). The role of hybridization in plant speciation. Annual Review of Plant Biology, 60, 561–88.CrossRefGoogle ScholarPubMed
Soltis, P. S., Soltis, D. E. & Doyle, J. J. (1992). Molecular systematics of plants. New York: Chapman & Hall.CrossRefGoogle Scholar
Soltis, P. S., Plunkett, G. M., Novak, S. J. & Soltis, D. E. (1995). Genetic variation in Tragopogon species: additional origins of the allopolyploids T. mirus and T. miscellus (Compositae). American Journal of Botany, 82, 1329–41.CrossRefGoogle Scholar
Soltis, P. S. & Soltis, D. E. (1991). Multiple origins of the allotetraploid Tragopogon mirus (Compositae): rDNA evidence. Systematic Botany, 16, 407–13.CrossRefGoogle Scholar
Soltis, P. S.et al. (2009). Floral variation and floral genetics in basal angiosperms. American Journal of Botany, 96, 110–28.CrossRefGoogle ScholarPubMed
Soltis, P. S.et al., (2014). Polyploidy and novelty: Gottlieb's legacy. Philosophical Transactions of the Royal Society, 369; 20130351.http://dx.doi.org/10.1098/rstb.2012.0351.CrossRefGoogle ScholarPubMed
Solymosi, P. & Lehoczki, E. (1989a). Characterization of a triple (atrazine-pyrazon-pyridate) resistant biotype of Common Lambs Quarters (Chenopodium album L.)Journal of Plant Physiology, 134, 685–90.CrossRefGoogle Scholar
Solymosi, P. & Lehoczki, E. (1989b). Co-resistance of atrazine-resistant Chenopodium and Amaranthus biotypes to other photosystem II inhibiting herbicides. Zeitschrift für Naturforschung, 44C, 119–27.Google Scholar
Song, B. H.et al. (2003). Cytoplasmic composition in Pinus densata and population establishment of a diploid hybrid pine. Molecular Ecology, 12, 2995–3001.CrossRefGoogle ScholarPubMed
Song, K. M., Osborn, T. C. & Williams, P. H. (1988). Brassica taxonomy based on nuclear restriction length polymorphisms (RFLPs). 1. Genome evolution of diploid and amphidiploid species. Theoretical and Applied Genetics, 75, 784–94.CrossRefGoogle Scholar
Soorae, P. S. (ed.) (2011). Global re-introduction perspectives: 2011. More case studies from around the globe. Gland, Switzerland: IUCN/SSC.Google Scholar
Sørenson, T. & Gudjónsson, G. (1946). Spontaneous chromosome-aberrants in apomictic Taraxaca. Biologiske Skrifter, K. Danske Videnskabernes Selskab, 4, No. 2.Google Scholar
Sork, V. L. (1984). Examination of seed dispersal and survival in red oak, Quercus rubra (Fagaceae), using metal-tagged acorns. Ecology, 65, 1020–2.CrossRefGoogle Scholar
Soukup, J. & Holec, J. (2004). Crop–wild interaction within the Beta vulgaris complex: agronomic aspects of weed beet in the Czech Republic. In Introgression from genetically modified plants into wild relatives, ed. den Nijs, H. C. M., Bartsch, D. & Sweet, J., pp. 203–218. Wallingford, UK: CAB International.Google Scholar
Soulé, M. E. (1980). Thresholds for survival: maintaining fitness and evolutionary potential. In Conservation: an evolutionary–ecological perspective, ed. Soulé, M. E. & Wilcox, B. A., pp. 151–70. Sunderland, MA: Sinauer.Google Scholar
Soutullo, A. (2010). Extent of the global network of terrestrial protected areas. Conservation Biology, 24, 362–3.Google ScholarPubMed
Specht, C. D. & Bartlett, M. E. (2009). Flower evolution: the origin and subsequent diversification of the angiosperm flower. Annual Review of Ecology, Evolution & Systematics, 40, 217–43.CrossRefGoogle Scholar
Spencer, K. C. (1988). Chemical mediation of coevolution. San Diego, CA: Academic Press.Google Scholar
Sprengel, C. K. (1793). Das entdeckte Geheimniss der Natur im Bau und in der Befruchtung der Blumen. Berlin.CrossRefGoogle Scholar
Srb, A. M. & Owen, R. D. (1958). General genetics. San Francisco: Freeman.Google Scholar
Stace, C. A. (1975). Hybridization and the flora of the British Isles. London: Academic Press.Google Scholar
Stace, C. A. (1980). Plant taxonomy and biosystematics. London: Edward Arnold.Google Scholar
Stace, C. A. (1989). Plant taxonomy and biosystematics, edn. London: Edward Arnold.Google Scholar
Stace, C. A. (1993). The importance of rare events in polyploid evolution. In Evolutionary patterns and processes, ed. Lees, D. R. & Edwards, D.. Linnean Society Symposium Series, 14, 157–68. London: Published for the Linnean Society by Academic Press.Google Scholar
Stace, C. A. (2010a). New flora of the British Isles, edn. Cambridge: Cambridge University Press.Google Scholar
Stace, C. A. (2010b). The new molecular classification: relevance to the flora of the British Isles. Botanical Society of the British Isles: BSBI News, 113, 4–6.Google Scholar
Stace, C. A. (2010c). Classification by molecules: what's in it for field botanists?Watsonia, 28: 103–22Google Scholar
Stace, C. A., Preston, C. D. & Pearman, D. A. (2015). Hybrid flora of the British Isles. Bristol: Botanical Society of Britain and Ireland.Google Scholar
Stapledon, R. G. (1928). Cocksfoot grass (Dactylis glomerata L.): ecotypes in relation to the biotic factor. Journal of Ecology, 16, 72–104.CrossRefGoogle Scholar
Stearn, W. T. (1957). Introduction to facsimile edition of Linnaeus’ Species plantarum. London: Ray Society, British Museum.Google Scholar
Stearns, S. C. (1992). The evolution of life histories. Oxford: Oxford University Press.Google Scholar
Stebbins, G. L. (1947). Types of polyploids: their classification and significance. Advances in Genetics, 1, 403–29.Google ScholarPubMed
Stebbins, G. L. (1950). Variation and evolution in plants. London: Oxford University Press; New York: Columbia University Press.Google Scholar
Stebbins, G. L. (1957). Self fertilization and population variability in the higher plants. The American Naturalist, 91, 337–54.CrossRefGoogle Scholar
Stebbins, G. L. (1966). Processes of organic evolution. Englewood Cliffs, NJ: Prentice-Hall. [edn 1971.]Google Scholar
Stebbins, G. L. (1971). Chromosomal evolution in higher plants. London: Arnold.Google Scholar
Stebbins, G. L. (1974). Flowering plants. evolution above the species level. London: Arnold; Cambridge, MA: Belknap Press.CrossRefGoogle Scholar
Stebbins, G. L. (1984). Polyploidy and the distribution of the arctic–alpine flora: new evidence and a new approach. Botanica Helvetica, 94, 1–13.Google Scholar
Stebbins, G. L. & Dawe, J. C. (1987). Polyploidy and distribution in the European flora: a reappraisal. Botanische Jahrbücher fur Systematik, Pflanzengeschichte und Pflanzengeographie, 108, 343–54.Google Scholar
Stebbins, G. L.et al. (1963). Identification of the ancestry of an amphiploid Viola with the aid of paper chromatography. American Journal of Botany, 50, 830–9.CrossRefGoogle Scholar
Steen, S. W.et al. (2000). Same parental species, but different taxa: molecular evidence for hybrid origins of the rare endemics Saxifraga opdalensis and S. svalbardensis (Saxifragaceae). Botanical Journal of the Linnean Society, 132, 153–64.Google Scholar
Steffen, W.et al. (2011). The Anthropocene: conceptual and historical perspectives. Philosophical Transactions of the Royal Society, A, 369, 842–67.CrossRefGoogle ScholarPubMed
Stegemann, S.et al. (2012). Horizontal transfer of chloroplast genomes between plant species. Proceedings of the National Academy of Sciences, USA, 109, 2434–8.CrossRefGoogle ScholarPubMed
Stehli, F. G. & Webb, S. D. (eds.) (1985). The great American biotic interchange. New York & London: Plenum Press.CrossRefGoogle Scholar
Steinger, T., Körner, C. & Schmid, B. (1996). Long-term persistence in a changing climate: DNA analysis suggests very old ages of clones of alpine Carex curvula. Oecologia, 105, 94–9.CrossRefGoogle Scholar
Stelleman, P. (1978). The possible role of insect visits in pollination of reputedly anemophilous plants, exemplified by Plantago lanceolata and syrphid flies. In The pollination of flowers by insects, ed. Richards, A. J., pp. 41–6. London: Academic Press.Google Scholar
Sterck, L.et al. (2007). How many genes are there in plants (… and why are they there)?Current Opinion in Plant Biology, 10, 199–203.CrossRefGoogle Scholar
Sterelny, K. (2007). Dawkins vs Gould: survival of the fittest. Cambridge: Icon Books.Google Scholar
Stern, C. & Sherwood, E. R. (1966). The origin of genetics. A Mendel source book. London & San Francisco: Freeman.Google Scholar
Sternberg, L. (1976). Growth forms of Larrea tridentata. Madroño, 23, 408–17.Google Scholar
Stevens, P. F. (2006). An end to all things? – plants and their names. Australian Systematic Botany, 19, 115–33.CrossRefGoogle Scholar
Stevens, P. F. (2012). Angiosperm Phylogeny Website – Missouri Botanical Garden www.mobot.org/MOBOT/research/APweb/
Stewart, J. R.et al. (2010). Refugia revisited: individualistic responses of species in space and time. Proceedings of the Royal Society of London, B, 277, 661–71.CrossRefGoogle ScholarPubMed
Stewart, R. N. (1947). The morphology of somatic chromosomes in Lilium. American Journal of Botany, 34, 9–26.CrossRefGoogle ScholarPubMed
Stift, M.et al. (2008). Segregation models for disomic, tetrasomic and intermediate inheritance in tetraploids: a general procedure applied to Rorippa (Yellow Cress) microsatellite data. Genetics, 179, 2113–23.CrossRefGoogle ScholarPubMed
Stigler, S. M. (1986). The history of statistics: the measurement of uncertainty before 1900. Cambridge, MA: Belknap Press of Harvard University Press.Google Scholar
Stiller, J. W., Reel, D. C. & Johnson, J. C. (2003). A single origin of plastids revisited: convergent evolution in organellar gene content. Journal of Phycology, 39, 95–105.CrossRefGoogle Scholar
Stoeckle, M. Y.et al. (2011). Commercial teas highlight plant DNA barcode identification successes and obstacles. Scientific Reports, 1, 42, DOI:101038/srep00042.CrossRefGoogle ScholarPubMed
Stone, J. L., Thomson, J. D. & Dent-Acosta, S. J. (1995). Assessment of pollen viability in hand-pollination experiments: a review. American Journal of Botany, 82, 1186–97.CrossRefGoogle Scholar
Stork, N. E. (2010). Re-assessing current extinction rates. Biodiversity and Conservation, 19, 357–71.CrossRefGoogle Scholar
Stott, R. (2012). Darwin's ghosts: in search of the first evolutionists. London: Bloomsbury.Google Scholar
Stowe, M. K. (1988). Chemical mimicry. In Chemical mediation of coevolution, ed. Spencer, K. C., pp. 513–80. San Diego, CA: Academic Press.Google Scholar
Strasburg, J. L.et al. (2012). What can patterns of differentiation across plant genomes tell us about adaptation and speciation?Philosophical Transactions of the Royal Society of London, 367, 364–73.Google ScholarPubMed
Strasburger, E. (1910). Chromosomenzahl. Flora, 100, 398–446.Google Scholar
Strauss, W. & Mainwaring, S. J. (1984). Air pollution. London: Edward Arnold.Google Scholar
Streisfeld, M. A, Young, W. N. & Sobel, J. M. (2013). Divergent selection drives genetic differentiation in an R2R3-MYB transcription factor that contributes to incipient speciation in Mimulus aurantiacus. PLOS Genetics, 10, DOI:10.1371/journal.pgen.1003385.Google Scholar
Strid, A. (1970). Studies in the Aegean flora. XVI. Biosystematics of the Nigella arvensis complex. With special reference to the problem of non-adaptive radiation. Opera Botanica, No. 28. Lund: Gleerup.Google Scholar
Stuart, A. (1984). The ideas of sampling. High Wycombe: Charles Griffin.Google Scholar
Stuart, S. N.et al. (2010). The barometer of life. Science, 328, 177.CrossRefGoogle ScholarPubMed
Stuessy, T. F. (1990). Plant taxonomy: the systematic evaluation of comparative data. New York: Columbia University Press.Google Scholar
Stuessy, T. F. (2004). A transitional-combinational theory of the origin of the angiosperms. Taxon, 53, 3–16.CrossRefGoogle Scholar
Stuessy, T. F. (2009). Plant taxonomy: the systematic evaluation of comparative data, edn. New York: Columbia University Press.Google Scholar
Stuessy, T. F. (2010). Paraphyly and the origin and classification of angiosperms. Taxon, 59, 689–93.Google Scholar
Sturtevant, A. H. (1965). A history of genetics. New York: Harper Row.Google Scholar
Suda, J.et al. (2007). Complex distribution patterns of di-, tetra-, and hexaploid cytotypes in the European high mountain plant Senecio carniolicus (Asteraceae). American Journal of Botany, 94, 1391–1401.CrossRefGoogle Scholar
Sugii, N. & Lamoureux, C. (2004). Tissue culture as a conservation method: an empirical view from Hawaii. In Ex situ plant conservation, ed. Guerrant, E. O. Jr, Havens, K. & Maunder, M., pp. 189–205. Washington DC, Covelo, CA, & London: Island Press.Google Scholar
Sultan, S. E. (1987). Evolutionary implications of phenotypic plasticity in plants. Evolutionary Biology, 21, 127–78.Google Scholar
Sutherland, W. J. & Hill, D. A. (eds.) (1995). Managing habitats for conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Sutherland, W. J.et al. (2004). The need for evidence-based conservation. Trends in Ecology & Evolution, 19, 305–8.CrossRefGoogle ScholarPubMed
Sutton, W. S. (1902). On the morphology of the chromosome group in Brachystola magna. Biological Bulletin. Marine Biological Laboratory, Woods Hole, Mass., 4, 24–39.CrossRefGoogle Scholar
Sutton, W. S. (1903). The chromosomes in heredity. Biological Bulletin, Marine Biological Laboratory, Woods Hole, Mass., 4, 231–48.CrossRefGoogle Scholar
Swensen, S. M.et al. (1995). Genetic analysis of the endangered island endemic Malacothamnus fasciculatus (Nutt.) Green var. nesioticus (Rob.) Kearn (Malvaceae). Conservation Biology, 9, 404–15.CrossRefGoogle Scholar
Swenson, U., Hill, R. S. & McLoughlin, S. (2001). Biogeography of Nothofagus supports the sequence of Gondwana break-up. Taxon, 50, 1025–41.CrossRefGoogle Scholar
Swofford, D. L. & Olsen, G. J. (1990). Phylogenetic reconstruction. In Molecular systematics, ed. Hillis, D. K. & Moritz, C., pp. 411–501. Sunderland, MA: Sinauer.Google Scholar
Sybenga, J. (1996). Aneuploid and other cytological tester sets in Rye. Euphytica, 89, 143–51.CrossRefGoogle Scholar
Sytsma, K. J. & Schaal, B. A. (1985). Phylogenetics of the Lisianthius skinneri (Gentianaceae) species complex in Panama utilizing DNA restriction fragment analysis. Evolution, 39, 594–608.CrossRefGoogle ScholarPubMed
Syvanen, M. (2012). Evolutionary implications of horizontal gene transfer. Annual Review of Genetics, 46, 341–58.CrossRefGoogle ScholarPubMed
Szadkowski, E.et al. (2010). The first meiosis of resynthesized Brassica napus, a genome blender. New Phytologist, 186, 102–12.CrossRefGoogle ScholarPubMed
Täckholm, G. (1922). Zytologische studien über die Gattung Rosa. Acta Horti Bergiani, 7, 97–381.Google Scholar
Taggart, J. B., McNally, S. F. & Sharp, P. M. (1990). Genetic variability and differentiation among founder populations of the Pitcher Plant (Sarracenia purpurea L.) in Ireland. Heredity, 64, 177–83.CrossRefGoogle Scholar
Tahara, M. (1915). Cytological studies on Chrysanthemum. Botanical Magazine (Tokyo), 29, 48–50.Google Scholar
Takayama, S. I. & Isogai, A. (2005). Self-incompatibility in plants. Annual Review of Plant Biology, 56, 467–8.CrossRefGoogle ScholarPubMed
Takhtajan, A. (1969). Flowering plants – origin and dispersal. Authorised translation from the Russian by Jeffrey, C.. Edinburgh: Oliver & Boyd.Google Scholar
Takhtajan, A. L. (1987). Sistema Magnoliofitov (Systema Magnoliophytorum). Acad. Sciences, USSR, Nauka, Leningrad. [In Russian: see Taxon, , 37, 422–4 (1988) for details.]Google Scholar
Talianova, M. & Janousek, B. (2011). What can we learn from tobacco and other Solanaceae about horizontal DNA transfer?American Journal of Botany, 98, 1231–42.CrossRefGoogle ScholarPubMed
Tang, C.et al. (2007). The evolution of selfing in Arabidopsis thaliana. Science, 317, 1070–2.CrossRefGoogle ScholarPubMed
Tang, H.et al. (2008a). Synteny and collinearity in plant genomes. Science, 320, 486–8.CrossRefGoogle ScholarPubMed
Tang, H.et al. (2008b). Unraveling ancient hexaploidy through multiply-aligned angiosperm gene maps. Genome Research, 18, 1944–54.CrossRefGoogle ScholarPubMed
Tansley, A. G. (1945). Our heritage of wild nature: a plea for organized nature conservation. Cambridge: Cambridge University Press.Google Scholar
Tardif, B. & Morisset, P. (1991). Chromosomal C–band variation in Allium schoenoprasum (Lilaceae) in Eastern–North America. Plant Systematics & Evolution, 174, 125–37.CrossRefGoogle Scholar
Tate, J.Soltis, D. E. & Soltis, P. S. (2005). Polyploidy in plants. In The evolution of the genome, ed. Gregory, T. R., pp. 371–426. London: Elsevier Academic Press.Google Scholar
Tate, J. A.et al. (2006). Evolution and expression of homeologous loci in Tragopogon miscellus (Asteraceae), a recent and reciprocally formed allopolyploid. Genetics, 173, 1599–1611.CrossRefGoogle Scholar
Tate, J. Aet al. (2009). On the road to diploidization? Homoelog loss in independently formed populations of the allopolyploid Tragopogon miscellus (Asteraceae). BMC Plant Biology, 9, 80, DOI:10.1186/1471-2229-9-80CrossRefGoogle Scholar
Tautz, D.et al. (2003). A plea for DNA taxonomy. Trends in Ecology and Evolution, 18, 70–4.CrossRefGoogle Scholar
Tayalé, A. & Parisod, C. (2013). Natural pathways to polyploidy in plants and consequences for genome reorganization. Cytogenetic and Genome Research, DOI:10.1159/000351318.CrossRef
Taylor, G. E. Jr & Murdy, W. H. (1975). Population differentiation of an annual plant species, Geranium carolinianum in response to sulfur dioxide. Botanical Gazette, 136, 212–15.CrossRefGoogle Scholar
Taylor, G. E., Pitelka, L. F. & Clegg, M. T. (1991). Ecological genetics and air pollution. New York: Springer-Verlag.CrossRefGoogle Scholar
Taylor, H. R. & Harris, W. E. (2012). An emergent science on the brink of irrelevance: a review of the past 8 years of DNA barcoding. Molecular Ecology Resources, 12, 377–88.CrossRefGoogle Scholar
Taylor, J. S. & Raes, J. (2005). Small-scale gene duplications. In The evolution of the genome, ed. Gregory, T. R., pp. 289–327. London: Elsevier Academic Press.Google Scholar
Taylor, K. & Markham, B. (1978). Ranunculus ficaria L. Biological flora of the British Isles. Journal of Ecology, 66, 1011–31.Google Scholar
Telwala, Y.et al. (2013). Climate-induced elevation range shifts and increase in plant species richness in a Himalayan biodiversity epicenter. PLOS ONE, 8, e57103.CrossRefGoogle Scholar
Theaker, A. J. & Briggs, D. (1993). Genecological studies of groundsel (Senecio vulgaris L.). IV. Rate of development in plants from different habitat types. New Phytologist, 123, 185–94.Google Scholar
Theissen, G. (2001). Development of floral organ identity: stories from the MADS house. Current Opinion in Plant Biology, 4, 75–85.CrossRefGoogle ScholarPubMed
Theissen, G. (2006). The proper place of hopeful monsters in evolutionary biology. Theory in Biosciences, 124, 349–69.Google ScholarPubMed
Theissen, G. & Melzer, R. (2007). Molecular mechanisms underlying origin and diversification of the angiosperm flower. Annals of Botany, 100, 603–19.CrossRefGoogle ScholarPubMed
Theissen, G.et al. (2002). How land plants learned their floral ABCs: the role of MADS-box genes in the evolutionary origin of flowers. In Developmental genetics and plant evolution, ed. Cronk, Q. C. B., Bateman, R. M. & Hawkins, J. A., pp. 173–205. London: Taylor & Francis.Google Scholar
Theobald, D. L. (2010). A formal test of the theory of universal common ancestry. Nature, 465, 219–22.CrossRefGoogle ScholarPubMed
Theodoridis, S.et al. (2013). Divergent and narrower climatic niches characterize polyploid species of European primroses in Primula sect. Aleuritia. Journal of Biogeography, 40, 1278–89.CrossRefGoogle Scholar
Thiele, K. & Yeates, D. (2002). Tension arises from duality at the heart of taxonomy. Nature, 419, 337.CrossRefGoogle ScholarPubMed
Thiselton-Dyer, W. T. (1895a). Variation and specific stability. Nature, 51, 459–61.CrossRefGoogle Scholar
Thiselton-Dyer, W. T. (1895b). Origin of the cultivated Cineraria. Nature, 52, 3–4, 78–9, 128–9.CrossRefGoogle Scholar
Thoday, J. M. (1972). Disruptive selection. Proceedings of the Royal Society of London, B, 182, 109–43.CrossRefGoogle ScholarPubMed
Thomas, C. D.et al. (2004). Extinction risk from climate change. Nature, 427, 145–8.CrossRefGoogle ScholarPubMed
Thomas, D. A. & Barber, H. N. (1974). Studies of leaf characteristics of a cline of Eucalyptus urnigera from Mount Wellington, Tasmania. II. Reflection, transmission and absorption of radiation. Australian Journal of Botany, 22, 701–7.Google Scholar
Thomas, R. B.et al. (2013). Evidence of recovery of Juniperus virginiana trees from sulfur pollution after the Clean Air Act. Proceedings of the National Academy of Sciences, USA, 110, 15319–24.CrossRefGoogle ScholarPubMed
Thomas, S. G. & Franklin-Tong, V. E. (2004). Self-incompatibility triggers programmed cell death in Papaver pollen. Nature, 429, 305–9.CrossRefGoogle ScholarPubMed
Thompson, D. Q.et al. (1987). Spread, impact, and control of Purple Loosestrife (Lythrum salicaria) in North American wetlands. US Fish and Wildlife Service. Jamestown, ND: Northern Prairie Wildlife Research Center Online.Google Scholar
Thompson, J. D. & Lumaret, R. (1992). The evolutionary dynamics of polyploid plants: origins, establishment and persistence. Trends in Ecology and Evolution, 7, 302–6.CrossRefGoogle ScholarPubMed
Thompson, J. N. (1998). Rapid evolution as an ecological process. Trends in Ecology and Evolution, 13, 329–32.CrossRefGoogle ScholarPubMed
Thompson, S. L. & Whitton, J. (2006). Patterns of recurrent evolution and geographic parthenogenesis within apomictic Easter daisies (Townsendia hookeri). Molecular Ecology, 15, 3389–400.CrossRefGoogle Scholar
Thompson, S. L.et al. (2008). Cryptic sex within male-sterile polyploid populations of the Easter Daisy, Townsendia hookeri. International Journal of Plant Science, 169, 183–93.CrossRefGoogle Scholar
Thomson, J. D., Herre, E. A., Hamrick, J. L. & Stone, J. L. (1991). Genetic mosaics in Strangler Fig trees: implications for tropical conservation. Science, 254, 1214–16.CrossRefGoogle ScholarPubMed
Thrall, P. H.et al. (1998). Metapopulation collapse: the consequences of limited gene flow in spatially structured populations. In Modeling spatiotemporal dynamics in ecology, ed. Bascompte, J. & Sole, R. V., pp. 83–104. Berlin: Springer-Verlag.Google Scholar
Throop, W. (2004). A response to the article (Hobbs, 2004) ‘Restoration Ecology: the challenge of social values and expectation’. Frontiers in Ecology and the Environment, 2, 47–8.Google Scholar
Thuiller, W. (2003). BIOMOD – optimizing predictions of species distributions and projecting potential future shifts under global change. Global Change Biology, 9, 1353–62.CrossRefGoogle Scholar
Thuiller, W.et al. (2008). Predicting global change impacts on plant species’ distributions: future challenges. Perspectives in Plant Ecology, Evolution and Systematics, 9, 137–52.CrossRefGoogle Scholar
Thurston, J. M., Dyke, G. V. & Williams, E. D. (1976). The Park Grass experiment on the effect of fertilisers and liming on the botanical composition of permanent grassland and on the yield of hay. Harpenden: Rothamsted Experimental Station.Google Scholar
Tian, D.et al. (2002). Signature of balancing selection in Arabidopsis. Proceedings of the National Academy of Sciences, USA, 99, 11525–30.CrossRefGoogle ScholarPubMed
Till, I. (1987). Variability of expression of cyanogenesis in White Clover (Trifolium repens L.). Heredity, 59, 265–71.CrossRefGoogle Scholar
Tindall, K. R. & Kunkel, T. A. (1988). Fidelity of DNA synthesis by the Thermus aquaticus DNA polymerase. Biochemistry, 27, 6008–13.CrossRefGoogle ScholarPubMed
Tischler, G. (1950). Die Chromosomenzahlen der Gefässpflanzen Mitteleuropas. 'S-Gravenhage: Junk.CrossRefGoogle Scholar
Tobgy, H. A. (1943). A cytological study of Crepis fuliginosa, C. neglecta and their F1 hybrid, and its bearing on the mechanism of phylogenetic reduction in chromosome number. Journal of Genetics, 45, 67–111.CrossRefGoogle Scholar
Todhunter, I. (1873). The conflict of studies and other essays. London: Macmillan.Google Scholar
Toomajian, C.et al. (2006). A nonparametric test reveals selection for rapid flowering in the Arabidopsis genome. PLoS Biology, 4, 732–8.CrossRefGoogle ScholarPubMed
Tower, W. L. (1902). Variation in the ray-flowers of Chrysanthemum leucanthemum L. at Yellow Springs, Green County, O, with remarks upon the determination of the modes. Biometrika, 1, 309–15.CrossRefGoogle Scholar
Townsend Peterson, A. (2011). Ecological niche conservatism: a time-structured review of evidence. Journal of Biogeography, 38, 817–27.Google Scholar
Traill, L. W., Bradshaw, C. J. A. & Brook, B. W. (2007). Minimum viable population size: a meta-analysis of 30 years of published estimates. Biological Conservation, 139, 159–66.CrossRefGoogle Scholar
Traill, L. W.et al. (2010). Pragmatic population viability targets in a rapidly changing world. Biological Conservation, 143, 28–34.CrossRefGoogle Scholar
Trakhtenbrot, A.et al. (2005). The importance of long-distance dispersal in biodiversity conservation. Diversity & Distributions, 11, 173–81.CrossRefGoogle Scholar
Tralau, H. (1968). Evolutionary trends in the genus Ginkgo. Lethaia, 1, 63–101.CrossRefGoogle Scholar
Tremetsberger, K.et al. (2002). Infraspecific genetic variation in Biscutella laevigata (Brassicaceae): new focus on Irene Manton's hypothesis. Plant Systematics and Evolution, 233, 163–81.CrossRefGoogle Scholar
Trevisan, L. (1988). Angiospermous pollen (monosulcate-trichotomosulcate phase) from very early Lower Cretaceous of Southern Tuscany (Italy): some aspects. In Proceedings of the 7th International Palynological Congress, Brisbane, Australia, Abstr. 165. Amsterdam: Elsevier.Google Scholar
Trewick, S. A., Morgan-Richards, M. & Chapman, H. M. (2004). Chloroplast DNA diversity of Hieracium pilosella (Asteraceae) introduced to New Zealand: reticulation, hybridization and invasion. American Journal of Botany, 91, 73–85.CrossRefGoogle ScholarPubMed
Trusty, J. L.et al. (2005). Molecular phylogenetics of the Macaronesian-endemic genus Bystropogon (Lamiaceae): palaeo-islands, ecological shifts and interisland colonizations. Molecular Ecology, 14, 1177–89.CrossRefGoogle ScholarPubMed
Tsuchimatsu, T.et al. (2010). Evolution of self-compatibility in Arabidopsis by a mutation in the male specificity gene. Nature, 464, 1342–6.CrossRefGoogle ScholarPubMed
Tubeuf, K. F.. (1923). Monographie der Mistel. Munich & Berlin: Oldenbourg.Google Scholar
Tucker, M. R. & Koltuno, A. M. G. (2009). Sexual and asexual (apomictic) seed development in flowering plants: molecular, morphological and evolutionary relationships. Functional Plant Biology, 36, 490–504.CrossRefGoogle Scholar
Turesson, G. (1922a). The species and variety as ecological units. Hereditas, 3, 100–13.Google Scholar
Turesson, G. (1922b). The genotypical response of the plant species to the habitat. Hereditas, 3, 211–350.Google Scholar
Turesson, G. (1925). The plant species in relation to habitat and climate. Hereditas, 6, 147–236.Google Scholar
Turesson, G. (1927a). Erbliche Transpirationsdifferenzen zwischen Ökotypen derselben Pflanzen Art. Hereditas, 11, 193–206.Google Scholar
Turesson, G. (1927b). Untersuchungen über Grenzplasmolyse und Saugkraftwerte in verschiedenen Ökotypen derselben Art. Jahrbücher für wissenschaftliche Botanik, 66, 723–47.Google Scholar
Turesson, G. (1930). The selective effect of climate upon the plant species. Hereditas, 14, 99–152.Google Scholar
Turesson, G. (1943). Variation in the apomictic microspecies of Alchemilla vulgaris L. Botaniska Notiser, 1943, 413–27.
Turesson, G. (1961). Habitat modifications in some widespread plant species. Botaniska Notiser, 114, 435–52.Google Scholar
Turkington, R. (1989). The growth, distribution and neighbour relationships of Trifolium repens in a permanent pasture. V. The co-evolution of competitors. Journal of Ecology, 77, 717–33.Google Scholar
Turkington, R. A. & Harper, J. L. (1979a). The growth, distribution and neighbour relationships of Trifolium repens in a permanent pasture. I. Ordination pattern and contact. Journal of Ecology, 67, 201–18.Google Scholar
Turkington, R. A. & Harper, J. L. (1979b). The growth, distribution and neighbour relationships of Trifolium repens in a permanent pasture. II. Fine scale biotic differentiations. Journal of Ecology, 67, 245–54.Google Scholar
Turnbull, C. (2011). Long-distance regulation of flowering time. Journal of Experimental Botany, 62, 4399–4413.CrossRefGoogle ScholarPubMed
Turner, I. M., Tan, H. T. W., Wee, Y. C., Ibrahim, A. B., Chew, P. T. & Corlett, R. T. (1994). A study of plant species extinction in Singapore: lessons for the conservation of tropical biodiversity. Conservation Biology, 8, 705–12.CrossRefGoogle Scholar
Turner, R. K. & Daily, G. C. (2008). The ecosystems services framework and natural capital conservation. Environmental and Resource Economics, 39, 25–35.CrossRefGoogle Scholar
Turner, T. L.et al. (2008). Population resequencing reveals local adaptation of Arabidopsis lyrata to serpentine soils. Nature Genetics, 42, 260–3.Google Scholar
Turrill, W. B. (1938). The expansion of taxonomy with special reference to Spermatophyta. Biological Reviews, 13, 342–73.CrossRefGoogle Scholar
Turrill, W. B. (1940). Experimental and synthetic plant taxonomy. In The new systematics, ed. Huxley, J. S., pp. 47–71. Oxford: Clarendon Press.Google Scholar
Tutin, T. G. (1957). A contribution to the experimental taxonomy of Poa annua L. Watsonia, 4, 110.
Tutin, T. G., Heywood, V. H., Burges, N. A., Moore, D. M., Valentine, D. H., Walters, S. M. & Webb, D. A. (1964–80). Flora Europaea. Cambridge: Cambridge University Press.Google Scholar
Twyford, A. D. & Ennos, R. A. (2012 ).Next-generation hybridization and introgression. Heredity, 108, 179–89.CrossRefGoogle ScholarPubMed
Udall, J. A., Quijada, P. A. & Osborn, T. C. (2005). Detection of chromosomal rearrangements derived from homoeologous recombination in four mapping populations of Brassica napus L. Genetics, 169, 967–79.
Uesugi, R.et al. (2007). Restoration of genetic diversity from soil seed banks in a threatened aquatic plant, Nymphoides peltata. Conservation Genetics, 8, 111–21.Google Scholar
Uhl, C. H. (1978). Chromosomes of Mexican Sedum, section Pachysedum. Rhodora, 80, 491–512.Google Scholar
Uphof, J. C. Th. (1938). Cleistogamic flowers. The Botanical Review, 4, 21–49.CrossRefGoogle Scholar
Valente, L. P., Silva, M. C. C. & Jansen, L. E. T. (2012). Temporal control of epigenetic centromere specification. Chromosome Research, 20, 481–92.CrossRefGoogle ScholarPubMed
Valentine, D. H. (1941). Variation in Viola riviniana Rchb. New Phytologist, 40, 189–209.CrossRefGoogle Scholar
Valentine, D. H. (1956). Studies in British Primulas. V. The inheritance of seed incompatibility. New Phytologist, 55, 305–18.CrossRefGoogle Scholar
Valentine, D. H. (1975). Primula. In Hybridization and the flora of the British Isles, ed. Stace, C. A., pp. 346–8. London & New York: Academic Press.Google Scholar
Vallejo-Marín, M., Dorken, M. E. & Barrett, S. C. H. (2010). The ecological and evolutionary consequences of clonality for plant mating. Annual Reviews of Ecology, Evolution and Systematics, 41, 193–213.CrossRefGoogle Scholar
Van de Peer, Y. (2003). Phylogeny inference based on distance methods. In The phylogenetic handbook: a practical guide to DNA and protein phylogeny. ed. Salemi, M. & Vandamme, A.-M., pp. 101–19. Cambridge: Cambridge University PressGoogle Scholar
van der Pilj, L. & Dodson, C. H. (1966). Orchid flowers: their pollination and evolution. Coral Gables, FL: University of Miami Press.Google Scholar
Van der Putten, W. H., Macel, M. & Visser, M. E. (2010). Predicting species distribution and abundance responses to climate change: why it is essential to include biotic interactions across trophic levels. Philosophical Transactions of the Royal Society, B, 365, 2025–34.CrossRefGoogle ScholarPubMed
van Dijk, H. (2004). Gene exchange between wild and crop in Beta vulgaris: how easy is hybridization and what will happen in later generations? In Introgression from genetically modified plants into wild relatives, ed. den Nijs, H. C. M., Bartsch, D. & Sweet, J., pp.53–61. Wallingford, UK: CAB International.Google Scholar
Van Dijk, P. J. (2003). Ecological and evolutionary opportunities of apomixis: insights from Taraxacum and Chondrilla. Philosophical Transactions of the Royal Society of London, B, 358, 1113–21.CrossRefGoogle ScholarPubMed
Van Dijk, P. (2009). Apomixis: basics for non-botanists. In Lost sex: the evolutionary biology of parthenogenesis, ed. Schön, I., Martens, K. & Dijk, P. J. van, pp. 47–62. Berlin: Springer Publications.Google Scholar
van Dijk, P. & Bijlsma, R. (1994). Simulation of flowering time displacement between two cytotypes that form inviable hybrids. Heredity, 72, 522–35.CrossRefGoogle Scholar
Van Dijk, P. J. & Vijverberg, K. (2005). The significance of apomixis in the evolution of the angiosperms: a reappraisal. In Plant species-level systematics. New perspectives on pattern and process, ed. Bakker, F. Y. T., Chatrou, L. W., Gravendeel, B. & Pelser, P. B., pp. 101–16. Regnum Vegetabile, 143. Ruggell, Liechtenstein: A. R. G. Gantner Verlag.
van Groenendael, J. M. (1986). Life history characteristics of two ecotypes of Plantago lanceolata. Acta Botanica Neerlandica, 35, 71–86.CrossRefGoogle Scholar
van Teuren, R.et al. (1991). The significance of genetic erosion in the process of extinction. I. Genetic differentiation in Salvia pratensis and Scabiosa columbaria in relation to population size. Heredity, 66, 181–9.Google Scholar
van Teuren, R.et al. (1993). The significance of genetic erosion in the process of extinction. IV. Inbreeding depression and heterosis effects caused by selfing and outcrossing in Scabiosa columbaria. Evolution, 47, 1669–80.Google Scholar
van Tienderen, P. H. & van der Toorn, J. (1991a). Genetic differentiation between populations of Plantago lanceolata. I. Local adaptation in three contrasting habitats. Journal of Ecology, 79, 27–42.CrossRefGoogle Scholar
Van Tienderen, P. H. & van der Toorn, J. (1991b). Genetic differentiation between populations of Plantago lanceolata. II. Phenotypic selection in a transplant experiment in three contrasting habitats. Journal of Ecology, 79, 43–59.CrossRefGoogle Scholar
Van Valen, L. (1976). Ecological species, multispecies, and oaks. Taxon, 25, 223–39.Google Scholar
van Wyk, A. E. (2007). The end justifies the means. Taxon, 56, 645–8.CrossRefGoogle Scholar
Vandamme, A.-M. (2003). Basic concepts of molecular evolution. In The phylogenetic handbook: a practical guide to DNA and protein phylogeny, ed. Salemi, M. & Vandamme, A.-M., pp. 1–23. Cambridge: Cambridge University Press.Google Scholar
Vangronsveld, J.et al. (2009). Phytoremediation of contaminated soils and groundwater: lessons from the field. Environmental Science and Pollution Research, 16, 765–94.CrossRefGoogle ScholarPubMed
Vanzolini, P. E. & Williams, E. E. (1970). South American anoles: the geographic differentiation and evolution of the Anolis chrysolepis species group (Sauria, Iguanidae). Arquivos de Zoologia (São Paulo), 19, 1–298.Google Scholar
Vargas, P.et al. (1999). Polyploid speciation in Hedera (Araliaceae): phylogenetic and biogeographic insights based on chromosome counts and ITS sequences. Plant Systematics and Evolution, 219, 165–79.CrossRefGoogle Scholar
Vasek, F. C. (1980). Creosote Bush: long-lived clones in the Mojave Desert. American Journal of Botany, 67, 246–55.CrossRefGoogle Scholar
Vashisht, D.et al. (2011). Natural variation of submergence tolerance amongst Arabidopsis thaliana accessions. New Phytologist, 190, 299–310.CrossRefGoogle Scholar
Vaughan, D. A., Balazs, E. & Heslop-Harrison, J. S. (2007). From crop domestication to super-domestication. Annals of Botany, 100, 893–901.CrossRefGoogle ScholarPubMed
Vaughton, G., Ramsey, M. & Johnson, S. D. (2011). Pollination and late-acting self-incompatibility in Cyrtanthus breviflorus (Amaryllidaceae): implications for seed production. Annals of Botany, 106, 547–55.Google Scholar
Vekeman, X. & Lefèbvre, C. (1997). On the evolution of heavy metal tolerant populations in Armeria maritima: evidence from allozyme variation and reproductive barriers. Journal of Evolutionary Biology, 10, 175–91.Google Scholar
Vekemans, X. (2010). What's good for you may be good for me: evidence for adaptive introgression of multiple traits in wild sunflower. New Phytologist, 187, 6–9.CrossRefGoogle ScholarPubMed
Vekemans, X. & Hardy, O. J. (2004). New insights from fine-scale spatial genetic structure analyses in plant populations. Molecular Ecology, 13, 921–35.CrossRefGoogle ScholarPubMed
Vellinga, E. C., Wolfe, B. E. & Pringle, A. (2009).Global patterns of ectomycorrhiza introductions. New Phytologist, 181, 960–73.CrossRefGoogle Scholar
Venable, D. L. & Levin, D. A. (1985). Ecology of achene dimorphism in Heterotheca latifolia. I. Achene structure, germination and dispersal. Journal of Ecology, 73, 133–45.CrossRefGoogle Scholar
Vergeer, P., Wagemaker, N. & Ouborg, N. J. (2012). Evidence for an epigenetic role in inbreeding depression. Biology Letters, 8, 798–801.CrossRefGoogle ScholarPubMed
Verhoeven, K. J. F.et al. (2010). Stress-induced DNA methylation changes and their heritability in asexual dandelions. New Phytologist, 185, 1108–18.CrossRefGoogle ScholarPubMed
Vernon, H. M. (1903). Variation in animals and plants. London: Kegan Paul.CrossRefGoogle Scholar
Verschaffelt, E. (1899). Galton's regression to mediocrity bij ongeslachtelijke verplanting. In Livre Jubilaire dédié à Charles wan Bambeke, pp. 1–5. Brussels: Lamerton.Google Scholar
Vesteg, M., Vacula, R. & Krajčovič, J. (2009). On the origin of chloroplasts, import mechanisms and chloroplast-targeted proteins, and loss of photosynthetic ability – review. Folia Microbiologica, 54, 303–21.CrossRefGoogle ScholarPubMed
Vickery, R. K. (1964). Barriers to gene exchange between members of the Mimulus guttatus complex (Scrophulariaceae). Evolution, 18, 52–69.CrossRefGoogle Scholar
Vietmeyer, N. (1995). Applying biodiversity. Journal of the Federation of American Scientists, 48, 1–8.Google Scholar
Viette, M., Tettamanti, C. & Saucy, F. (2000). Preference for acyanogenic white clover (Trifolium repens) in the vole Arvicola terrestris. II. Generalization and further investigations. Journal of Chemical Ecology, 26, 101–22.CrossRefGoogle Scholar
Viktora, M., Savidge, R. A. & Rajora, O. P. (2011). Clonal and nonclonal genetic structure of subarctic Black Spruce (Picea mariana) populations in Yukon Territory. Botany-Botanique, 89, 133–40.CrossRefGoogle Scholar
Vila-Aiub, M. M., Neve, P. & Powles, S. B. (2009). Fitness costs associated with evolved herbicide resistance alleles in plants. New Phytologist, 184, 751–67.CrossRefGoogle ScholarPubMed
Viosca, P. Jr (1935). The irises of southeastern Louisiana: a taxonomic and ecological interpretation. Bulletin of the American Iris Society, 57, 3–56.Google Scholar
Vitousek, P. M.et al. (1997). Human domination of the Earth's ecosystems. Science, 277, 494–9.CrossRefGoogle Scholar
Vittoz, P.et al. (2013). Climate change impacts on biodiversity in Switzerland. Journal for Nature Conservation, 21, 154–62.CrossRefGoogle Scholar
Volis, S. & Blecher, M. (2010). Quasi in situ: a bridge between ex situ and in situ conservation of plants. Biodiversity and Conservation, 19, 2441–54.CrossRefGoogle Scholar
von der Lippe, M.& Kowarik, I. (2007). Crop spillage along roads: a factor of uncertainty in the containment of GMO. Ecography, 30, 483–90.CrossRefGoogle Scholar
Vorontsova, M. S. & Simon, B. K. (2012). Updating classifications to reflect monophyly: 10 to 20 percent of species names change in Poaceae. Taxon, 61, 735–46.Google Scholar
Vorzimmer, P. J. (1972). Charles Darwin: the years of controversy. London: University of London Press.Google Scholar
Vouillamoz, J., Maigre, D. & Meredith, C. P. (2003). Microsatellite analysis of ancient alpine grape cultivars: Pedigree reconstruction of Vitis vinifera L. ‘Cornalin du Valais’. Theoretical and Applied Genetics, 107, 448–54.CrossRefGoogle ScholarPubMed
Vreysen, M. J. B., Robinson, A. S. & Hendrichs, J. (2007). The Mountain Pine Beetle Dendroctonus ponderosae in western North America: potential for area-wide integrated management. 2007. In Area-wide control of insect pests from research to field implementation, ed. Vreysen, M. J. B., Robinson, A. S. and Hendrichs, J., pp. 297–307. New York: Springer.CrossRefGoogle Scholar
Waddington, C. H. (1966). Mendel and evolution. In G. Mendel Memorial Symposium, 1865–1965, ed. Sosna, M., pp. 145–50. Prague: Academia Publishing House of the Czechoslovak Academy of Sciences.Google Scholar
Wägele, et al. (2011). The taxonomist – an endangered race. A practical proposal for its survival. Frontiers in Zoology, 8, 25–30.CrossRefGoogle Scholar
Wagenius, S. et al. (2012). Seedling recruitment in the long-lived perennial, Echinacea angustifolia: a 10-year experiment. Restoration Ecology, 20, 352–9.CrossRefGoogle Scholar
Wagner, M. (1868). Die Darwin'sche Theorie und das Migrationgesetz der Organismen. Leipzig: Duncker & Humblot.Google Scholar
Wakeley, J. (2008). Coalescent theory: an introduction. Greenwood Village, CO: Roberts & Company Publishers.Google Scholar
Waldron, L. R. (1912). Hardiness in successive Alfalfa generations. The American Naturalist, 46, 463–9.CrossRefGoogle Scholar
Waller, D. M. (1979).The relative costs of self- and cross-fertilized seeds in Impatiens capensis (Balsaminaceae). American Journal of Botany, 66, 313–20.CrossRefGoogle Scholar
Waller, D. M. (1984). Differences in fitness between seedlings derived from cleistogamous and chasogamous flowers in Impatiens capensis. Evolution, 38, 427–40.CrossRefGoogle ScholarPubMed
Waller, D. M., O'Malley, D. M. & Gawler, S. C. (1987). Genetic variation in the extreme endemic Pedicularis furbishiae (Scrophulariaceae). Conservation Biology, 1, 335–40.CrossRefGoogle Scholar
Walley, K. A., Khan, M. S. I. & Bradshaw, A. D. (1974). The potential for evolution of heavy metal tolerance in plants. I. Copper and zinc tolerance in Agrostis tenuis. Heredity, 32, 309–19.CrossRefGoogle Scholar
Wallis, G. P. & Trewick, S. A. (2009). New Zealand phylogeography: evolution on a small continent. Molecular Ecology, 18, 3548–580.CrossRefGoogle ScholarPubMed
Walsh, B. (2008). Using molecular markers for detecting domestication, improvement, and adaptation genes. Euphytica, 16, 1–17.Google Scholar
Walsh, N. E.et al. (1997). Experimental manipulations of snow-depth: effects on nutrient content of caribou forage. Global Change Biology, 3, 158–164.CrossRefGoogle Scholar
Walter, K. S. & Gillett, H. J. (1997) IUCN red list of threatened plants, compiled by the World Conservation Monitoring Centre. Gland, Switzerland, & Cambridge: IUCN – The World Conservation Union.Google Scholar
Walters, S. M. (1961). The shaping of angiosperm taxonomy. New Phytologist, 60, 74–84.CrossRefGoogle Scholar
Walters, S. M. (1962). Generic and specific concepts in the European flora. Preslia, 34, 207–26.Google Scholar
Walters, S. M. (1970). Dwarf variants of Alchemilla L. Fragmenta Floristica et Geobotanica, 16, 91–8.
Walters, S. M. (1972). Endemism in the genus Alchemilla in Europe. In Taxonomy, phytogeography and evolution, ed. Valentine, D. H., pp. 301–5. London & New York: Academic Press.Google Scholar
Walters, S. M. (1979). Progress in biological conservation in Cambridge. In Landscape towards 2000: conservation or desolation, ed. Smith, D., pp. 56–8. London: Landscape Institute.Google Scholar
Walters, S. M. (1986a). Alchemilla: a challenge to biosystematists. Acta Universitatus Upsaliensis, Symbolae Botanicae Upsalienses, XXVII, 193–8.Google Scholar
Walters, S. M. (1986b). The name of the rose: a review of ideas on the European bias in angiosperm classification. New Phytologist, 104, 527–46.CrossRefGoogle Scholar
Walters, S. M. (1989a). Obituary of John Scott Lennox Gilmour. Plant Systematics & Evolution, 167, 93–5.CrossRefGoogle Scholar
Walters, S. M. (1989b). Experimental and orthodox taxonomic categories and the deme terminology. Plant Systematics & Evolution, 167, 1–2.CrossRefGoogle Scholar
Walters, S. M. & Stow, E. A. (2001). Darwin's mentor: John Stevens Henslow, 1796–1861. Cambridge: Cambridge University Press.Google Scholar
Walther, G.-R. (2010). Community and ecosystem responses to recent climate change. Philosophical Transactions of the Royal Society, B, 365, 2019–24.CrossRefGoogle ScholarPubMed
Walther, G.-R.et al. (2002). Ecological responses to recent climate change. Nature, 416, 389–95.CrossRefGoogle ScholarPubMed
Walther-Hellwig, K. & Frankl, R. (2003). Foraging habitats and foraging distances of bumblebees, Bombus spp. (Hym., Apidae), in an agricultural landscape. Journal of Applied Entomology, 124, 299–306.Google Scholar
Wang, E.et al. (2008). Control of rice grain-filling and yield by a gene with a potential signature of domestication. Nature Genetics, 40, 1370–4.CrossRefGoogle ScholarPubMed
Wang, H.et al. (2005). The origin of the naked grains of Maize. Nature, 436, 714–19.CrossRefGoogle ScholarPubMed
Wang, W.-X.et al. (2007). MicroRNAs (miRNAs) and plant development. In Handbook of plant science, ed. Roberts, K., pp. 640–9. Chichester: Wiley.Google Scholar
Ward, D. B. (1974). The ‘ignorant man’ technique of sampling plant populations. Taxon, 23, 325–30.CrossRefGoogle Scholar
Ward, J.et al. (2000). Is atmospheric CO2 a selective agent on model C3 annuals?Oecologia, 123, 330–41.CrossRefGoogle ScholarPubMed
Warren, J. (2009). Extra petals in the buttercup (Ranunculus repens) provide a quick method to estimate the age of meadows. Annals of Botany, 104, 785–8.CrossRefGoogle ScholarPubMed
Warren, M. S.et al. (2001). Rapid responses of British butterflies to opposing forces of climate and habitat change. Nature, 414, 65–9.CrossRefGoogle ScholarPubMed
Warren, R.et al. (2013) Quantifying the benefit of early climate change mitigation in avoiding biodiversity lossNature Climate Change, 3, 678–82.CrossRefGoogle Scholar
Warwick, S. I. (1990a). Allozyme and life history variation in five northwardly colonizing North American weed species. Plant Systematics & Evolution, 169, 41–54.CrossRefGoogle Scholar
Warwick, S. I. (1990a). Genetic variation in weeds – with particular reference to Canadian agricultural weeds. In Biological approaches and evolutionary trends in plants, ed. Kawano, S., pp. 3–18. New York: Academic Press.Google Scholar
Warwick, S. I. (1991). Herbicide resistance in weedy plants: physiology and population biology. Annual Review of Ecology & Systematics, 22, 95–114.CrossRefGoogle Scholar
Warwick, S. I. & Al-Shehbaz, I. A. (2006). Brassicaceae: chromosome number index and database on CD-Rom. Plant Systematics and Evolution, 259, 237–48.Google Scholar
Warwick, S. I. & Briggs, D. (1978a). The genecology of lawn weeds. I. Population differentiation in Poa annua L. in a mosaic environment of bowling green lawns and flower beds. New Phytologist, 81, 711–23.Google Scholar
Warwick, S. I. & Briggs, D. (1978b). The genecology of lawn weeds. II. Evidence for disruptive selection in Poa annua L. in a mosaic environment of bowling green lawns and flower beds. New Phytologist, 81, 725–37.Google Scholar
Warwick, S. I. & Briggs, D. (1979). The genecology of lawn weeds. III. Cultivation experiments with Achillea millefolium L., Bellis perennis L., Plantago lanceolata L., Plantago major L. and Prunella vulgaris L. collected from lawns and contrasting grassland habitats. New Phytologist, 83, 509–36.CrossRefGoogle Scholar
Warwick, S. I. & Briggs, D. (1980a). The genecology of lawn weeds. IV. Adaptive significance of variation in Bellis perennis L. as revealed in a transplant experiment. New Phytologist, 85, 275–88.Google Scholar
Warwick, S. I. & Briggs, D. (1980b). The genecology of lawn weeds. V. The adaptive significance of different growth habit in lawn and roadside populations of Plantago major L. New Phytologist, 85, 289–300.
Warwick, S. I. & Briggs, D. (1980c). The genecology of lawn weeds. VI. The adaptive significance of variation in Achillea millefolium L. as investigated by transplant experiments. New Phytologist, 85, 451–60.Google Scholar
Warwick, S. I. & Gottlieb, L. D. (1985). Genetic divergence and geographic speciation in Layia (Compositae). Evolution, 39, 1236–41.CrossRefGoogle Scholar
Warwick, S. I., Phillips, D., Andrews, C. (1986). Rhizome depth: the critical factor in winter survival of Sorghum halepense (L.) Pers. (Johnson grass). Weed Research, 26, 381–7.CrossRefGoogle Scholar
Warwick, S. I., Thompson, B. K. & Black, L. D. (1987). Genetic variation in Canadian and European populations of the colonising weed species Apera spica-venti. New Phytologist, 106, 301–17.CrossRefGoogle Scholar
Warwick, S. I., Bain, J. F., Wheatcroft, R. & Thompson, B. K. (1989). Hybridization and introgression in Carduus nutans and C. acanthoides re-examined. Systematic Botany, 14, 476–94.CrossRefGoogle Scholar
Warwick, S. I., Beckie, H. J. & Hall, L. M. (2009). Gene flow, invasiveness, and ecological impact of gentically modified crops. Annals of the New York Academy of Sciences, 1168, 72–99.CrossRefGoogle Scholar
Watrud, L. S.et al. (2004) Evidence for landscape-level, pollen-mediated gene flow from genetically modified creeping bentgrass with CP4 EPSPS as a marker. Proceedings of the National Academy of Sciences, USA, 101, 14533–8.CrossRefGoogle ScholarPubMed
Watson, J. D. (1968). The double helix. London: Weidenfeld & Nicolson.Google Scholar
Watson, J. D. & Crick, F. H. C. (1953). A structure of deoxyribose nucleic acid. Nature, 171, 737–8.Google ScholarPubMed
Watson, J. D.et al. (2014). Molecular biology of the gene, edn. Boston: Pearson.Google ScholarPubMed
Watson, P. J. (1969). Evolution in closely adjacent plant populations. VI. An entomophilous species, Potentilla erecta, in two contrasting habitats. Heredity, 24, 407–22.CrossRefGoogle Scholar
Weart, S. (2011). Global warming: how skepticism became denial. Bulletin of the Atomic Scientists, 67, 41–50.CrossRefGoogle Scholar
Weber, B. H. & Depew, D. J. (2007). Darwinism, design, and complex systems dynamics. In Debating design. From Darwin to DNA, ed. Dembski, W.A & Ruse, M., pp. 173–90. Cambridge: Cambridge University Press.Google Scholar
Weber, E. & Schmidt, B. (1998). Latitudinal population differentiation in two species of Solidago (Asteraceae) introduced into Europe. American Journal of Botany, 85, 1110–21.CrossRefGoogle ScholarPubMed
Webster, S. D. (1988). Ranunculus penicillatus (Dumort.) Bab. in Great Britain and Ireland. Watsonia, 17, 1–22.Google Scholar
Weeden, N. F. & Gottlieb, L. D. (1979). Isolation of cytoplasmic enzymes from pollen. Plant Physiology, 66, 400–3.Google Scholar
Weeden, N. F. & Wendel, J. F. (1990) Genetics of plant isozymes. In Isoenzymes in plant biology, ed. Soltis, D. E & Soltis, P. S., pp. 46–72. London: Chapman & Hall.Google Scholar
Wegener, A. (1912). Die Herausbildung der Grossformen der Erdrinde (Kontinente und Ozeane), auf geophysikalischer Grundlage. Petermanns Geographische Mitteilungen, 63: 185–95, 253–6, 305–9. Presented at the annual meeting of the German Geological Society, Frankfurt am Main (6 January 1912).Google Scholar
Weigel, D. (2012). Natural variation in Arabidopsis: from molecular genetics to ecological genomics. Plant Physiology, 158, 2–22.CrossRefGoogle ScholarPubMed
Weigel, D. & Colot, V. (2012). Epialleles in plant evolution. Genome Biology, 13, 249–54.CrossRefGoogle ScholarPubMed
Weimark, H. (1945). Experimental taxonomy in Aethusa cynapium. Botaniska Notiser, 1945, 351–80.Google Scholar
Weinstein, A. (1977). How unknown was Mendel's paper?Journal of the History of Biology, 10, 341–64.CrossRefGoogle Scholar
Weir, J. & Ingram, R. (1980). Ray morphology and cytological investigations of Senecio cambrensis Rosser. New Phytologist, 86, 237–41.CrossRefGoogle Scholar
Weising, K.et al. (2005). DNA fingerprinting in plants. Principles, methods and applications, edn. London: Taylor & Francis.CrossRefGoogle Scholar
Weismann, A. (1883). Uber die Vererbung. English translation, On heredity (1889), translated by Shipley, A. E.: Oxford: Clarendon Press.Google Scholar
Weiss, H. & Maluszynska, J. (2000). Chromosomal rearrangement in autotetraploid plants of Arabidopsis thaliana. Hereditas, 133, 255–61.Google ScholarPubMed
Weiss, S. F. (2013). The Nazi symbiosis: human genetics and politics in the Third Reich. Chicago: University of Chicago Press.Google Scholar
Weiss-Schneeweiss, H.et al. (2013). Evolutionary consequences, constraints and potential of polyploidy in plants. Cytogenetic and Genome Research, 140, 137–50.CrossRefGoogle ScholarPubMed
Weldon, W. F. R. (1895a). The origin of the cultivated Cineraria. Nature, 52, 54, 104, 129.CrossRefGoogle Scholar
Weldon, W. F. R. (1895b). Remarks on variation in animals and plants. Proceedings of the Royal Society of London, 57, 379–82.Google Scholar
Weldon, W. F. R. (1898). Presidential address. Section D. Zoology. Nature, 58, 499–506.Google Scholar
Weldon, W. F. R. (1902a). On the ambiguity of Mendel's categories. Biometrika, 2, 44–55.CrossRefGoogle Scholar
Weldon, W. F. R. (1902b). Seasonal changes in the characters of Aster prenanthoides Muhl. Biometrika, 2, 113–14.Google Scholar
Wells, C. L.& Pigliucci, M. (2000). Adaptive phenotypic plasticity: the case of heterophylly in aquatic plants. Perspectives in Plant Ecology, Evolution and Systematics, 3, 1–18.CrossRefGoogle Scholar
Wells, W. C. (1818). An account of a White female, part of whose skin resembles that of a Negro. [Paper given at the Royal Society, 1813.] In Two essays upon dew and single vision. London.Google Scholar
Wendel, J. F. & Doyle, J. J. (1998). Phylogenetic incongruence: window into genome history and molecular evolution. In Molecular systematics of plants II, ed. Soltis, P. and Doyle, J., pp. 265–96. Boston: Kluwer.Google Scholar
Wendel, J. F. & Doyle, J. J. (2005). Polyploidy and evolution in plants. In Plant diversity and evolution, ed. Henry, R. J., pp. 97–117. Wallingford, UK: CABI Publishing.Google Scholar
Wenny, D. G. (2000). Seed dispersal, seed predation, and seedling recruitment of a neotropical montane tree. Ecological Monographs, 70, 331–51.CrossRefGoogle Scholar
Werth, C. R., Riopel, J. L. & Gillespie, N. W. (1984). Genetic uniformity in an introduced population of Witchweed (Striga asiatica) in the United States. Weed Science, 32, 645–8.Google Scholar
Werth, C. R., Guttman, S. I. & Eshbaugh, W. H. (1985a). Electrophoretic evidence of reticulate evolution in the Appalachian Asplenium complex. Systematic Botany, 10, 184–92.CrossRefGoogle Scholar
Werth, C. R., Guttman, S. I. & Eshbaugh, W. H. (1985b). Recurring origins of allopolyploid species in Asplenium. Science, 228, 731–3.CrossRefGoogle ScholarPubMed
Westerbergh, A. (1975). Serpentine and non-serpentine Silene dioica plants do not differ in nickel tolerance. Plant and Soil, 167, 297–303.Google Scholar
Westergaard, K. B.et al. (2008). Genetic diversity and distinctiveness in Scottish alpine plants. Plant Ecology & Diversity, 1, 329–38.CrossRefGoogle Scholar
Westerling, A. L.et al. (2006). Warming and earlier spring increase Western U.S. forest wildfire activity. Science, 313, 940–3.CrossRefGoogle ScholarPubMed
Western, D. (2001). Human modified ecosystems and future evolution. Proceedings of the National Academy of Sciences, USA, 98, 5458–65.CrossRefGoogle ScholarPubMed
Western, D., Russell, S. & Cuthill, I. (2009). The status of wildlife in protected areas compared to non-protected areas of Kenya. PLoS ONE, 4, 1–6.CrossRefGoogle ScholarPubMed
Wettstein, R. (1895). Der Saison-Dimorphismus als Ausgangpunkt für die Bildung neuer Arten im Pflanzenreich. Berichte der Deutschen botanischen Gesellschaft, 13, 303–13.Google Scholar
Wheeler, B. D. & Shaw, S. C. (1995). Restoration of damaged peatlands. London: HMSO.Google Scholar
Wheeler, Q. D. (ed.) (2008). The new taxonomy. Systematics Association Special Volume, Series No. 76. Boca Raton, FL, & London: CRC Press.CrossRefGoogle Scholar
Wheeler, W. C. (2012). Clocks and rates. In Systematics: a course of lectures. Wiley: published online: 17 May 2012, DOI:10.1002/9781118301081CrossRefGoogle Scholar
White, G. M., Boshier, D. H. & Powell, W. (2002). Increased pollen flow counteracts fragmentation in a tropical dry forest: an example from Swietenia humilis Zuccarini. Proceedings of the National Academy of Sciences, USA, 99, 2038–42.CrossRefGoogle Scholar
White, M. J. D. (1978). Modes of speciation. San Francisco: Freeman & Company.Google Scholar
White, O. E. (1917). Inheritance studies in Pisum. 2. The present state of knowledge of heredity and variation in peas. Proceedings of the American Philosophical Society, 56, 487–588.Google Scholar
Whitehouse, H. L. K. (1950). Multiple-allelomorph incompatibility of pollen and style in the evolution of the angiosperms. Annals of Botany, 14, 199–216.CrossRefGoogle Scholar
Whitehouse, H. L. K. (1959). Cross- and self-fertilisation in plants. In Darwin's biological work, ed. Bell, P. R., pp. 207–61. London: Cambridge University Press.Google Scholar
Whitehouse, H. L. K. (1965). Towards an understanding of the mechanism of heredity. London: Arnold. [edn, 1973.]Google Scholar
Whitney, G. G. (1994). From coastal wilderness to fruited plain: a history of environmental change in temperate North America, 1500 to the present. Cambridge: Cambridge University Press.Google Scholar
Whitney, K. D., Randell, R. A. & Rieseberg, L. H. (2010). Adaptive introgression of abiotic tolerance traits in the sunflower Helianthus annuus. New Phytologist, 187, 230–9.CrossRefGoogle ScholarPubMed
Whitton, J.et al. (2008). The dynamic nature of apomixis in the angiosperms. International Journal of Plant Science, 169, 169–82.CrossRefGoogle Scholar
Wichmann, M. C.et al. (2009). Human-mediated dispersal of seeds over long distances. Proceedings of the Royal Society, B, 276, 523–32.CrossRefGoogle ScholarPubMed
Wickler, W. (1968). Mimicry in plants and animals. New York: McGraw-Hill.Google Scholar
Widén, B. (1991). Phenotypic selection on flowering phenology in Senecio integrifolius, a perennial herb. Oikos, 61, 205–15.CrossRefGoogle Scholar
Widén, B. (1993). Demographic and genetic effects on reproduction as related to population size in a rare, perennial herb, Senecio integrifolius. Biological Journal of the Linnean Society, 50, 179–95.CrossRefGoogle Scholar
Widén, M. (1992). Sexual reproduction in a clonal, gynodioecious herb Glechoma hederacea. Oikos, 63, 430–8.CrossRefGoogle Scholar
Wiens, D. (1978). Mimicry in plants. Evolutionary Biology, 11, 365–403.Google Scholar
Wikström, N., Savolainen, V. & Chase, M. W. (2007). Evolution of the angiosperms: calibrating the family tree. Proceedings of the Royal Society, B, 268, 2211–20.Google Scholar
Wilcove, D. S. & Master, L. L. (2005). How many endangered species are there in the United States?Frontiers of Ecology and the Environment, 3, 414–420.CrossRefGoogle Scholar
Wiley, E. O. (1981). Phylogenetics: the theory and practice of phylogenetic systematics. New York: Wiley.Google Scholar
Wiley, E. O. & Lieberman, B. S. (2011) Phylogenetics: the theory and practice of phylogenetic systematics, edn. Hoboken: Wiley-Blackwell.CrossRefGoogle Scholar
Wilkins, D. A. (1959). Sampling for genecology. Record of the Scottish Plant Breeding Station, 1959, 92–6.
Wilkins, D. A. (1960). Recognising adaptive variants. Proceedings of the Linnean Society of London, 171, 122–6.CrossRefGoogle Scholar
Wilkins, J. S. (2009). Species: a history of the idea. Berkeley, Los Angeles & London: University of California Press.Google Scholar
Willerslev, E. & Cooper, A. (2005). Ancient DNA. Proceedings of the Royal Society, B, 272, 3–16.CrossRefGoogle ScholarPubMed
Willerslev, E.et al. (2004). Long-term persistence of bacterial DNA. Current Biology, 14, R9–R10.CrossRefGoogle ScholarPubMed
Willerslev, E., Hansen, A. J. & Poinar, H. N. (2004). Isolation of nucleic acids and cultures from ice and permafrost. Trends in Ecology and Evolution, 19, 141–7.CrossRefGoogle ScholarPubMed
Williams, K. & Gilbert, W. L. (1981). Insects as selective agents on plant vegetative morphology: egg mimicry reduces egg laying by butterflies. Science, 212, 467–9.CrossRefGoogle ScholarPubMed
Williams, M. (1989). Americans and their forests. Cambridge: Cambridge University Press.Google Scholar
Williamson, P. G. (1981). Morphological stasis and developmental constraint: real problems for Neo-Darwinism. Nature, 294, 214–15.CrossRefGoogle Scholar
Willis, K. & McElwain, J. (2014). The evolution of plants, edn. Oxford: Oxford University Press.Google Scholar
Willis, K. J., Gillson, L. & Brncic, T. M. (2004). How ‘virgin’ is virgin rainforest?Science, 304, 402–3.Google ScholarPubMed
Willson, M. F. (1979). Sexual selection in plants. American Naturalist, 113, 777–90.CrossRefGoogle Scholar
Willson, M. F. (1983). Plant reproductive ecology. New York: Wiley.Google Scholar
Willyard, A., Conn, R. & Liston, A. (2009). Reticulate evolution and incomplete lineage sorting among the ponderosa pines. Evolution, 52, 498–511.Google ScholarPubMed
Wilmott, A. J. (1949). Intraspecific categories of variation. In British flowering plants and modern systematic methods, ed. Wilmott, A. J., pp. 28–45. London: Botanical Society of the British Isles.Google Scholar
Wilson, E. O. (1985). The biological diversity crisis. BioScience, 35, 700–6.CrossRefGoogle Scholar
Wilson, G. B. & Bell, J. N. B. (1985). Studies of the tolerance to sulphur dioxide of grass populations in polluted areas. IV. The spatial relationship between tolerance and a point source of pollution. New Phytologist, 102, 563–74.Google Scholar
Wilson, M. A., Gaut, B. & Clegg, M. T. (1990). Chloroplast DNA evolves slowly in the Palm family (Arecaceae). Molecular Biology & Evolution, 7, 303–14.Google Scholar
Winchester, A. M. (1966). Genetics. Boston: Houghton.Google Scholar
Winfield, M. & Parker, J. (2000). A molecular analysis of Gentianella in Britain. English Nature Species Recovery Programme/Plantlife Report, No. 155.
Winge, Ø. (1917). The chromosomes, their numbers and general importance. Comptes Rendus des Travaux du Laboratoire Carlsberg, 13, 131–275.Google Scholar
Winge, Ø. (1940). Taxonomic and evolutionary studies in Erophila based on cytogenetic investigations. Comptes Rendus des Travaux du Laboratoire Carlsberg (Ser. Physiol.), 23, 41–74.Google Scholar
Winkler, H. (1908). Uber Parthenogenesis und Apogamie im Pflanzenreich. Progressus rei Botanicae, 2, 293–454.Google Scholar
Winkler, H. (1916). Uber die experimentelle Erzeugung von Pflanzen mit abweichenden Chromosomenzahlen. Zeitschrift für Botanik, 8, 417–531.Google Scholar
Winsor, M. P. (1995). The English debate on taxonomy and phylogeny, 1937–1940. History and Philosophy of the Life Sciences, 17, 227–52.Google ScholarPubMed
Wolff, K., Rogstad, S. H. & Schaal, B. A. (1994). Population and species variation of minisatellite DNA in Plantago. Theoretical and Applied Genetics, 87, 733–40.CrossRefGoogle ScholarPubMed
Woltereck, R. (1909). Weitere experimentelle Untersuchungen über Artveränderung, speziel über das Wesen quantitativer Artunterschiede bei Daphniden. Verhandlungen der deutschen zoologischen Gesellschaft, 19, 110–73.Google Scholar
Wood, T. E., Burke, J. M. & Rieseberg, L. H. (2005). Parallel genotypic adaptation: when evolution repeats itself. Genetica, 123, 157–70.CrossRefGoogle Scholar
Woodell, S. R. J. (1965). Natural hybridization between the Cowslip (Primula veris L.) and the Primrose (P. vulgaris Huds.) in Britain. Watsonia, 6, 190–202.Google Scholar
Woodson, R. E. Jr (1964). The geography of flower color in Butterflyweed. Evolution, 18, 143–63.CrossRefGoogle Scholar
Woodwell, G. M. (1990). The Earth in transition: patterns and processes of biotic impoverishment. Cambridge: Cambridge University Press.Google Scholar
Wookey, P. A.et al. (1993). Comparative responses of phenology and reproductive development to simulated environmental change in sub-Arctic and high-Arctic plants. Oikos, 67, 490–502.CrossRefGoogle Scholar
Wright, J. W. (1953). Pollen dispersion studies: some practical applications. Journal of’ Forestry, 51, 114–18.Google Scholar
Wright, K. M. (2013). Indirect evolution of hybrid lethality due to linkage with selected locus in Mimulus guttatus. PLOS Biology: published 26 February 2013. DOI:10.1371/journal.pbio.1001497.CrossRef
Wright, S. (1931). Evolution in Mendelian populations. Genetics, 16, 97–159.Google ScholarPubMed
Wright, S. (1938). Size of population and breeding structure in evolution. Science, 87, 430–1.Google Scholar
Wright, S. (1943). Isolation by distance. Genetics, 28, 114–28.Google ScholarPubMed
Wright, S. (1946). Isolation by distance under diverse systems of mating. Genetics, 31, 39–59.Google Scholar
Wright, S. (1950). Genetical structure of populations. Nature, 166, 247–9.CrossRefGoogle ScholarPubMed
Wright, S. (1966). Mendel's ratios. In The origin of genetics, ed. Stern, C. & Sherwood, E. R., pp. 173–5. London & San Francisco: Freeman.Google Scholar
Wright, S. (1977). Evolution and the genetics of populations, vol. 3: Experimental results and evolutionary deductions. Chicago & London: University of Chicago Press.Google Scholar
Wright, S. I., Kalisz, S. & Slotte, T. (2013). Evolutionary consequences of self-fertilization in plants. Proceedings of the Royal Society of London, B, 280, http://dx.doi.org/10.1098/rspb.2013.0133CrossRefGoogle ScholarPubMed
Wu, J., Hettenhausen, C. & Baldwin, I. (2006). Evolution of proteinase inhibitor defenses in North American allopolyploid species of Nicotiana. Planta, DOI:10.1007/s00425-006-0256-6.CrossRef
Wu, L. (1990). Colonization and establishment of plants in contaminated sites. In Heavy metal tolerance in plants: evolutionary aspects, ed. Shaw, A. J., pp. 269–84. Boca Raton, FL: CRC Press.Google Scholar
Wu., L., Bradshaw, A. D. & Thurman, D. A. (1975). The potential for evolution of heavy metal tolerance in plants. III. The rapid evolution of copper tolerance in Agrostis stolonifera. Heredity, 34, 165–87.CrossRefGoogle Scholar
Wu, L., Till-Bottraud, I. & Torres, A. (1987). Genetic differentiation in temperature-enforced seed dormancy among golf course populations of Poa annua L. New Phytologist, 107, 623–31.
Wyatt, R. (1988). Phylogenetic aspects of the evolution of self-pollination. In Plant evolutionary biology, ed. Gottleib, L. D. & Jain, S. K., pp. 109–31. London: Chapman & Hall.Google Scholar
Wyatt, R. E., Evans, A. & Sorenson, J. C. (1992). The evolution of self pollination in granite outcrop species of Arenaria (Caryophyllaceae). VI. Electrophoretically detectable genetic variation. Systematic Botany, 17, 201–9.CrossRefGoogle Scholar
Wyse Jackson, P. S. & Sutherland, L. A. (2000). International agenda for botanic gardens in conservation. Kew: Botanic Gardens Conservation International.Google Scholar
Xi, Z.et al. (2012). Horizontal transfer of expressed genes in a parasitic flowering plant. BMC Genomics, 13, 227.CrossRefGoogle Scholar
Xiao, Z. S., Jansen, P. A. & Zhang, Z. B. (2006). Using seed-tagging methods for assessing post-dispersal seed fate in rodent-dispersed trees. Forest Ecology & Management, 223, 18–23.CrossRefGoogle Scholar
Yang, L.et al. (2014). Next-generation sequencing, FISH mapping and synteny-based modeling reveal mechanisms of decreasing dysploidy in Cucumis. Plant Journal, 77, 16–30.CrossRefGoogle ScholarPubMed
Yannic, G., Baumel, A. & Ainouche, M. (2004). Uniformity of the nuclear and chloroplast genomes of Spartina maritima (Poaceae), a salt-marsh species in decline along the Western European Coast. Heredity, 93, 182–8.CrossRefGoogle Scholar
Yates, F. (1960). Sampling methods for censuses and surveys, edn. London: Griffin.Google Scholar
Yates, F. (1981). Sampling methods for censuses and surveys, edn. London: Charles Griffin.Google Scholar
Ye, Q., Bunn, E. & Dixon, K. W. (2011). Failure of sexual reproduction found in micropropagated critically endangered plants prior to reintroduction: a cautionary tale. Botanical Journal of the Linnean Society, 165, 278–84.CrossRefGoogle Scholar
Yeo, P. F. (1975). Some aspects of heterostyly. New Phytologist, 75, 147–53.CrossRefGoogle Scholar
Yoder, J. B.et al. (2014). Genomic signature of adaptation to climate in Medicago truncatula. Genetics, 196, 1263–75.CrossRefGoogle ScholarPubMed
Yoshida, S.et al. (2010). Horizontal gene transfer by the parasitic plant Striga hermonthica. Science, 328, 1128.CrossRefGoogle ScholarPubMed
Young, A. G. & Merriam, H. G. (1994) Effects of forest fragmentation on the spatial genetic structure of Acer saccharum Marsh. (sugar maple) populations. Heredity, 72, 201–8.CrossRefGoogle Scholar
Young, M. & Edis, T. (2004). Why intelligent design fails: a scientific critique of the new Creationism. New Brunswick, NJ, & London: Rutgers University Press.Google Scholar
Youngner, V. B. (1960). Environmental control of initiation of the inflorescence, reproductive structures and proliferations in Poa bulbosa. American Journal of Botany, 47, 753–7.CrossRefGoogle Scholar
Yu, Q.et al. (2009). Distinct non-target-site mechanisms endow resistance to glyphosate, ACCase and ALS-inhibiting herbicides in multiple herbicide-resistant Lolium rigidum populations. Planta, 230, 713–23.CrossRefGoogle Scholar
Yule, G. U. (1902). Mendel's laws and their probable relations to intra-racial heredity. New Phytologist, 1, 193–207, 222–38.CrossRefGoogle Scholar
Zalapa, J. E., Brunet, J. &. Guries, R. P. (2009). Patterns of hybridization and introgression between invasive Ulmus pumila (Ulmaceae) and native U. rubra. American Journal of Botany, 96, 1116–28.CrossRefGoogle ScholarPubMed
Zander, B. & Wiegleb, G. (1987). Biosystematische Untersuchungen an Populationen von Ranunculus subgen. Batrachium in Nordwest-Deutschland. Botanische Jahrbücher für Systematik, Planzengeschichte und Planzengeographie, 109, 81–130.Google Scholar
Zander, R. H. (2007). Paraphyly and the species concept, a reply to Ebach et al. Taxon, 56, 642–4.CrossRefGoogle Scholar
Zangerl, A. R. & Berenbaum, M. R. (2005). Increase in toxicity of an invasive weed after reassociation with its coevolved herbivore. Proceedings of the National Academy of Sciences, USA, 102, 15529–32.CrossRefGoogle ScholarPubMed
Zapiola, M. L. & Mallory-Smith, C. A. (2012) Crossing the divide: gene flow produces intergeneric hybrid in feral transgenic Creeping Bentgrass population. Molecular Ecology, 21, 4672–80.CrossRefGoogle ScholarPubMed
Zapiola, M. L.et al. (2008). Escape and establishment of transgenic glyphosate-resistant creeping bentgrass Agrostis stolonifera in Oregon, USA: a 4-year study. Journal of Applied Ecology, 45, 486–94.CrossRefGoogle Scholar
Zavada, M. S. (2007). The identification of fossil angiosperm pollen and its bearing on the time and place of the origin of angiosperms. Plant Systematics and Evolution, 263, 117–34.CrossRefGoogle Scholar
Zelikova, T. J.et al. (2013). Eco-evolutionary responses of Bromus tectorum to climate change: implications for biological invasions. Ecology & Evolution, 3, 1374–87.CrossRefGoogle ScholarPubMed
Zevenhuizen, E. (2000). Keeping and scrapping: the story of a Mendelian Lecture Plate of Hugo de Vries. Annals of Science, 57, 329–52.CrossRefGoogle Scholar
Zhang, W.et al. (2013). Species-specific identification from incomplete sampling: applying DNA barcodes to monitoring invasive Solanum plants. PLoS ONE, 8, 1–7.Google ScholarPubMed
Zhao, K. Y.et al. (2010). Genomic diversity and introgression in O. sativa reveal the impact of domestication and breeding on the Rice genome. PLoS ONE, 5, e10780CrossRefGoogle Scholar
Zhou, Z., Yang, X. & Yang, Q. (2006). Land bridge and long-distance dispersal – old views, new evidence. Chinese Science Bulletin, 51, 1030–8.CrossRefGoogle Scholar
Ziegenhagen, B., Bialozyt, R. & Liepelt, S. (2004). Contrasting molecular markers reveal gene flow via pollen is much more effective than gene flow via seeds. In Biological resources and migration, ed. Werner, D., pp.239–51. Heidelberg: Springer.Google Scholar
Zietkiewicz, E., Rafalski, A. & Labuda, D. (1994). Genome fingerprinting by simple sequence repeat (Ssr)-anchored polymerase chain-reaction amplification. Genomics 20, 176–83.CrossRefGoogle ScholarPubMed
Zirkle, C. (1941). Natural selection before the ‘Origin of species’. Proceedings of the American Philosophical Society, 84, 71–123.Google Scholar
Zirkle, C. (1966). Some anomalies in the history of Mendelism. In G. Mendel Memorial Symposium 1865–1965, ed. Sosna, M., pp. 31–7. Prague: Academia Publishing House of the Czechoslovak Academy of Sciences.Google Scholar
Ziska, L. H., Faulkner, S. & Lydon, J. (2004). Changes in biomass and root : shoot ratio of field grown Canada thistle (Cirsium arvense), a noxious invasive weed, with elevated CO2: implications for control with glyphosate. Weed Science, 52, 584–8.CrossRefGoogle Scholar
Zohary, D. (2004). Unconscious selection and the evolution of domesticated plants. Economic Botany, 58, 5–10.CrossRefGoogle Scholar
Zohary, D. & Feldman, M. (1962). Hybridisation between amphidiploids and the evolution of polyploids in the Wheat (Aegilops triticum) group. Evolution, 16, 44–61.CrossRefGoogle Scholar
Zohary, D. & Hopf, M. (1993). Domestication of plants in the Old World: the origin and spread of cultivated plants in West Asia, Europe and the Nile valley, edn. Oxford: Clarendon.Google Scholar
Zohary, D. & Nur, V. (1959). Natural triploids in the Orchard Grass Dactylis glomerata polyploid complex and their significance for gene flow from diploid to tetraploid levels. Evolution, 13, 311–17.CrossRefGoogle Scholar
Zuckerkandl, E. & Pauling, L. (1965). Molecules as documents of evolutionary history. Journal of Theoretical Biology, 8, 357–66.CrossRefGoogle ScholarPubMed

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • David Briggs, University of Cambridge, S. Max Walters, University of Cambridge Botanic Garden
  • Book: Plant Variation and Evolution
  • Online publication: 05 June 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139060196.023
Available formats
×

Save book to Dropbox

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

  • References
  • David Briggs, University of Cambridge, S. Max Walters, University of Cambridge Botanic Garden
  • Book: Plant Variation and Evolution
  • Online publication: 05 June 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139060196.023
Available formats
×

Save book to Google Drive

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

  • References
  • David Briggs, University of Cambridge, S. Max Walters, University of Cambridge Botanic Garden
  • Book: Plant Variation and Evolution
  • Online publication: 05 June 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139060196.023
Available formats
×