Skip to main content Accessibility help
×
Hostname: page-component-cd9895bd7-dzt6s Total loading time: 0 Render date: 2024-12-27T22:18:00.511Z Has data issue: false hasContentIssue false

Selected Bibliography

Published online by Cambridge University Press:  11 November 2024

Christopher N. Page
Affiliation:
University of Exeter
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
Evolution of the Arborescent Gymnosperms
Pattern, Process and Diversity
, pp. 620 - 670
Publisher: Cambridge University Press
Print publication year: 2024

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

Selected Bibliography

Aalbersberg, G. & Litt, T. 1998. Multiproxy climate reconstructions for the Eemian and early Weichselian. Journal of Quaternary Science 13: 367390.3.0.CO;2-I>CrossRefGoogle Scholar
Aase, H.C. 1915. Vascular anatomy of the megasporophylls of conifers. Botanical Gazette 60: 277313.CrossRefGoogle Scholar
Abbott, R.J. & Brochmann, C. 2003. History and evolution of the arctic flora: in the footsteps of Eric Hulten. Molecular Ecology 12: 299313.CrossRefGoogle ScholarPubMed
Adams, D.C. & Jackson, J.F. 1997. A phylogenetic analysis of the southern pines (Pinus subsect. Australes Loudon): biogeographical and ecological implications. Proceedings Biological Society of Washington 110: 681692.Google Scholar
Adams, R.P. 1987. Yields and seasonal variation of phytochemicals from Juniperus species of the United States. Biomass 12: 129139.CrossRefGoogle Scholar
Adams, R.P. 2007. Juniperus maritima, the seaside Juniper, a new species from Puget Sound, North America. Phytologia 89: 263283.Google Scholar
Adams, R.P. 2009. Analyses and taxonomic utility of the cedarwood oils of the serrate leaf junipers of the western hemisphere. Phytologia 91: 117139.Google Scholar
Adams, R.P. & Kistler, J.R. 1991. Hybridisation between Juniperus erythrocarpa Cory and Juniperus pinchotii Sudworth in the Chisos Mountains, Texas. Southwestern Naturalist 36: 295301.CrossRefGoogle Scholar
Adams, R.P., Schwarzbach, A.E., Morris, J.A. & González Elizondo, S. 2007. Juniperus compacta, a new species from Mexico (Cupressaceae). Phytologia 89: 361369.Google Scholar
Adams, R.P., Morris, J.A. & Schwarzbach, A.E. 2008. The evolution of Caribbean Juniperus (Cupressaceae): terpenoids, RAPDs and DNA SNPs data. Phytologia 90: 103120.Google Scholar
Adams, R.P., Morris, J.A. & Schwarzbach, A.E. 2008. Taxonomic affinity of Rushforth’s Bhutan juniper and Juniperus indica using SNPs from nrDNA and cp trnC–trnD, terpenoids and RAPD data. Phytologia 90: 233245.Google Scholar
Adams, R.P., Fontinha, S.S., Ruiz, B.R., & Nogales, M. 2010. Speciation of the Juniperus cedrus and J. maderensis in the archipelagos of Canaries and Madeira based on terpenoids and petN-psbM sequences. Phytologia 92: 4455.Google Scholar
Adzmi, Y., Suhaimi, W.C., Husini, M.S.A., et al. 2010. Heterogeneity of soil morphology and hydrology on the 50Ha long-term ecological research plot at Pasoh, Peninsular Malaysia. Journal of Tropical Forest Science 22: 2135.Google Scholar
Agarwal, A. & Rajanikanth, A. 2004. Podocarpaceae wood from the Cretaceous of Cauveery Basin. Palaeobotanist 53: 173176.Google Scholar
Agee, J.K. 1993. Fire Ecology of Pacific Northwest Forests. Washington, DC: Island Press.Google Scholar
Agee, J.K., Wakimoto, R.H. & Biswell, H.H. 1978. Fire and fuel dynamics of Sierra Nevada conifers. Forest Ecology and Management 1: 255265.CrossRefGoogle Scholar
Aiba, S. & Kohyama, T. 1997. Crown architecture and life-history traits of 14 tree species in a warm-temperate rain forest: significance of spatial heterogeneity. Journal of Ecology 85: 611624.CrossRefGoogle Scholar
Albert, V.A., Backlund, A., Bremer, K., et al. 1994. Functional constraints and rbcL evidence for plant phylogeny. Annals of the Missouri Botanical Garden 81: 534567.CrossRefGoogle Scholar
Alison, T.D., Shimizu, T., Ohara, M. & Yamanaka, N. 2008. Variation in sexual reproduction in Taxus cuspidata Sieb. & Zucc. Plant Species Biology 23: 2532.CrossRefGoogle Scholar
Allen, J.C., Krieger, S.M., Walters, J.R. & Collazo, J.A. 2006. Associations of breeding birds with fire-influenced and riparian-upland gradients in a longleaf pine ecosystem. Auk 123: 11101128.Google Scholar
Allen, K.J., Cook, E.R., Francey, R.J. & Michael, K. 2001. The climatic response of Phyllocladus aspleniifolius (Labill.) Hook. F. in Tasmania. Journal of Biogeography 28: 305316.CrossRefGoogle Scholar
Allnutt, T., Thomas, P., Newton, A.C. & Gardner, M.F. 1998. Genetic variation in Fitzroya cupressoides cultivated in the British Isles, assessed using RAPDS. Edinburgh Journal of Botany 55: 329341.CrossRefGoogle Scholar
Allnutt, T.R., Newton, A.C., Premoli, A. & Lara, A. 2003. Genetic variation in the threatened South American conifer Pilgerodendron uviferum (Cupressaceae), detected using RAPD markers. Biological Conservation 114: 245253.CrossRefGoogle Scholar
Almaraz-Abarca, N., Gonzáles-Elizondo, M.S., Tena-Flores, J.A., et al. 2006. Foliar flavenoids distinguish Pinus leiophylla and Pinus chihuanuana (Coniferales: Pinaceae). Proceedings of the Biological Society of Washington 119: 426436.CrossRefGoogle Scholar
Alvin, K.K. & Boulter, M.C. 1974. A controlled method of comparative study for Taxodiaceous leaf cuticles. Botanical Journal of the Linnean Society 69(4): 277.CrossRefGoogle Scholar
Alvin, K.L. 1982. Cheiroledidaceae: biology, structure, and paleoecology. Review of Palaeobotany and Palynology 37: 7198.CrossRefGoogle Scholar
Andersen, K.M., Turner, B.J. & Dalling, J.W. 2014. Seedling performance trade-offs influencing habitat filtering along a soil nutrient gradient in a tropical forest. Ecology 95: 33993413.CrossRefGoogle Scholar
Anderson, C.L., Channing, A. & Zamuner, A.B. 2009. Life, data and fossilization on Gran Canaria: implications for Macaronesian biogeography and molecular dating. Journal of Biogeography 36: 21892201.CrossRefGoogle Scholar
Anderson, E. 1940. The concept of the genus. Bulletin of the Torrey Botanical Club 67: 363369.CrossRefGoogle Scholar
Anderson, E. 1953. Introgressive hybridisation. Biological Review 28: 280307.CrossRefGoogle Scholar
Anderson, H.M. 1978. Podozamites and associated cones and scales from the Upper Triassic Molteno Formation, Karoo basin, South Africa. Palaeontologica Africana 21: 5777.Google Scholar
Anderson, J.A.R. & Muller, J. 1975. Palynological study of a Holocene peat and a Miocene coal deposit from N.W. Borneo. Review of Palaeobotany and Palynology 19: 291351.CrossRefGoogle Scholar
Anderson, J.E. & Romme, W.H. 1991. Initial floristics in lodgepole pine (Pinus contorta) forests following the 1988 Yellowstone fires. International Journal of Wildfire 1: 119124.Google Scholar
Anderson, J.M. & Anderson, H.M. 2003. Heyday of the Gymnosperms: Systematics and Biodiversity of the Late Triassic Molteno Fructifications. Pretoria: National Botanical Institute.Google Scholar
Anderson, J.M., Anderson, H.M. & Cleal, C.J. 2007. Brief history of the Gymnospermae: classification, biodiversity, phytogeography and ecology. Strelizia 20: 1280.Google Scholar
Anderson, R.S. 1996. Postglacial biogeography of Sierra lodgepole pine (Pinus contorta var. murrayana) in California. Ecoscience 3: 343351.CrossRefGoogle Scholar
Andrews, E.C. 1916. The geological history of the Australian flowering plants. American Journal of Science 42: 171232.CrossRefGoogle Scholar
Andrews, H.N. 1961. Studies in Palaeobotany. New York: Wiley.Google Scholar
Andrews, H.N. 1963. Early seed plants. Science 142: 925931.CrossRefGoogle ScholarPubMed
Andruchow-Colombo, A.R., Wilf, P. & Escapa, I.C. 2019. A South American relative of Phyllocladus? Huncocladus laubenfelsii gen. et sp. nov. (Podocarpaceae) from the early Eocene of Laguna del Hunco, Patagonia, Argentina. Australian Systematic Botany 32: 291319.Google Scholar
Ann, W., Syring, J., Gernandt, D.S., Liston, A. & Cronn, R. 2007. Fossil calibration of molecular divergence infers a moderate mutation rate and recent radiation for Pinus. Molecular Biology and Evolution 24: 90101.Google Scholar
Antonovics, J., Bradshaw, A.D. & Turner, R.G. 1971. Heavy metal tolerance in plants. Advances in Ecological Research 7: 185.CrossRefGoogle Scholar
Archangelsky, S. 1963. A new Mesozoic flora of Tico, Santa Cruz, Argentina. Bulletin of the British Museum (Natural History), Geology 8: 4792.CrossRefGoogle Scholar
Archangelsky, S. & Cuneo, R. 1987. Ferrugiocladaceae, a new conifer family from the Permian of Gondwana. Review of Palaeobotany and Palynology 51: 330.CrossRefGoogle Scholar
Archangelsky, S. & Del Fueyo, G. 1989. Squamastrobus gen nov., a fertile podocarp from the early Cretaceous of Patagonia, Argentina. Review of Palaeobotany and Palynology 59: 109126.CrossRefGoogle Scholar
Archangelsky, S. & Gamerro, J.C. 1967. Pollen grains found in coniferous cones from the Lower Cretaceous of Patagonia (Argentina). Review of Palaeobotany and Palynology 5: 179182.CrossRefGoogle Scholar
Archangelsky, S. & Romero, E.J. 1987. Polen de Gymnospermas (Coniferás) del Cretácico Superior y Paleógeno e Patagonia. Ameghiniana 11: 217236.Google Scholar
Areces, A. 1987. Consideraciones sobre la supuesta presencia de Pinus sylvestris L. en Oligoceno de Cuba. Publicationes Centre Investigaciones Petrol, Series Geologique 2: 2740.Google Scholar
Arianoutsou, M., Rundel, P.W. & Berry, W.L. 1993. Serpentine endemics as biological indicators of soil elemental concentrations. Pp 179189 in Markert, B. (ed.), Plants as Biomonitors: Indicators of Heavy Metals in the Terrestrial Environment. Weinheim: VCH.Google Scholar
Arkley, R.J. 1981. Soil moisture use by a mixed conifer forest in a summer-dry climate. Journal of the Soil Science Society of America 45: 423427.CrossRefGoogle Scholar
Armesto, J.J., Mitchell, J.D. & Villagran, C. 1986. A comparison of spatial patterns of trees in some tropical and temperate forests. Biotropica 18: 111.CrossRefGoogle Scholar
Arno, S. 1977. Northwest Trees. Seattle, WA: The Mountaineers.Google Scholar
Arno, S. & Habeck, J.R. 1972. Ecology of alpine larch (Larix lyallii) in the Pacific Northwest. Ecological Monographs 42: 417450.CrossRefGoogle Scholar
Arnold, M.L. 2006. Evolution through Genetic Exchange. Oxford: Oxford University Press.Google Scholar
Ash, J. 1983. Growth rings in Agathis robusta and Araucaria cunninghamii from tropical Australia. Australian Journal of Botany 31: 269275.CrossRefGoogle Scholar
Ash, J. 1985. Growth rings and longevity of Agathis vitiensis (Seeman) Benth. & Hook. F. ex Drake in Fiji. Australian Journal of Botany 33: 8188.CrossRefGoogle Scholar
Ashton, P.S 1964. Ecological studies in the mixed dipterocarp forests of Brunei state. Oxford Forestry Memoirs 25: 175.Google Scholar
Ashworth, A.C. & Hoganson, J.W. 1993. The magnitude and rapidity of the climate change marking the end of the Pleistocene in the mid-latitudes of South America. Palaeogeography, Palaeoclimatology, Palaeoecology 101: 263270.CrossRefGoogle Scholar
Atkinson, I.A.E. & Greenwood, R.M. 1989. Relationships between moas and plants. New Zealand Journal of Ecology 12(suppl.): 6796.Google Scholar
Aubréville, A. 1960. Flore du Cambodge, du Laos et du Viêt-Nam. Paris: Museum National d’Histoire Naturelle.Google Scholar
Aučina, A., Rudawska, M., Leski, T., et al. 2007. Growth and mycorrhizal community structure of Pinus sylvestris seedlings following the addition of forest litter. Applied and Environmental Microbiology 73: 48674873.CrossRefGoogle ScholarPubMed
Audley-Charles, M.G. 1987. Dispersal of Gondwanaland: relevance to evolution of the angiosperms. Pp 525 in Whitmore, T.C. (ed.), Biogeographical Evolution of the Malay Archipelago. Oxford: Clarendon Press.Google Scholar
Audley-Charles, M.G., Balantyne, P.D. & Hall, R. 1988. Mesozoic–Cenozoic rift–drift sequence of Asian fragments from Gondwanaland. Tectonophysics 155: 317330.CrossRefGoogle Scholar
Averdieck, F.R. 1971. Zur postglazialen Geschichte der Eibe (Taxus baccata) in Nordwestdeutschland. Flora 160: 2842.CrossRefGoogle Scholar
Awashi, N., Mehrotra, R.C. & Lakhanpal, R.N. 1992. Occurrence of Podocarpus and Messua in the Oligocene sediments of Mankum Coalfield, Assam, India. Geophytology 22: 193198.Google Scholar
Axelrod, D.I. 1962. A Pliocene Sequoiadendron forest from western Nevada. University of California Publications in Geological Sciences 39: 195268.Google Scholar
Axelrod, D.I. 1976. History of the Coniferous Forests of California and Nevada. Stanford, CA: University of California Press.Google Scholar
Axelrod, D.I. 1984. An interpretation of Cretaceous and Tertiary biota in polar regions. Palaeogeography, Palaeoclimatology, Palaeoecology 45: 105147.CrossRefGoogle Scholar
Axelrod, D.I. & Raven, P.H. 1982. Paleobiogeography and origin of the New Guinea flora. Pp 919941 in Gressit, J.L. (ed.), Biogeography and Ecology of New Guinea. The Hague: W.Junk.CrossRefGoogle Scholar
Bagnell, C.R. 1975. Species distinction among pollen grains of Abies, Picea and Pinus in the Rocky Mountain area, a scanning electron microscope study. Review of Palaeobotany and Palynology 19: 203220.CrossRefGoogle Scholar
Bagnoli, F., Vendramin, G.C., Buonamici, A., et al. 2009. Is Cupressus sempervirens native in Italy? An answer from genetic palaeobotanical data. Molecular Ecology 18: 22762286.CrossRefGoogle ScholarPubMed
Bailey, D.K. 1970. Phytogeography and taxonomy of Pinus subsection Balfourianae. Annals of the Missouri Botanical Garden 57: 210249.CrossRefGoogle Scholar
Bailey, D.K. 1975. Pinus albicaulis. Kew Magazine 691: 141147.Google Scholar
Bailey, D.K. 1987. A study of Pinus subsection Cembroides I: the single-needle pinyons of the Californias and the Great Basin. Notes from the Royal Botanic Garden Edinburgh 44: 275310.Google Scholar
Bailey, D.K. & Hawksworth, F.G. 1979. Pinyons of the Chihuahuan Desert Region. Phytologia 44: 129133.Google Scholar
Bailey, D.K. & Hawksworth, F.G. 1983. Pinaceae of the Chihuahuan Desert Region. Phytologia 53: 226234.Google Scholar
Baille, I.C., Ashton, P.S., Court, M.N., et al. 1987. Site characteristics and the distribution of tree species in mixed dipterocarp forest on Tertiary sediments in Central Darawak, Malaysia. Journal of Tropical Ecology 3: 201220.CrossRefGoogle Scholar
Bajpai, U. 1991. On Ginkgoites leaves from the early Permian of Rajmahal Hills, Bihar, India. Ameghiniana 28: 145148.Google Scholar
Baker, G. 1965. Basement complex rocks in the Cyclops ranges, Sentai region of Dutch New Guinea. Nova Guinea n.s. 6: 307328.Google Scholar
Baker, H.G. 1959. The contribution of autecological and genecological studies to our knowledge of the past migration of plants. American Naturalist 93: 255272.CrossRefGoogle Scholar
Baker, R.T. 1903. On a new species of Callitris from eastern Australia. Proceedings of the Linnean Society of New South Wales 28: 839841.CrossRefGoogle Scholar
Baker, R.T. & Smith, H.G. 1910. Pines of Australia. Sydney.Google Scholar
Baker, R.T. & Smith, H.G. 1910. A research on the pines of Australia: the genus Athrotaxis. New South Wales Technical Education Series 16: 303313.Google Scholar
Baldeck, C.A., Harms, K.E., Yavitt, J.B., John, R. & Turner, B.J. 2013. Soil resource and topography shape local tree community structure in tropical forests. Proceedings of the Royal Society: Biological Sciences 280: 17.Google ScholarPubMed
Ban, Y., Xu, H., Bergeron, Y. & Kneeshaw, D.D. 1998. Gap regeneration of shade-intolerant Larix gmelini in old growth boreal forests of northeastern China. Journal of Vegetation Science 9: 529536.CrossRefGoogle Scholar
Banda, M.B. & Prakash, U. 1984. A podocarpaceaous fossil wood from the Deccan intertrappean beds of Malabar Hills, Bombay, India. Geophytology 14: 171177.Google Scholar
Banerji, I. & Jama, B.N. 2003. Petrified araucarian remains from Sonajori, Rajmahal Basin, India. Palaeobotanist 52: 5562.Google Scholar
Bannister, M.H. 1958. Specimens of two pine trees from Guadalupe Island, Mexico. New Zealand Journal of Forestry 7: 8187.Google Scholar
Bannister, M.H. & McDonald, J.R. 1983. Turpentine composition of the pines of Guadeloupe and Cedros Islands, Baja California. New Zealand Journal of Botany 21: 373377.CrossRefGoogle Scholar
Bannister, P. & Neuner, G. 2001. Frost resistance and the distribution of conifers. Pp 321 in Bigras, F. & Colombo, S. (eds.), Tree Physiology. Vol 1. Conifer Cold Hardiness. Dordrecht: Springer.CrossRefGoogle Scholar
Barale, G. 1992. De noveaux restes fossils attributés aux Araucariacées dan les calcaires lithographiques du Crétacé inférieur du Montsec (Province de Lérida, Espagne). Review of Palaeobotany and Palynology 75: 5374.CrossRefGoogle Scholar
Baranova, I.U.P. & Grinenko, O.V. (eds.). 1989. Paleogene and Neogene of Northeastern USSR. Siberia: Yakut Scientific Centre.Google Scholar
Barbour, M. 1988. California upland forests and woodlands. Pp 131164 in Barbour, M.G. & Bilings, W.D. (eds.), North American Terrestrial Vegetation. Cambridge: Cambridge University Press.Google Scholar
Barbour, M.G. & Major, J. (eds.). Terrestrial Vegetation of California. New York: Wiley.Google Scholar
Barden, L.S. 2000. Population maintenance of Pinus pungens Lam. (Table Mountain pine) after a century without fire. Natural Areas Journal 20: 227233.Google Scholar
Barker, M.G. & Ashenden, T.W. 1993. Foliar injury in young Betula pendula Roth, Salix purpurea L. & Ilex aquifolium L. trees and in propagated Taxus baccata L. shoots exposed to intermittent fog at a range of acidities. Environmental Pollution 80: 123127.CrossRefGoogle Scholar
Baron, E.J. 1983. A warm equable Cretaceous: the nature of the problem. Earth Science Reviews 19: 305338.CrossRefGoogle Scholar
Barreda, V.D. 2012. Cretaceous/Paleogene floral turnover in Patagonia: drop in diversity, low extinction and a Classopollis spike. PLoS One 7(12): e52455.CrossRefGoogle Scholar
Barrett, P.M. 1999. A sauropod dinosaur from the lower Lufeng formation (Lower Jurassic) of Yunnan province, Peoples’ Republic of China. Journal of Vertebrate Paleontology 19: 785787.CrossRefGoogle Scholar
Barrett, P.M., McGowan, A.J. & Page, V. 2009. Dinosaur diversity and the rock record. Proceedings of the Royal Society of London B 276: 26672674.Google ScholarPubMed
Bartholomew, B., Boufford, D.E. & Sponberg, S.A. 1983. Metasequoia glyptostroboides: its present status in Central China. Journal of the Arnold Arboretum 64: 105128.CrossRefGoogle Scholar
Basinger, J.F. 1981. The vegetative body of Metasequoia milleri from the Middle Eocene of southern British Columbia. Canadian Journal of Botany 59: 23792410.CrossRefGoogle Scholar
Bauer, K., Grauvogel-Stamm, L., Kustatscher, E. & Krings, M. 2013. Fossil ginkgophyte seedlings from the Triassic of France resemble modern Ginkgo biloba. BMC Evolutionary Biology 13(127): 8.CrossRefGoogle ScholarPubMed
Beadle, N.C.W. 1962. An alternative hypothesis to account for the generally low phosphate content of Australian soils. Australian Journal of Agricultural Research 13: 434442.CrossRefGoogle Scholar
Beadle, N.C.W. 1981. The Vegetation of Australia. Cambridge: Cambridge University Press.Google Scholar
Beck, C.B. 1970. The appearance of gymnospermous structure. Biological Review 45: 379400.CrossRefGoogle Scholar
Beck, C.B. (ed). 1973. Origin and Early Evolution of Angiosperms. New York: Columbia University Press.Google Scholar
Beck, C.B. 1981. Archaeopteris and its role in vascular plant evolution. Pp 193223 in Niklas, K.J. (ed.), Paleobotany, Paleoecology and Evolution. New York: Praeger.Google Scholar
Becker, P. 2000. Competition and the regeneration niche between conifers and angiosperms; Bond’s slow seedling hypothesis. Functional Ecology 14: 401412.CrossRefGoogle Scholar
Becking, J.H. 1966. Mycorrhize de Podocarpus. Physiologie et morphologies. Annales de L’Institut Pasteur 111: 295302.Google Scholar
Ben Brahim, M., Loustau, D., Gaudillère, J.P. & Saaur, E. 1996. Effects of phosphate deficiency on photosynthesis and accumulation of starch and soluble sugars in 1-year old seedings of maritime pine (Pinus pinaster Ait.). Annales Science Forestières 53: 801810.CrossRefGoogle Scholar
Benkman, C.W. 1995. Wind dispersal capacity of pine seeds and the evolution of different seed dispersal modes in pines. Oikos 73: 221224.CrossRefGoogle Scholar
Benkman, C.W., Balda, R.P. & Smith, C.C. 1984. Adaptations for seed dispersal and the compromises due to seed predation in limber pine. Ecology 65: 632642.CrossRefGoogle Scholar
Benowicz, A. & El Kassaby, Y.A. 1999. Genetic variation in mountain hemlock (Tsuga mertensiana Bong.): quantitative and adaptive attributes. Forest Ecology and Management 123: 205215.CrossRefGoogle Scholar
Benson, L. & Darrow, R.A. 1981. Trees and Shrubs of the Southwestern Deserts. Tucson, AZ: University of Arizona Press.Google Scholar
Berbee, M.L. & Taylor, J.W. 1993. Dating the evolutionary radiation of the true fungi. Canadian Journal of Botany 71: 11141127.CrossRefGoogle Scholar
Bergstedt, J. & Milberg, P. 2001. The impact of logging intensity on field-layer vegetation in Swedish boreal forests. Forest Ecology and Management 154: 105115.CrossRefGoogle Scholar
Bernal, R., Gradstein, S.R. & Cells, U. (eds.). 2015. Catâlogo de plantas y ligeres de Colombia. Bogata: Universidad Nacional de Colombia, Instituto de Ciencias.Google Scholar
Berry, E.W. 1916. The lower Eocene floras of southeastern North America. U.S. Geological Survey Professional Papers 91: 1481.Google Scholar
Berry, E.W. 1925. Fossil plants from the Tertiary of Patagonia and their significance. Proceedings of the National Academy of Sciences, USA 11: 404405.CrossRefGoogle ScholarPubMed
Berry, E.W. 1929. The Kootenai and Lower Blairmore floras. National Museum of Canada Bulletin 58: 2854.Google Scholar
Berry, E.W. 1935. A tertiary Ginkgo from Patagonia. Torreya 35: 1113.Google Scholar
Berry, E.W. 1937. A Paleocene Flora from Patagonia. Baltimore, MD: Johns Hopkins University Press.Google Scholar
Berube, Y., Ritland, C. & Ritland, K. 2003. Isolation, characterization, and cross-species utility of microsatellites in yellow cedar (Chamaecyparis nootkatensis). Genome 46: 353361.CrossRefGoogle ScholarPubMed
Beveridge, A.E. 1964. Dispersal and destruction of seed in central North Island podocarp forest. Proceedings of the New Zealand Ecological Society 11: 4855.Google Scholar
Beveridge, A.E. 1973 Regeneration of podocarps in a central North Island forest. New Zealand Journal of Forestry 18: 2335.Google Scholar
Bharathan, G. & Zimmer, E.A. 1995. Early branching events in monocotyledons: partial 18S ribosomal DNA sequence analysis. Pp 81107 in Rudall, P.J., Cribb, P.J., Cutler, D.F. & Humphries, C.J. (eds.), Monocotyledons: Systematics and Evolution. Kew: Royal Botanic Gardens.Google Scholar
Bieleski, R.L. 1959. Factors affecting growth and distribution of kauri (Agathis australis Salisb.). Australian Journal of Botany 7: 252294.CrossRefGoogle Scholar
Bingham, M. & Simard, S. 2012. Mycorrhizal networks affect ectomycorrhizal fungal community similarity between conspecific trees and seedlings. Mycorrhiza 22: 317326.CrossRefGoogle ScholarPubMed
Bingham, R.T. 1972. Taxonomy, crossability, and relative blister-rust resistance of 5-needled white pines. Pp 271278 in Bingham, R.T. (ed.), Biology of Rust Resistance in Forest Trees. Washington, DC: USDA.Google Scholar
Biondi, F. & Fessenden, J.E. 1999. Radiocarbon analysis of Pinus lagunae tree rings: implications for tropical dendrochronology. Radiocarbon 41: 241249.CrossRefGoogle Scholar
Biondi, F., Cayan, D.R. & Berger, W.H. 1997. Dendroclimatology of Torrey pine (Pinus torreyana Parry ex Carr.). American Midland Naturalist 138: 237251.CrossRefGoogle Scholar
Birks, H.J.B. & Birks, H.H. 2004. The rise and fall of forests. Science 305: 484485.CrossRefGoogle ScholarPubMed
Blackburn, D.T. 1981. Tertiary megafossil flora of Maslin bay, South Australia: numerical taxonomic study of selected leaves. Alcheringa 5: 928.CrossRefGoogle Scholar
Blasubramanian, S.V, Alderfer, J.L. & Straubinger, R.M. 2006. Solvent- and concentration- dependent molecular interactions of taxol (Paraclitaxel). Journal of Pharmaceutical Science 83: 14701476.CrossRefGoogle Scholar
Blatt, H. & Jones, R.L. 1975. Proportions of exposed igneous, metamorphic and sedimentary rocks. Bulletin of the Geological Society of America: 21: 724730.Google Scholar
Block, W.M., Ganey, J.L., Scott, P.E. & King, R. 2005. Prey ecology of Mexican spotted owls in pine-oak forests of northern Arizona. Journal of Wildlife Management 69: 618629.CrossRefGoogle Scholar
Blyakharchuk, T.A., Wright, H.E., Borodavko, P.S., van der Knaap, W.O. & Ammann, B. 2004. Late Glacial and Holocene vegetational history of the Altai Mountains (southwestern Tuva Republic, Siberia). Palaeogeography, Palaeoclimatology, Palaeoecology 245: 518534.CrossRefGoogle Scholar
Bobrov, A.V.F.C. 1997. System und Phylogenie der Ordnung Taxales nach den Angaben der Samenstruktur. Scripta Botanica Belgica 15: 29.Google Scholar
Bobrov, A.V.F.Ch. 1997. Structure of seeds of representatives of genus Torreya Arn. (Torreyaceae Nakai) in connection with systematic position. P 42 in Proceedings of the VI Conference of Young Botanists, St. Petersburg (in Russian).Google Scholar
Bobrov, A.V.F.C. 1997. Morphology, anatomy and ultrastructure of seeds of Cephalotaxus Siebold & Zucc. (Cephalotacaceae Neger) in connection with problems of system and phylogeny of the genus. Pp 18–19 in Proceedings of the International Conference of Plant Anatomy, St. Petersburg (in Russian).Google Scholar
Bobrov, A.V.F.C. 1997. Systematic position and phylogeny of genus Amentotaxus Pilg. (Amentotaxaceae Kudo & Yamamoto) on base of data of structure of female reproductive organs. Pp 19–20 in Proceedings of the International Conference of Plant Anatomy, St. Petersburg (in Russian).Google Scholar
Bobrov, A.V.F.C., Melikian, A.P. & Yembaturova, E.Y. 1999. Seed morphology, anatomy and ultrastructure of Phyllocladus L.C. & A. Rich. ex E Mirb. (Phyllocladaceae (Pilg.) Bessey) in connection with the generic system and phylogeny. Annals of Botany 83: 601618.CrossRefGoogle Scholar
Bocher, T.W. 1964. Morphology of the vegetative body of Metasequoia glyptostroboides. Danish Botanical Arkiv 24: 170.Google Scholar
Bohra, D.R. & Sharma, B.D. 1980. Araucarites mittrii new species of petrified megastrobilus from the Rajmahal Hills, India. Ameghiniana 17: 38.Google Scholar
Bonan, G.B., Pollard, D. & Thompson, S.L. 1992. Effects of boreal forest vegetation on global climate. Nature 359: 716718.CrossRefGoogle Scholar
Bonaparte, J.F. 1979. Dinosaure: a Jurassic assemblage from Patagonia. Science 205: 13771379.CrossRefGoogle Scholar
Bonaparte, J.F. 1999. Evolucion de las vertebras presacras en Sauropodomorpha. Ameghiniana 36: 115189.Google Scholar
Bond, W.J., Lee, W.E. & Craine, J.H. 2002. Plant structural defences against browsing birds, a legacy of New Zealand’s extinct moas. Oikos 104: 500508.CrossRefGoogle Scholar
Bonell, M. & Gilmour, D.A. 1978. The development of overland flow in a tropical rainforest catchment. Journal of Hydrology 39: 365382.CrossRefGoogle Scholar
Booth, T.H., Nguyễn, H.N., Kirachbaum, M.U.F. & Jovanovic, T. 1999. Assessing impacts of possible climate change for forestry in Vietnam. Climate Change 41: 109129.CrossRefGoogle Scholar
Bootle, K.R. 1985. Wood in Australia: Types, Properties and Uses. Sydney: McGraw-Hill.Google Scholar
Borchert, M. 1985. Serotiny and cone-habit variation in populations of Pinus coulteri (Pinaceae) in the southern Coast Ranges of California. Madrõno 32: 2948.Google Scholar
Borchert, M., Johnson, M., Schreiner, D.S. & Vander Wall, S.B. 2003. Early postfire seed dispersal, seedling establishment and seedling mortality of Pinus coulteri (D.Don) in central coastal California. Plant Ecology 168: 207220.CrossRefGoogle Scholar
Bormann, F.H. & Likens, G.E. 1979. Pattern and Processes in a Forested Ecosystem. New York: Springer.CrossRefGoogle Scholar
Bose, M.N. 1955. Sciadopitytes variabilis n. sp. from the Arctic of Canada. Norske Geologisk Tidsskrift 35: 5368.Google Scholar
Bose, M.N. 1955. On two new conifers from the Jurassic of Yorkshire. Annals and Magazine of Natural History 8: 111120.CrossRefGoogle Scholar
Bose, M.N. 1959. Leaf-cuticle and other plant microfossils from the Mesozoic rocks of Andoya, Norway. Palaeobotanist 8: 17.Google Scholar
Bose, M.N., & Manum, S.B. 1991. Additions to the family Miroviaceae (Coniferae) from the Lower Cretaceous of West Greenland and Germany: Mirovia groenlandica n.sp., Tritaenia crassa (Seward) comb.nov., and Tritaenia linkii Magdefrau et Rudolph emend. Polar Research 9: 920.Google Scholar
Boulter, M.C. 1970. Lignified guard cell thickenings in the leaves of some modern and fossil species of Taxodiaceae (Gymnospermae). Biological Journal of the Linnean Society 2: 4146.CrossRefGoogle Scholar
Boulter, M.L. 1997. Plant evolution through the Phanerozoic. Geology Today 1997: 102106.CrossRefGoogle Scholar
Boureau, E. 1939. Rechercehes anatomiques et expérimentales sur l’ontogénie des paalntules de Pinacées et ses raports avec la phylogénie. Annales Sciences Naturelle Botanique 11(1): 1216.Google Scholar
Bowden, G.D. 1981. Coping with low nutrients. Pp 3464 in Pate, J.S. & McComb, R.J. (eds.), The Biology of Australian Plants. Nedlands: University of Western Australia Press.Google Scholar
Bowden, R.D., Geballe, G.T. & Bowden, W.B. 1989. Foliar uptake of 15N from simulated cloud water by red spruce (Picea rubens) seedlings. Canadian Journal of Forest Research 139: 383386.Google Scholar
Bowen, J.G. & Zachos, J.C. 2010. Rapid carbon sequestration at the termination of the Palaeocene–Eocene Thermal Maximum. Nature Geosciences 3: 866869.CrossRefGoogle Scholar
Boyce, C.K., Brodribb, T.J., Field, T.S. & Zwoeniecki, M.A. 2009. Angiosperm leaf-vein evolution: was it physiologically and environmentally transformative? Proceedings of the Royal Society B 276: 17711776.CrossRefGoogle Scholar
Boyd, R.S. & Jaffré, T. 2009. Elemental concentrations of eleven New Caledonian species from serpentine soils: elemental correlations and leaf-age. Northeastern Naturalist 16: 93110.CrossRefGoogle Scholar
Brady, P.V. & Carroll, S.A. 1994. Direct effects of CO2 and temperature on silicate weathering: possible implications for climate control. Geochimica Cosmochimica Acta 58: 18531856.CrossRefGoogle Scholar
Brea, M., Bellosi, E. & Krause, M. 2009. Taxaceoxylon katuatenkum n.sp. en la Formación Koluei-kaike (Eocene inferior-medio), Chubut, Argentina: un componente de los bosques subtropicales paleógenos de Patagonia. Ameghiniana 46: 127140.Google Scholar
Brennan, M.S. & Doyle, J. 1956. The gametophytes and embryogeny of Athrotaxis. Science Proceedings of the Royal Dublin Society 27: 193252.Google Scholar
Bretschneider, E. 1898. History of Botanical Discoveries in China. London.Google Scholar
Breuninger, M., Einig, W., Magel, E. Cardoso, E. & Hampp, R. 2000. Mycorrhiza of Brazil Pine (Araucaria angustifolia [Bert. O, Ktze]). Plant Biology 2: 410.CrossRefGoogle Scholar
Brodribb, T. 2011. A functional analysis of podocarp ecology. Pp 165173 in Turner, B.L. & Cernusak, L.A. Ecology of the Podocarpaceae in Tropical Forests. Washington, DC: Smithsonian Institution Scholarly Press.Google Scholar
Brodribb, T. & Field, T.S. 2008. Evolutionary significance of a flat-leaved Pinus in Vietnamese rainforest. New Phytologist 179: 201209.CrossRefGoogle Scholar
Brodribb, T. & Hill, R.S. 1996. The drought physiology of a diverse group of Southern Hemisphere conifer species is correlated with minimum seasonal rainfall. Functional Ecology 12: 465471.CrossRefGoogle Scholar
Brodribb, T. & Hill, R.S. 1997. Light response characteristics of a morphologically diverse group of southern hemisphere conifers as measured by chlorophyll fluorescence. Oecologia 110: 1017.CrossRefGoogle ScholarPubMed
Brodribb, T. & Hill, R.S. 1997. Imbricacy and stomatal wax plugs reduce maximum leaf-conductance in Southern Hemisphere conifers. Australian Journal of Botany 45: 657668.CrossRefGoogle Scholar
Brodribb, T.J., Holbrook, N.M. & Hill, R.S. 2005. Seedling growth in conifers and angiosperms: impacts of constraining xylem structures. Australian Journal of Botany 53: 749755.CrossRefGoogle Scholar
Brooks, R.R. 1987. Serpentine and Its Vegetation: A Multidisciplinary Approach. Portland, OR: Dioscoroides Press.Google Scholar
Brouwers, E., Clemens, W.A., Spicer, R.A., et al. 1987. Dinosaurs of the North Slope, Alaska: Latest Cretaceous environments. Science 237: 16081610.CrossRefGoogle ScholarPubMed
Brownlie, G. 1953. Embryology of the New Zealand species of the genus Podocarpus section Eupodocarpus. Phytomorphology 3: 295306.Google Scholar
Brummitt, R.K. 2007. Report on Nomenclature Committee for Vascular Plants. Taxon 59: 12891296.CrossRefGoogle Scholar
Brummitt, R.K. 1992. Vascular Plant Families and Genera. London: Royal Botanic Gardens, Kew.Google Scholar
Brummitt, R.K. & Powell, C.E. 1992. Authors of Plant Names. London: Royal Botanic Gardens, Kew.Google Scholar
Brundrett, M. 2004. Diversity and classification of mycorrhizal associations. Biological Review 79: 473495.CrossRefGoogle Scholar
Brunstein, F.C. & Yamaguchi, D.K. 1992. The oldest known Rocky Mountain bristlecone pines (Pinus aristata Engelm.). Arctic and Alpine Research 24: 253256.Google Scholar
Buchholz, J.T. 1919. Studies concerning the evolutionary significance of polycotyledony. American Journal of Botany 6: 106119.CrossRefGoogle Scholar
Buchholz, J.T. 1929. The embryogeny of conifers. Pp 359–392 in Proceedings of the International Congress of Plant Science, New York.Google Scholar
Buchholz, J.T. 1933. Determinate cleavage polyembryony, with special reference to Dacrydium. Botanical Gazette 94: 579588.CrossRefGoogle Scholar
Buchholz, J.T. 1934. The classification of Coniferales. Transactions of the Illinois State Academy of Science 25: 112113.Google Scholar
Buchholz, J.T. 1939. The embryogeny of Sequoia sempervirens with a comparison of the Sequoias. American Journal of Botany 26: 248257.CrossRefGoogle Scholar
Buchholz, J.T. 1939. The morphology and embryology of Sequoia gigantea. American Journal of Botany 26: 93101.CrossRefGoogle Scholar
Buchholz, J.T. 1939. Generic segregation of the Sequoias. American Journal of Botany 26: 535538.CrossRefGoogle Scholar
Buchholz, J.T. 1940. The embryogeny of Cunninghamia. American Journal of Botany 27: 877883.CrossRefGoogle Scholar
Buchholz, J.T. 1940. The embryogeny of Torreya, with a note on Austrotaxus. Bulletin of the Torrey Botanical Club 67: 731754.CrossRefGoogle Scholar
Buchholz, J.T. 1948. Generic and subgeneric distribution of the Coniferales. Botanical Gazette 110: 8091.CrossRefGoogle Scholar
Buchholz, J.T. 1948. A taxonomic revision of Podocarpus II. The American species of Podocarpus: Section Stachycarpus. Journal of the Arnold Arboretum 29(1): 6476.CrossRefGoogle Scholar
Buchholz, J.T. 1948. A taxonomic revision of Podocarpus III. The American species of Podocarpus: Section Polypodiopsis. Journal of the Arnold Arboretum 29(2): 117122.CrossRefGoogle Scholar
Buchholz, J.T. 1948. A taxonomic revision of Podocarpus IV. The American species of Section Eupodocarpus, sub-sections C and D. Journal of the Arnold Arboretum 29(2): 123151.CrossRefGoogle Scholar
Buchholz, J.T. 1951. A taxonomic revision of Podocarpus V. The South Pacific species of Podocarpus: Section Stachycarpus. Journal of the Arnold Arboretum 32(1): 8292.Google Scholar
Buchholz, J.T. & Gray, N.E. 1948. A taxonomic revision of Podocarpus I. The sections of the genus and their subdivisions with special reference to leaf anatomy. Journal of the Arnold Arboretum 29: 4963.CrossRefGoogle Scholar
Buchholz, J.T. & Gray, N.E. 1948. A taxonomic revision of Podocarpus IV. The American species of Eupodocarpus subsections C and D. Journal of the Arnold Arboretum 29: 123151.CrossRefGoogle Scholar
Buchholz, J.T. & Kaeiser, M. 1940. A statistical study of two variables in the Sequoias: pollen grain size and cotyledon number. American Naturalist 74: 279283.CrossRefGoogle Scholar
Buckley, B.M., Cook, E.R., Peterson, M.J., & Barbetti, M. 1997. A changing temperature response with elevation for Lagarostrobos franklinii in Tasmania, Australia. Climatic Change 36: 477498.CrossRefGoogle Scholar
Buckley, B.M., Ogden, J., Palmer, J.G., Fowler, A. & Salinger, J. 2000. Dendroclimatic interpretation of tree-rings in Agathis australia (kauri). 1. Climate correlation functions and master chrononlogy. Journal of the Royal Society of New Zealand 30: 263276.CrossRefGoogle Scholar
Bugala, W. 1984. Metasequoia glyptostroboides: 35 years in cultivation at the Kórnik Arboretum. Arboretum Kórnickie 28: 101112 (in Polish).Google Scholar
Bulgin, N.E., Lovelius, N.V. & Forsov, G.A. 1989. Response of Metasequoia glyptostroboides in Leningrad to changes in temperature and moisture regime. Botanicheskii Zhurnal 74: 13231328.Google Scholar
Bullock, S., Summerell, B.A. & Gunn, L.V. 2000. Pathogens of the wollemi pine, Wollemia nobilis. Australasian Plant Pathology 29: 211214.CrossRefGoogle Scholar
Bunn, A.G., Waggoner, L.A. & Graumlich, L.J. 2005. Topographic mediation of growth in high elevation foxtail pine (Pinus balfouriana Grev. et Balf.) forests in the Sierra Nevada, USA. Global Ecology and Biogeography 14: 103114.CrossRefGoogle Scholar
Burley, J.F. 1965. Genetic variation in Picea sitchensis. Commonwealth Forestry Review 44: 4759.Google Scholar
Burlingame, L.L. 1908. The staminate cone and male gametophyte of Podocarpus. Botanical Gazette 46: 161178.CrossRefGoogle Scholar
Burlingame, L.L. 1913. The morphology of Araucaria brasiliensis. Botanical Gazette 55: 97114.CrossRefGoogle Scholar
Burlingame, L.L. 1914. The morphology of Araucaria brasiliensis, II. The ovulate cone and female gametophyte. Botanical Gazette 56: 490508.CrossRefGoogle Scholar
Burlingame, L.L. 1915. The morphology of Araucaria brasiliensis, III. Botanical Gazette 59: 138.CrossRefGoogle Scholar
Burlingame, L.L. 1915. The origin and relationships of the araucarians. I. Botanical Gazette 60: 126.CrossRefGoogle Scholar
Burlingame, L.L. 1915. The origin and relationships of the araucarians. II. Botanical Gazette 60: 89114.CrossRefGoogle Scholar
Burns, R.M. & Honkala, B.H. (eds.). Silvics of North America. Vol. 1. Conifers. Washington, DC: USDA.Google Scholar
Burrett, C., Duhig, N., Berry, R. & Varne, R. 1991. Asian and southwestern Pacific continental terranes derived from Gondwana, and their biogeographic significance. Australian Systematic Botany 4: 1324.CrossRefGoogle Scholar
Burrows, G.E. & Stockey, R.A. 1994. The developmental anatomy of cryptogeal germination in Bunya pine (Araucaria bidwillii). International Journal of Plant Science 155: 519537.CrossRefGoogle Scholar
Burrows, J.E. & Burrows, S.M. 1987. Mount Mulanje, last refuge of the giants. Veld and Flora 73: 122124.Google Scholar
Burslem, D.F.R.P., Whitmore, T.C. & Brown, G.C. 2000. Short-term effects of cyclone impact and long term recovery of tropical rain forest on Kolombangara, Solomon Islands. Journal of Ecology 88: 10631087.CrossRefGoogle Scholar
Burt Davy, J. 1926. A Manual of the Flowering Plants and Ferns of the Transvaal with Swaziland. London: Longmans Green & Co.Google Scholar
Busgen, M., Munch, E. & Thomson, T. 1929. The Structure and Life of Forest Trees. London: Chapman & Hall.Google Scholar
Businsky, R. 1999. Study of Pinus dalatensis Ferré and the enigmatic ‘Pin du Moyen Annam’. Candollea 54: 125143.Google Scholar
Businsky, R. 2004. A revision of the Asian Pinus subsection Strobus (Pinaceae). Willdenovia 34: 209257.CrossRefGoogle Scholar
Butala, J.R. & Cridland, A.A. 1978. Nomenclature of fossil Glyptostrobus in North America. Taxon 27: 1520.CrossRefGoogle Scholar
Butts, B. & Buchholz, J.T. 1940. Cotyledon numbers in conifers. Transactions of the Illinois Academy of Sciences 33: 5862.Google Scholar
Cai, Q. & Liu, Y. 2006. Temperature variability since 1776 inferred from tree rings of Pinus tabulaeformis in Mt. Helan. Acta Geographica Sinica 61: 929936.Google Scholar
Cain, S.A. 1944. Foundations of Plant Geography. New York: Academic Press.Google Scholar
Calder, J.A. & Taylor, R.L. 1968. Flora of the Queen Charlotte Islands. Part 1. Systematics of the Vascular Plants. Ottowa: Queen’s Printer.Google Scholar
Calvo-Polanco, M., Zwiazek, J., Jones, M. & MacKinnon, M. 2008. Responses of mycorrhizal jack pine (Pinus banksiana) seedlings to NaCl and boron. Trees: Structure and Function 22: 825834.CrossRefGoogle Scholar
Cameron, R.J. 1960. Natural regeneration of podocarps in the forests of the Whirinaki River Valley. New Zealand Journal of Forestry 8: 337354.Google Scholar
Camp, W.H. & Hubbard, M.M. 1963. On the origins of the ovule and cupule in lyginopterid pteridosperms. American Journal of Botany 50: 235243.CrossRefGoogle Scholar
Campbell, K.A. & Hawkins, C.D.B. 2003. Paper birch and lodgepole pine root reinforcement in coarse-, medium-, and fine-textured soils. Canadian Journal of Forest Research 33: 15801586.CrossRefGoogle Scholar
Campo, J. & Dirzo, R. 2003. Leaf quality and herbivory responses to soil nutrient addition in secondary tropical dry forests of Yucatan, Mexico. Journal of Tropical Ecology 19: 525530.CrossRefGoogle Scholar
Campo-Duplan, M. Van. 1951. Recherches sur la phylogenie des Taxodiacees d’apres leurs grains de pollen. Travaux du Laboratoire forestier de Toulouse 11, 4: 114.Google Scholar
Camus, A. 1914. Les Cypres: Encyclopedie Economique de Silviculture 2. Paris: Paul Lechevalier.Google Scholar
Camus, A. 1926. Le Cupressus dupreziana A. Camus, Cypres Nouveau du Tassili. Bulletin Societe Denrologique de France 58: 3944.Google Scholar
Cantrill, D.J. & Falcon-Lang, H.J. 2001. Cretaceous (Late Albian) coniferales of Alexander Island, Antarctica: leaves, reproductive structures and roots. Review of Palaeobotany and Palynology 115: 119145.CrossRefGoogle ScholarPubMed
Cantrill, D.J. & Poole, I. 2005. Taonomic turnover and abundance in Cretaceous to Tertiary wood floras of Antarctica: implications for changes in forest ecology. Palaeogeography, Palaeoclimatology, Palaeoecology 215: 205219.CrossRefGoogle Scholar
Cantrill, D.J., Tosolini, A.M.P. & Francis, J.E. Paleocene flora from Seymour Island, Antarctica: revision of Dusen’s (1908) pteridophyte and conifer taxa. Alcheringa 35: 309328.CrossRefGoogle Scholar
Carpenter, R.J., Bannister, J.M., Jordan, G.J. & Lee, D.E. 2010. Leaf fossils of Proteaceae tribe Persoonieae from the Late Oligocene–Early Miocene of New Zealand. Australian Systematic Botany 23: 115.CrossRefGoogle Scholar
Carrer, M. & Urbanati, C. 2004. Age-dependent tree-ring growth responses to climate in Larix decidua and Pinus cembra. Ecology 85: 730740.CrossRefGoogle Scholar
Carrer, M., Nola, P., Eduard, J.L., Motta, R. & Urbinati, C. 2007. Regional variability of climate-growth relationships in Pinus cembra high elevation forests in the Alps. Journal of Ecology 95: 10721083.CrossRefGoogle Scholar
Carrutheres, W. 1866. On araucarian cones from the Secondary Beds of Britain. Geological Magazine 3: 249252.CrossRefGoogle Scholar
Carvalho, A., Rebelo, A. & Dias, J. 1999. Distribution and natural regeneration of yew trees in the national parks of Peneda-Geres (Portugal) and Baixa Limia Serra-Xures (Spain). Revista de Biologia 17: 4349.Google Scholar
Casadio, S., Nelson, C., Taylor, P., Griffin, M. & Gordon, D. West Antarctic Rift system: a possible New Zealand–Patagonia Oliogocene paleobiogeographic link. Ameghiniana 47: 129132.CrossRefGoogle Scholar
Casey, W.H. & Sposito, G. 1992. On the temperature dependence of mineral dissolution rates. Geochimica Cosmochimica Acta 56: 38253830.CrossRefGoogle Scholar
Cassana, F.F. & Dillenburgh, L.R. 2013. The periodic wetting of leaves enhances water relations and growth of the long-lived conifer Araucaria angustifolia. Plant Biology 15: 7583.CrossRefGoogle ScholarPubMed
Castanera, D., Barco, J.L., Díaz-Martínez, I., et al. 2011. New evidence of a herd of titanosauriform sauropods from the lower Berriasian of the Iberian Range (Spain). Palaeogeography, Palaeoclimatology, Palaeoecology 310: 227237.CrossRefGoogle Scholar
Cázares, E., Trappe, J.M. & Jumpponen, A. 2005. Mycorrhizal-plant colonisation patterns on a subalpine glacier forefront as a model system of primary succession. Mycorrhiza 15: 405406.CrossRefGoogle Scholar
Cecchi, F.A., Mariotti, L.M. & Tani, G. 1991. Ultrastructural observation on the megasporocyte of Taxus baccata L. Taxaceae in relation to megaspore behaviour. Caryologia 44: 4554.Google Scholar
Chabot, B.F. & Hicks, D.J. 1982. The ecology of leaf life spans. Annual Review of Ecology and Systematics 13: 229259.CrossRefGoogle Scholar
Chamber, C.L. & Mast, J.N. 2005. Ponderosa pine snag dynamics and cavity excavation following wildfire in northern Arizona. Forest Ecology and Management 216: 227240.CrossRefGoogle Scholar
Chamberlain, C.J. 1935. Gymnosperms, Structure and Evolution. Chicago, IL: University of Chicago Press.Google Scholar
Chandler, M.E.J. 1922. Sequoia couttsiae Heer, at Hordle, Hants: a study of the characters which serve to distinguish Sequoia from Athrotaxis. Annals of Botany 36: 385390.CrossRefGoogle Scholar
Chandler, M.E.J. 1978. Supplement to the Lower Tertiary floras of southern England, UK. Part 5. Tertiary Research Special Papers 4: 147.Google Scholar
Chaney, R.W. 1948. The bearing of living Metasequoia on problems of Tertiary paleobotany. Proceedings of the National Academy of Sciences USA 34: 503515.CrossRefGoogle ScholarPubMed
Chaney, R.W. 1949. The Miocene occurrence of Sequoia and related conifers in the John Day Basin. Proceedings of the National Academy of Sciences USA 35: 125129.CrossRefGoogle Scholar
Chaney, R.W. 1954. A new pine from the Cretaceous of Minnesota and its paleoecological significance. Ecology 35: 145151.CrossRefGoogle Scholar
Chapin, F.S., Autumn, K. & Pugnaire, F. 1993. Evolution of suites of traits in response to environmental stress. American Naturalist 142: S78S92.CrossRefGoogle Scholar
Chapman, A.L. 1885. Torreya taxifolia Arnott: a reminiscence. Botanical Gazette 10: 251254.CrossRefGoogle Scholar
Chapman, J.D. & White, F. 1970. The Evergreen Forests of Malawi. Oxford: Commonwealth Forestry Institute.Google Scholar
Chaw, S.M., Sung, H., Long, H., Zharkikh, A. & Li, W.H. 1995. The phylogenetic positions of the conifer genera Amentotaxus, Phylocladus and Nageia inferred from 18S rRNA sequences. Journal of Molecular Evolution 41: 224230.CrossRefGoogle ScholarPubMed
Cheddai, R., Lamb, H.F., Guiot, J. & van der Kaars, S. 1998. Holocene climatic change in Morocco: a quantative reconstruction from pollen data. Climate Dynamics 14: 883890.CrossRefGoogle Scholar
Cheddadi, R., Vendramin, G.C., Litt, T., et al. 2006. Imprints of glacial refugia in the modern genetic diversity of Pinus sylvestris. Global Ecology and Biogeography 15: 271282.CrossRefGoogle Scholar
Cheeseman, T.F. 1925. Manual of the New Zealand Flora, 2nd edn. Wellington: Government Printer.Google Scholar
Chen, J., Tauer, C.G. & Huang, Y. 2002. Paternal chloroplast inheritance patterns in pine hybrids detected with trnL–trnF intergenic region polymorphism. Theoretical and Applied Genetics 104: 13071311.CrossRefGoogle ScholarPubMed
Chen, J.Q., Franklin, J.F. & Spies, T.A. 1992. Vegetation responses to edge environments in old-growth Douglas-fir forests. Ecological Applications 2: 387396.CrossRefGoogle ScholarPubMed
Chen, Z.K. & Wang, F.H. 1979. The gametophytes and fertilisation in Pseudotaxus. Acta Botanica Sinica 21: 1029 (in Chinese with English summary).Google Scholar
Chen, Z.-K. & Wang, F.-H. 1981. The early embryogeny of the genus Fokienia with a note on its systematic position. Acta Phtotaxonomica Sinica 19: 2328 (in Chinese, with English summary).Google Scholar
Cheng, S.-S., Wu, C.-L., Chang, H.-T. & Chang, S.-T. 2004. Anti-termitic and antifungal activities of essential oil of Calocedrus formosana leaf and its composition. Journal of Chemical Ecology 30: 19571967.CrossRefGoogle Scholar
Cheng, W.C. 1931. A new spruce from western China. Contributions from the Biological Laboratory of Science Society of China 6: 3336.Google Scholar
Cheng, W.C. 1933. Enumeration of gymnosperms from Kweichow collected by Y. Tsiang. Sinensia 2: 103.Google Scholar
Cheng, W.C. 1934. An enumeration of vascular plants from Chekiang. III. Contributions from the Biological Laboratory of the Science Society of China 9: 240241.Google Scholar
Cheng, W.C. 1961. [On Nothotsuga]. Nanjing Academy Forestry Research 32: 13 (in Chinese).Google Scholar
Cheng, W.C. & Fu, L.K. 1978. Flora Republicae Popularis Sinica, 7: Gymnospermae. Beijing: Science Press (in Chinese).Google Scholar
Cherrier, J.-F., Gondran, M., Woltz, P. & Vogt, G. 1992. Parasitisme interspécifique chez les Gymnospermes; Donnés inédites chez deux Podocarpaceae endémiques Néo-Caledoniennes. Revue Cytologique et Végétal-Botaniques 15: 6587.Google Scholar
Chesson, P.L. & Warner, R.R. 1981. Environmental variability promotes coexistence in lottery competitive environments. American Naturalist 117: 923943.CrossRefGoogle Scholar
Chevalier, A. 1944. Notes sur les conifers de l’Indochine. Revue de Botanique Applique et d’Agriculture Tropicale 24: 734.CrossRefGoogle Scholar
Chin, K. & Gill, B.D. 1996. Dinosaurs, dung beetles, and conifers: participants in a Cretaceous food web. Palaios 11: 280285.CrossRefGoogle Scholar
Chochiyeva, K.I. 1974. Rod Metasequoia Miki v paleoflore zapinsy Gruzii. [The genus Metasequoia in the palaeoflora of western Georgia]. Sak’art’velos SSR Mec’nierebat Akademie 76: 213216 (in Russian).Google Scholar
Chowdhury, R.C. 1961. Morphology and embryology of Cedrus deodara (Roxb.) Loud. Phytomorphology 11: 283304.Google Scholar
Christiansen, H. 1950. A tetraploid Larix decidua Miller. Det. Kgl. Danske Videnskaps. Selsk. 19: 19.Google Scholar
Christiansen, P. 1999. On the head size of sauropodomorphs dinosaurs: implications for ecology and physiology. History of Biology 13: 269297.CrossRefGoogle Scholar
Christison, R. 1897. The exact measurement of trees. (Part 3). The Fortingall Yew. Transactions of the Botanical Society of Edinburgh 13: 410435.CrossRefGoogle Scholar
Christophel, D.C. 1973. Sciadopitophyllum canadense gen. et spec. nov: a new conifer from western Alberta. American Journal of Botany 60: 6166.CrossRefGoogle Scholar
Chu, C.C. & Sun, C.S. 1981. Chromosome numbers and morphology of Cathaya. Acta Phytotaxonomica Sinica 19: 444446.Google Scholar
Chun, W.-Y. & Kuang, K.-Z. 1962. De genere Cathaya Chun et Kuang. Acta Botanica Sinica 10: 245246 (in Chinese).Google Scholar
Chung, J.D., Lin, T.P., Tan, Y.C., Lin, M.Y. & Hwang, S.Y. 2004. Genetic diversity and biogeography of Cunninghamia konishii (Cupressaceae), an island species in Taiwan: a comparison with Cunninghamia lanceolata, a mainland species in China. Molecular Phylogenetics and Evolution 33: 791801.CrossRefGoogle ScholarPubMed
Chung-MacCoubrey, A.L. 2003. Monitoring long-term reuse of trees by bats in pinyon–juniper woodlands of New Mexico. Wildlife Society Bulletin 31: 7379.Google Scholar
Churchill, S.P., Balslev, H., Forero, E. & Luteyn, J.L. (eds.), Biodiversity and Conservation of Neotropical Montane Forests. New York: New York Botanic Garden.Google Scholar
Clarkson, B.R., Clarkson, B.D. & Patel, R.N. 1992. The pre-Taupo eruption (c. AD 130) forest of the Benneydale-Pueora district, central North Island, New Zealand. Journal of the Royal Society of New Zealand 22: 6176.CrossRefGoogle Scholar
Clemens, S. 2006. Toxic metal accumulation, responses to exposure and mechanisms of tolerance in plants. Biochimie 88: 17071719.CrossRefGoogle ScholarPubMed
Clement-Westerhof, J.A. 1984. Aspects of Permian paleobotany and palynology. 4. The conifer Ortiseia Florin from the Val Gardana Formation of the Dolomites and Vicentinian Alps (Italy) with special reference to a revised concept of Walchiaceae (Goppert) Schimper. Review of Palaeobotany and Palynology 41: 51166.CrossRefGoogle Scholar
Clement-Westerhof, J.A. 1987. Aspects of Permian palaeobotany and palynology. VII. The Majonicaeceae, a new family of Late Permian conifers. Review of Palaeobotany and Palynology 52: 375402.CrossRefGoogle Scholar
Cluzel, D., Adams, C.J. & Meffre, S. 2010. Discovery of Early Cretaceous rocks in New Caledonia: new geochemical and U–Pb zircon age constraints on the transition from subduction to marginal breakup in the southwest Pacific. Journal of Geology 118: 381397.CrossRefGoogle Scholar
Cobb, F.W. & Libby, W.J. 1968. Susceptibility of Monterey, Guadalupe Island and Bishop pines to Scirrhia (Dothistroma) pini, the cause of red band blight. Phytopathology 58: 8890.Google Scholar
Cockayne, L. 1899. On the burning and reproduction of subalpine scrub and its associated plants, with special reference to Arthur’s Pass district. Transactions of the New Zealand Institute 31: 398419.Google Scholar
Cockburn, P. 1980. Trees of Sabah. Sabah Forest Record 10. Sabah: Borneo Forest Department.Google Scholar
Coe, M.L., Dilcher, D.L., Farlow, J.O., Jarzen, D.M. & Russell, D.A. 1987. Dinosaurs and land plants. Pp 225258 in Friis, E.M., Chaloner, W.G. & Crane, P.R. (eds.), The Origin of Angiosperms and Their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Coffey, K.L. & Kirkman, L.K. 2006. Seed germination strategies of species with restoration potential in fire-maintained pine savanna. Natural Areas Journal 26: 289299.CrossRefGoogle Scholar
Coker, W.C. 1902. Notes on the gametophytes and embryo of Podocarpus. Botanical Gazette 33: 89107.CrossRefGoogle Scholar
Coker, W.C. 1903. On the gametophytes and embryo of Taxodium. Botanical Gazette 36: 114141.CrossRefGoogle Scholar
Coker, W.C. 1909. Vitality of pine seeds and the delayed opening of cones. The American Naturalist 43: 677681.CrossRefGoogle Scholar
Colbert, E.H. 1993. Feeding strategies and metabolism in elephants and sauropod dinosaurs. American Journal of Science 293: 119.CrossRefGoogle Scholar
Cole, K.L. & Wahl, E. 2000. A late Holocene paleoecological record from Torrey Pines State Reserve, California. Quaternary Research 53: 341351.CrossRefGoogle Scholar
Collinson, M.E. 1986. Use of modern generic names for plant fossils. Pp 91104 in Spicer, R.A. & Thomas, B.A. (eds.), Systematic and Taxonomic Approaches in Palaeobotany. Oxford: Clarendon Press.Google Scholar
Collinson, M.E. 1990. Plant evolution and ecology during the Cainozoic diversification. Advances in Botanical Research 17: 198.CrossRefGoogle Scholar
Collinson, M.E. 2002. The ecology of Cainozoic ferns. Review of Palaeobotany and Palynology 119: 5168.CrossRefGoogle Scholar
Collinvaux, P.A. 1967. Bering Land Bridge: evidence of spruce in late Wisconsin times. Science 156: 380383.CrossRefGoogle Scholar
CONAF (Corporacion Nacional Forestal, Chile). 1988. Informacion Estadistica Historica de Occurrencia y Dano de los Incendios Forestales: Periodo 1978–1988. Decima region de los Lagos, Puerto Montt, Chile. Santiago: CONAF.Google Scholar
Conduit, R., Ashton, P.S., Bunyavejchewin, S., Dattaraja, H.S. & Davis, S.J. 2006. Importance of demographic niches to tree diversity. Science 313: 98101.CrossRefGoogle Scholar
Conkle, M.T. 1973. Growth data for 29 years from the California elevational transect study of ponderosa pine. Forest Science 19: 3139.Google Scholar
Conkle, M.T. 1992. Genetic diversity: seeing the forest through the trees. New Forests 6: 522.CrossRefGoogle Scholar
Conner, R.N. & Saenz, D. 2005. The longevity of large pine snags in eastern Texas. Wildlife Society Bulletin 33: 700705.CrossRefGoogle Scholar
Conner, W.H., Toliver, J.R. & John, R. 1990. Observations on the regeneration of baldcypress (Taxodium distichum (L.) Rich.) in Louisiana swamps. Southern Journal of Applied Forestry 14: 115118.CrossRefGoogle Scholar
Connor, W.H., Gosselink, J.G. & Parrondo, R.T. 1981. Comparison of the vegetation of three Louisiana swamp sites with different flooding regimes. American Journal of Botany 68: 320331.CrossRefGoogle Scholar
Contreras-Medina, R. & Luna Vega, I. 2002. On the distribution of gymnosperms genera, their areas of endemism and cladistic biogeography. Australian Systematic Botany 15: 193203.CrossRefGoogle Scholar
Conway-Morris, S. 1988. The evolution of diversity in ancient ecosystems: a review. Philosophical Transactions of the Royal Society of London B 353: 327345.CrossRefGoogle Scholar
Cook, E.R., Bird, T., Peterson, M., et al. 1991. Climatic change in Tasmania inferred from a 1089-year tree-ring chronology of subalpine huon pine. Science 253: 12661268.CrossRefGoogle Scholar
Cook, E.R., Buckley, B., D’Arrogo, R. & Peterson, M.D. 2000. Warm season temperatures since 1600 BC reconstructed from Tasmanian tree rings and their relationship to large-scale sea-surface temperature anomalies. Climate Dynamics 16: 7991.CrossRefGoogle Scholar
Cook, E.R., Esper, J. & D’Arrigo, R. 2004. Extra-tropical Northern hemisphere temperature variability over the past 1000 years. Quaternary Science Reviews 23: 20632074.CrossRefGoogle Scholar
Cook, E.R., Buckley, B.M., Palmer, J.G., et al. 2006. Millennia-long tree-ring records from Tasmania and New Zealand: a basis for modelling climate variability and forcing, past, present and future. Journal of Quaternary Science 21: 689699.CrossRefGoogle Scholar
Cook, E.R., Seager, R., Cane, M.A., & Stahle, D.W. 2007. North American drought: reconstructions, causes, and consequences. Earth-Science Reviews 81: 93134.CrossRefGoogle Scholar
Cook, P. 1939. A new type of embryogeny in the conifers. American Journal of Botany 26: 138143.CrossRefGoogle Scholar
Cookson, I.C. 1957. On some Australian Tertiary spores and pollen grains that extend the geological and geographical distribution of living genera. Proceedings of the Royal Society of Victoria 69: 4153.Google Scholar
Coomes, D.A. & Grubb, P.J. 1996. Amazonian caatinga and related communities at La Esmerelda, Venezuela: forest structure, physiognomy and floristics, and control by soil factors. Vegetatio 122: 167191.CrossRefGoogle Scholar
Coomes, D.A., Allen, R.B. & Bentley, W.A. 2005. The hare, the tortoise and the crocodile: the ecology of angiosperm dominance, conifer persistence and fern filtering. Journal of Ecology 93: 918935.CrossRefGoogle Scholar
Cooper, R.A. & Millener, P.R. 1993. The New Zealand biota: historical background and new research. Trends in Ecology and Evolution 8: 429433.CrossRefGoogle ScholarPubMed
Cope, E.A. 1998. Taxaceae: the genera and cultivated species. Botanical Review 64: 291323.CrossRefGoogle Scholar
Coppenheaver, C.A., Grinter, L.E., Lorber, J.H., Neatrour, M.A. & Spinney, M.P. 2002. A dendroecological and dendroclimatic analysis of Pinus virginiana and Pinus rigida at two slope positions in the Virginia piedmont. Castanea 67: 302315.Google Scholar
Coria, R.A. 1994. On a monospecific assemblage of sauropod dinosaurs from Patagonia: implications for gregarious behaviour. Gaia 10: 209213.Google Scholar
Cornish, V. 1946. The Churchyard Yew and Immortality. London: Frederick Muler.Google Scholar
Cornwell, W.K., Bedford, B.L. & Chapin, C.T. 2001. Occurrence of arbuscular mycorrhizal fungi in a phosphorous-poor wetland and mycorrhizal response to phosphorous fertilisation. American Journal of Botany 88: 18241829.CrossRefGoogle Scholar
Corrigan, D., Timoney, R.F. & Donnelly, O.M.X. 1978. N-alkanes and w-hydroxyalkanoic acids from the needles of twenty-eight Picea species. Phytochemistry 17: 907910.CrossRefGoogle Scholar
Costanza, S.H. 1985. Pennsylvanioxylon of Middle and Upper Pennsylvanian coals from the Illinois Basin and its comparison with Mesoxylon. Palaeontographical B 197: 81121.Google Scholar
Cotton, M.H., Hicks, R.R. Jr. & Flake, R.H. 1975. Morphological variability among loblolly and shortleaf pines of east Texas with reference to natural hybridisation. Castanea 40: 309319.Google Scholar
Coulter, J.M. & Chamberlain, C.J. 1910. Morphology of Gymnosperms. Chicago, IL: University of Chicago Press.CrossRefGoogle Scholar
Couper, R.A. 1960. New Zealand Mesozoic and Cainozoic plant microfossils. Palaeontological Bulletin, Wellington 32: 187.Google Scholar
Couvert-Bratland, K.A., Block, W.M. & Theimer, T.C. 2006. Hairy woodpecker winter ecology in ponderosa pine forests representing different ages since wildfire. Journal of Wildlife Management 70: 13791392.CrossRefGoogle Scholar
Cox, C.B. & Moore, P.D. 1993. Biogeography: An Ecological and Evolutionary Approach. Oxford: Blackwell Scientific.Google Scholar
Cox, R.E., Yamamoto, S., Otto, A. & Simoneit, B.R.T. 2007. Oxygenated di- and tricyclic diterpenoids of southern hemisphere conifers. Biochemical Systematics and Ecology 35: 342362.CrossRefGoogle Scholar
Crame, J.A., Lomas, S.A., Pirrie, D. & Luther, A. 1996. Late Cretaceous extinction patterns in Antarctica. Journal of the Geological Society of London 153: 503506.CrossRefGoogle Scholar
Crame, J.A., Francis, J.E., Cantrill, D.J. & Pirrie, D. 2004. Maastrichtian stratigraphy of Antarctica. Cretaceous Research 25: 411423.CrossRefGoogle Scholar
Crane, P.R. 1985. Phylogenetic analyses of seed plants and the origin of angiosperms. Annals of the Missouri Botanical Garden 72: 716793.CrossRefGoogle Scholar
Crane, P.R. 1987. Vegetational consequences of angiosperm diversification. Pp 107144 in Friis, E.M., Chaloner, W.G. & Crane, P.R. (eds.), The Origin of Angiosperms and their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Cranwell, L.M. 1940. Pollen grains of the New Zealand conifers. New Zealand Journal of Science and Technology 22B: 117.Google Scholar
Cranwell, L.M. 1942. New Zealand pollen studies. 1. Key to the pollen grains of families and genera in the native flora. Records of the Auckland Institute and Museum 2: 280308.Google Scholar
Cranwell, L.M. 1961. Coniferous pollen types of the southern hemisphere. I. Aberration in Acmopyle and Podocarpus dacrydioides. Journal of the Arnold Arboretum 42: 416423.CrossRefGoogle Scholar
Cranwell, L.M. & Von Post, L. 1936. Post-Pleistocene pollen diagrams from the southern hemisphere. Geografiska Annaler 18: 308347.CrossRefGoogle Scholar
Crisp, M.D., Linder, H.P. & Weston, P.H. 1995. Cladistic biogeography of plants in Australia and New Guinea: congruent pattern reveals two endemic tropical tracks. Systematic Biology 4: 457473.CrossRefGoogle Scholar
Crisp, M.D., Laffan, S., Linder, H.P. & Monro, A. 2001. Endemism in the Australian flora. Journal of Biogeography 28: 183198.CrossRefGoogle Scholar
Crisp, M.D., Arroyo, M.T.K., Cook, L.G., et al. 2009. Phylogenetic biome conservatism on a global scale. Nature 458: 754756.CrossRefGoogle ScholarPubMed
Critchfield, W.B. 1963. Hybridization of the southern pines in California. Southern Forest Tree Improvement Committee Publication 22: 40–48.Google Scholar
Critchfield, W.B. 1970. Shoot growth and heterophylly in Ginkgo biloba. Botanical Gazette 131: 150162.CrossRefGoogle Scholar
Critchfield, W.B. 1980. Genetics of Lodgepole Pine. Washington, DC: USDA.Google Scholar
Critchfield, W.B. 1986. Hybridization and classification of the white pines (Pinus section Strobus). Taxon 35: 647656.CrossRefGoogle Scholar
Critchfield, W.B. & Little, E.L. Jr. 1971. Geographic Distribution of the Pines of the World. Washington, DC: USDA.Google Scholar
Crompton, A.W. & Attridge, J. 1986. Masticatory apparatus of the larger herbivores during late Triassic and early Jurassic times. Pp 223236 in Padian, K. (ed.), The Beginning of the Age of the Dinosaurs. Cambridge: Cambridge University Press.Google Scholar
Cronquist, A. 1968. The Evolution and Classification of Flowering Plants. New York: Houghton Mifflin.Google Scholar
Cronquist, A. 1981. An Integrated System of Classification of the Flowering Plants. New York: Columbia University Press.Google Scholar
Cuneo, R. 1985. Ejamplares fertiles de Genoites patagonica Ferugio (Buriadiaceae, Coniferopsida?) del permico de Chubut, Republica Argentina. Ameghiniana 22: 269279.Google Scholar
D’Arrigo, R., Wilson, R. & Jacoby, G. 2006. On the long-term context for late twentieth century warming. Journal of Geophysical Research 111.Google Scholar
D’Emic, M.D., Wilson, J.A. & Thompson, R. 2010. The end of the sauropod dinosaur hiatus in North America. Palaeogeography, Palaeoclimatology, Palaeoecology 297: 486490.CrossRefGoogle Scholar
Dahlgren, R.A., Boettinger, J.L., Huntingdon, G.L. & Amundsen, R.C. 1997. Soil development along an elevational transect in the western Sierra Nevada, California. Geoderma 78: 207236.CrossRefGoogle Scholar
Dahlgren, R.M.T. 1980. A revised system of classification of the angiosperms. Botanical Journal of the Linnean Society 80: 91124.CrossRefGoogle Scholar
Dahlgren, R.M.T. 1983. General aspects of angiosperm evolution and macrosystematics. Nordic Journal of Botany 3: 119149.CrossRefGoogle Scholar
Dahlgren, R.M.T. & Clifford, H.T. 1982. The Monocotyledons: A Comparative Study. London: Academic Press.Google Scholar
Dahlgren, R.M.T., Clifford, H.T. & Yeo, P.F. 1985. The Families of the Monocotyledons: Structure, Evolution and Taxonomy. Berlin: Springer.CrossRefGoogle Scholar
Dallimore, W. 1908. Holly, Yew and Box. London: John Lane.Google Scholar
Dallimore, W. & Jackson, A.B. 1966. A Handbook of Coniferae and Ginkgoaceae. London: Edward Arnold.Google Scholar
Dalling, J.W., Muller-Landau, H.C., Wright, S.J. & Hubbell, S.P. 2002. Role of dispersal in the recruitment limitation of neotropical pioneer species. Journal of Ecology 90: 714727.CrossRefGoogle Scholar
Danjon, F., Forcaud, T. & Bert, D. 2005. Root architecture and wind-firmness of mature Pinus pinaster. New Phytologist 168: 387400.CrossRefGoogle ScholarPubMed
Dargavel, J., Hart, D. & Libbis, B. 2001. Perfumed Pineries: Environmental History of Australia’s Callitris Forests. Canberra: Australian National University Centre for Resource and Environmental Studies.Google Scholar
Darlington, C.D. 1956. Chromosome Botany. London: Arnold.Google Scholar
Darlington, C.D. & Wylie, A.P. 1955. Chromosome Atlas of Flowering Plants. London: G. Allen.Google Scholar
Daubenmire, R. 1972. On the relation between Picea pungens and Picea engelmanii in the Rocky Mountains. Canadian Journal of Botany 50: 733742.CrossRefGoogle Scholar
Davis, P.H. 1965. Flora of Turkey. Cambridge: Cambridge University Press.Google Scholar
Davis, P.H. & Heywood, V.H. 1963. Principles of Angiosperm Taxonomy. Edinburgh: Oliver & Boyd.Google Scholar
Davis, P.H. & Heywood, V.H. 1973. Principles of Angiosperms Taxonomy. New York: Kreiger.Google Scholar
Davis, S.D., Heywood, V.H. & Hamilton, A.C. (eds.). 1994. Centres of Plant Diversity: a Strategy for their Conservation. Vol. 1. Europe, Africa, South West Asia and the Middle East. Cambridge: IUCN.Google Scholar
Davis, S.D., Sperry, J.S. & Hacke, U.G. 1999. The relationship between xylem conduit diameter and cavitation caused by freezing. American Journal of Botany 86: 13671372.CrossRefGoogle ScholarPubMed
Dawson, J.W. 1963. New Caledonia and New Zealand: a botanical comparison. Tuatara 11: 178194.Google Scholar
Day, J.J., Upchurch, P., Norman, D.B., Gale, A.S. & Powell, H.P. 2002. Sauropod trackways, evolution and behaviour. Science 296: 1659.CrossRefGoogle Scholar
Day, R.H., Doyle, T.W. & Draugelis-Dale, R.O. 2004. Interactive effects of substrate, hydroperiod, and nutrients on seedling growth of Salix nigra and Taxodium distichum. Environmental and Experimental Botany 55: 163174.CrossRefGoogle Scholar
De Azkue, D. 1982. Los chromosomas de Pilgerodendron uviferum (Don) Flor. Darwiniana 24: 1922.Google Scholar
De Ferré, Y. 1960. Une nouvelle espece de pin au Viet-Nam: Pinus dalatensis. Bulletin Société Histoire Naturelle Toulouse 95: 171180.Google Scholar
De Ferré, Y. 1965. Structure des plantules et systematique du genre Pinus. Bulletin Société Histoire Naturelle Toulouse 100: 150.Google Scholar
De Ferré, Y. 1966. Validite de l’espece Pinus pumila et affinities systematiques. Bulletin Société Histoire Naturelle Toulouse 102: 351356.Google Scholar
De Ferré, Y. & Gaussen, H. 1970. Affinites systematiques de Pinus peuce Gris. Zbornik ne simpoziumot za molokata Skopje 2: 3338.Google Scholar
De Ferré, Y. & Rouane, M.L. 1978. A propos du genre Amentotaxus Pilger. Travaux Laboratoire Forestiere Toulouse 9: 16.Google Scholar
De Ferré, Y., Rouane, M.L. & Woltz, P. 1975. Plantules des Podocarpacées. Bulletin de la Société d’Historire Naturelle de Toulouse 111: 303321.Google Scholar
De Ferré, M.Y., Rouane, M.I. & Wolz, M.P. 1977. Systematique et anatomie compares des feulles de taxaceae, Podocarpaceae et Cupressaceae de Nouvele-Caledonie. Cahiers du Pacifique 20: 241266.Google Scholar
De Las Heras, J., Moya, D., López-Serrano, F. & Condes, S. 2007. Reproduction of postfire Pinus halepensis Mill: stands six years after silvicultural treatments. Annals of Forest Science 64: 5966.CrossRefGoogle Scholar
De Laubenfels, D.J. 1959. Parasitic conifer found in New Caledonia. Science 130: 97.CrossRefGoogle Scholar
De Laubenfels, D.J. 1962. The primitiveness of polycotyledony considered with special reference to the cotyledonary condition in Podocarpaceae. Phytomorphology 12: 296299.Google Scholar
De Laubenfels, D.J. 1965. The relationships of Fitzroya cupressoides (Molina) Johnston and Diselma archeri J.D.Hooker based on morphological considerations. Phytomorphology 15: 414419.Google Scholar
De Laubenfels, D.J. 1967. Podocarpus vitiensis in the Moluccas (Taxaceae). Blumea 15: 440.Google Scholar
De Laubenfels, D.J. 1969. A revision of the Malesian and Pacific rainforest conifers. 1. Podocarpaceae. Journal of the Arnold Arboretum 50: 274314.CrossRefGoogle Scholar
De Laubenfels, D.J. 1972. Flore de la Nouvelle Caledonie: 4. Gymnosperms. Paris: Museum d’Histoire Naturelle.Google Scholar
De Laubenfels, D.J. 1978. The genus Prumnopitys (Podocarpaceae) in Malesia. Blumea 24: 189190.Google Scholar
De Laubenfels, D.J. 1978. The Moluccan dammars (Agathis, Araucariaceae). Blumea 24: 499504.Google Scholar
De Laubenfels, D.J. 1978. The Podocarpus species of Ambon (Podocarpaceae). Blumea 24(2): 495497.Google Scholar
De Laubenfels, D.J. 1978. The taxonomy of Philippine Coniferae and Taxaceae. Kalikasan 7: 117152.Google Scholar
De Laubenfels, D.J. 1984. Decussocarpus. Pacific Plant Areas 4: 212213.Google Scholar
De Laubenfels, D.J. 1988. Coniferales. Pp 337453 in Van Steenis, C.G.G.J. & De Wilde, W.J.J.O. (eds.), Flora Malesiana, Vol. 10. Dordrecht: Kluwer.Google Scholar
Debazac, E.F. & Tomassone, R. 1965. Contribution a une etude comparee des Pins Mediterranees de la section Halepenses. Annales Scientifique Forestières 22: 211256.Google Scholar
Degi, G.S. 1984. Seedling survival and growth rates in experimental cypress domes. Pp 141144 in Ewel, K.C. & Odum, H.T. (eds.). Cypress Swamps. Gainesville, FL: University of Florida Press.Google Scholar
Del Fueyo, G.M.S. 1998. Coniferous woods from the Upper Cretaceous of Patagonia, Argentina. Revue Espana Paleontol 13: 4350.Google Scholar
Del Fueyo, G.M.S., Archangelsky, S., Llorens, M. & Cuneqo, N.R. 2008. Coniferous ovulate cones from the Lower Cretaceous of Santa Cruz province, Argentina. International Journal of Plant Sciences 169: 799813.CrossRefGoogle Scholar
Del Tredici, P. 2001. Sprouting in temperate trees: a morphological and ecological review. The Botanical Review 67: 121140.CrossRefGoogle Scholar
Del-Castillo, R.F. & Trujillo, S. 2008. Effect of inbreeding depression on outcrossing rates among populations of a tropical pine. New Phytologist 177: 517524.CrossRefGoogle Scholar
Delevoryas, T. 1963. Morphology and Evolution of Fossil Plants. New York: Holt Reinhart & Wineston.Google Scholar
Delevoryas, T. & Hope, R.C. 1973. Fertile coniferophyte remains from the Late Triassic Deep River Basin, North Carolina. American Journal of Botany 60: 810818.CrossRefGoogle Scholar
Delevoryas, T.D. & Hope, R.C. 1987. Further observations on the Late Triassic conifers Compsostrobus neotericus and Volzia andrewsii. Review of Palaeobotany and Palynology 51: 5964.CrossRefGoogle Scholar
Delgado, P., Salas-Lizana, R., Vázquez-Lobo, A., et al. 2007. Introgressive hybridisation in Pinus montezumae Lamb. and Pinus pseudostrobus Lindl. (Pinaceae): morphological and molecular (cpSSR) evidence. International Journal of Plant Sciences 168: 861875.CrossRefGoogle Scholar
Dettmann, M.E. 1963. Upper Mesozoic microfloras from south-eastern Australia. Proceedings of the Royal Society of Victoria 77: 1148.Google Scholar
Dettman, M.E. & Playford, G. 1968. Taxonomy of some Cretaceous spores and pollen grains from eastern Australia. Proceedings of the Royal Society of Victoria 81: 6993.Google Scholar
Dettmann, M.E., Clifford, H.T. & Peters, M. 2012. Emwadea microcarpa gen. et sp. nov – anatomically preserved araucarian seed cones from the Winton Formation (late Albian), western Queensland, Australia. Alcheringa 36: 217237.CrossRefGoogle Scholar
Dezzotti, A. & Sancholuz, L. 1991. Los bosques de Austrocedrus chilensis en Argentina: Ubicacion, estructura y crecimiento. Bosques 12: 4352.CrossRefGoogle Scholar
Diehl, P., Mazzaarino, M.J. & Fontelna, S. 2008. Plant limiting nutrients in Andean–Patagonian woody species: effects of interannual rainfall variation, soil fertility and mycorrhizal infection. Forest Ecology and Management 255: 29732980.CrossRefGoogle Scholar
Dillenburgh, L.R., Rosa, L.M.G. & Mosena, M. 2010. Hypocotyl of seedlings of the large-seeded species Araucaria angustifolia: an important underground sink for the seed reserves. Trees: Structure and Function 24: 705711.CrossRefGoogle Scholar
Doak, C.C. 1937. Morphology of Cupressus arizonica: gametophytes and embryogeny. Botanical Gazette 98: 808815.CrossRefGoogle Scholar
Dobzhansky, T. 1941. Genetics and the Origin of Species. New York: Columbia University Press.Google Scholar
Dodd, N.L., Schweinsburg, R.E. & Boe, S. 2006. Landscape-scale forest habitats relationships to tassel-eared squirrel populations: implications for ponderosa pine forest restoration. Restoration Ecology 14: 537547.CrossRefGoogle Scholar
Dodson, J.R. 1976. Pollen spectra from Chatham Island, New Zealand. New Zealand Journal of Botany 14: 341347.CrossRefGoogle Scholar
Dodson, J.R. & Macphail, M.K. 2004. Palynological evidence for aridity events and vegetation change during the Middle Pliocene, a warm period in southwestern Australia. Global Planetary Change 41: 285307.CrossRefGoogle Scholar
Dogra, P.D. 1964. Gymnosperms of India: II. Chilgoza Pine (Pinus gerardiana Wall.). Bulletin of the National Botanic Gardens Lucknow 109: 147.Google Scholar
Dogra, P.D. 1966. Embryogeny of the Taxodiaceae. Phytomorphology 16: 125141.Google Scholar
Dogra, P.D. 1980. Embryogeny of gymnosperms and taxonomy-assessments. Glimpses in Plant Research 5: 114128.Google Scholar
Dogra, P.D. 1999. Regional action plan: conifers of the Himalayas and their endangered genetic resources. Pp 5054 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Dong, J., Wagner, D.B., Yanchuk, A.D., et al. 1992. Paternal chloroplast DNA inheritance in Pinus contorta and Pinus banksiana: independence of parental species or cross direction. Journal of Heredity 83: 419422.CrossRefGoogle Scholar
Donoghue, M.J., Doyle, J.A., Gaultier, J., Kluge, A.G. & Rowe, T. 1989. The importance of fossils in phylogeny reconstruction. Annual Review of Ecology and Systematics 20: 431460.CrossRefGoogle Scholar
Donoso, C. 2008. Árboles Nativos de Chile. Valdivia: Marisa Cuneo Ediciones.Google Scholar
Donoso, C., Sandoval, V. & Grez, R. 1990. Silvicultura de los bosques de Fitzroya cupressoides: Ficción o realidad ? Bosque 11: 5767.CrossRefGoogle Scholar
Donoso, C., Sandoval, V., Grez, R. & Rodríguez, J. 1993. Dynamics of Fitzroya cupressoides forests in southern Chile. Journal of Vegetation Science 4: 303312.CrossRefGoogle Scholar
Donoso, C., Lara, A., Escobar, B. Premoli, A & Souto, C. 2006. Fitzroya cupressoides (Molina) I.M.Johnston. Autoecologia 2006: 6881.Google Scholar
Dörken, V.M. & Stützel, T. 2011. Morphology and anatomy of anomalous cladodes in Sciadopitys verticillata Siebold & Zucc. (Sciadopityaceae). Trees 25: 199213.CrossRefGoogle Scholar
Dörken, V.M. & Stützel, T. 2012. Morphology, anatomy and vasculature of leaves in Pinus (Pinaceae) and their evolutionary meaning. Flora 207: 5762.CrossRefGoogle Scholar
Dorman, K.W. & Zobel, B.J. 1973. Genetics of Loblolly Pine. Washington, DC: USDAGoogle Scholar
Dorofeev, P.I. & Sveshnikova, I.N. 1959. On the discovery of the remains of the genus Sciadopitys S. & Z. in the Upper Cretaceous deposits of the Urals. Doklady Akademii Nauk S.S.S.R. (Earth Sciences) 128: 10141016 (in Russian).Google Scholar
Doweld, R.B. 2000. Botryopitys, a new generic name (Podocarpopsida). Turczaninowia 3(4): 363367.Google Scholar
Downie, D.G. 1923. Chinese species of Tsuga. Notes from the Royal Botanic Garden, Edinburgh 67: 1319.Google Scholar
Doyle, J. 1945. Developmental lines in pollination mechanisms in the Coniferales. Science Proceedings of the Royal Society of Dublin 24: 4362.Google Scholar
Doyle, J. 1954. Development in Podocarpus nivalis in relation to other Podocarps. III. General conclusions. Science Proceedings of the Royal Society of Dublin 26: 347377.Google Scholar
Doyle, J. & Brennan, S.J. 1971. Cleavage polyembryony in conifers and taxads: a survey. 1. Podocarps, taxads and taxodioids. Science Proceedings of the Royal Society of Dublin 4: 5788.Google Scholar
Doyle, J. & O’Leary, M. 1935. Pollination in Saxegothaea. Science Proceedings of the Royal Society of Dublin 21: 175179.Google Scholar
Doyle, J.A. 1998. Molecules, morphology, fossils, and the relationship of angiosperms and gnetales. Molecular Phylogenetics and Evolution 9: 448462.CrossRefGoogle Scholar
Doyle, J.A. & Donoghue, M.J. 1987. The importance of fossils in elucidating seed plant phylogeny and macroevolution. Review of Palaeobotany and Palynology 50: 6395.CrossRefGoogle Scholar
Doyle, J.A. & Donoghue, M.J. 1992. Fossils and seed plant phylogeny reanalyzed. Brittonia 44: 89106.CrossRefGoogle Scholar
Doyle, J.A. & Hotton, C.L. 1991. Diversification of early angiosperm pollen in a cladistic context. In Blackmore, S. & Barnes, S.H. (eds.), Pollen and Spores: Patterns of Diversification. Oxford: Clarendon Press.Google Scholar
Doyle, J.A. & Zimmer, E.A. 1994. Integration of morphological and ribosomal RNA data on the origin of angiosperms. Annals of the Missouri Botanical Garden 81: 419450.CrossRefGoogle Scholar
Doyle, M.F. 1999 . Species accounts: Fiji acmopyle (Acmopyle sahniana J. Buchholz and N.E. Gray). Pp 99100 in Farjon, A. & Page, C.N. (eds.), Conifers:. Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Doyle, M.F. 1999. Regional action plan: conifers of the oceanic islands of the Insular South Pacific (Fiji, Tonga, Solomon Islands and Vanuatu). Pp 7274 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Du, B.-X., Yan, D.-F., Sun, B.-N., et al. 2012. Cunninghamia praelanceolata sp. nov. with associated epiphyllous fungi from the upper Miocene of eastern Zhejiang, S.E. China and their palaeoecological implications. Review of Palaeobotany and Palynology 182: 3243.CrossRefGoogle Scholar
Duarte, L. 1993. Fossils of Araucariaceae from the Santana Formation: Crato Member (Aptian), northeastern Brazil. Anais da Academia Brasileira de Ciências 65:357362.Google Scholar
Dubois-Laduranrie, G. 1941. Revision de quelques especes d’Abietinees fossiles. Bulletin de la Société d’Historire Naturelle de Toulouse 76: 363402.Google Scholar
Duffield, J.W. & Righter, F.I. 1953. Annotated list of pine hybrids made at the Institute of Forest Genetics. US Department of Agriculture, Forest Service, Research Notes 86: 19.Google Scholar
Duigan, S.L. 1951. A catalogue of the Australian Tertiary flora. Proceedings of the Royal Society of Victoria 53: 4156.Google Scholar
Dunham, E., Overall, K.I., Porter, W.P. & Forster, C.A. 1989. Implications of ecological energetics and biophysical and developmental constraints for life-history variation in dinosaurs. Pp 119 in Farlow, J. (ed.), Palaeobiology of the Dinosaurs. Boulder, CO: Geological Society of America.Google Scholar
Dunn, S.T. & Tutcher, W.J. 1912. Flora of Kwangtung and Hong Kong (China). Kew Bulletin Miscellaneous Information, Additional Series 10: 1370.Google Scholar
Dupler, A.W. 1920. Ovuliferous structures of Taxus canadensis. Botanical Gazette 69: 492520.CrossRefGoogle Scholar
Dupont-Nivet, G., Krigsman, W. Langereis, C.G., et al. 2007. Tibetan Plateau aridification linked to global cooling at the Eocene–Oligocene transition. Nature 445: 635638.CrossRefGoogle ScholarPubMed
Dupuy, L., Forcaud, T., Stokes, A. & Danjon, F. 2005. A density-based approach for the modelling of root architecture: application to maritime pine (Pinus pinaster Ait.) root systems. Journal of Theoretical Biology 236: 323334.CrossRefGoogle ScholarPubMed
Durrieu, G. 1980. Phylogeny of Uredinales on Pinaceae. Reports of the Tottori Mycological Institute 18: 283290.Google Scholar
Dvorak, W.S. & Raymond, R.H. 1990. The taxonomic status of closely related closed cone pines in Mexico and Central America. New Forests 4: 291307.CrossRefGoogle Scholar
Dvorak, W.S., Hamrick, J.L. & Gutierrez, E.A. 2005. The origin of Caribbean pine in the seasonal swamps of the Yucatan. International Journal of Plant Science 166: 985994.CrossRefGoogle Scholar
Dy, F., Wu, B. & Wang, X. 1990. A preliminary study on the phenology of Taxus cuspidata Sieb. et Zucc. Journal of the Northeast Forestry University 7: 1116.Google Scholar
Dyer, A.F. & Page, C.N. (eds.). 1982. The Biology of Pteridophytes. Edinburgh: Royal Society of Edinburgh.Google Scholar
Eames, A.J. 1913. The morphology of Agathis australis. Annals of Botany 27: 138.CrossRefGoogle Scholar
Eames, A.J. 1952. Relationships of the Ephedrales. Phytomorphology 2: 79100.Google Scholar
Eckenwalder, J.E. 1976. Comments on ‘ A new classification of the conifers’. Taxon 25: 338339.CrossRefGoogle Scholar
Eckert, A.J. & Hall, B.D. 2006. Phylogeny, historical biogeography, and patterns of diversification for Pinus (Pinaceae): phylogenetic tests of fossil-based hypotheses. Molecular Phylogenetics and Evolution 40: 166182.CrossRefGoogle ScholarPubMed
Eckman, S.R. 1998. Pleistocene pollen stratigraphy from borehole 81/34, Devil’s Hole area, central North Sea. Quaternary Science Reviews 17: 855869.CrossRefGoogle Scholar
Edlin, H.L. 1958. The Living Forest. London: Thames & Hudson.Google Scholar
Eguiluz-Piedra, T. 1986. Taxonomic relationships of Pinus tecunumanii from Guatemala. Commonwealth Forestry Review 65: 303313.Google Scholar
Ehrendorfer, F. 1973. Evolutionary significance of chromosomal differentiation patterns in gymnosperms and primitive angiosperms. Pp 220240 in Beck, C.B. (ed.), Origin and Early Evolution of Angiosperms. New York: Columbia University Press.Google Scholar
Ehrendorfer, F., Krendl, F., Habeler, E. & Sauer, W. 1968. Chromososme numbers and evolution in primitive angiosperms. Taxon 17: 337353.CrossRefGoogle Scholar
Ehrenfeld, J.G. 1986. Wetlands of the New Jersey Pine Barrens: the role of species composition in community function. American Midland Naturalist 115: 301313.CrossRefGoogle Scholar
Ekanayake, J.C., Marden, M., Watson, A.J. & Rowan, D. 1997. Tree roots and slope stability: a comparison between Pinus radiata and kanuka. New Zealand Journal of Forest Science 27: 216233.Google Scholar
El-Lakany, M.H. & Dugle, J.R. 1972. DNA content on relation to phylogeny of selected boreal forest plants. Evolution 26: 427434.CrossRefGoogle ScholarPubMed
Ellenberg, H. 1988. Vegetation Ecology of Central Europe. Cambridge: Cambridge University Press.Google Scholar
Elliot, M.B. 1998. Late Quaternary pollen records of vegetation and climate change from Laitaia Bog, far northern New Zealand. Review of Palaeobotany and Palynology 99: 189202.CrossRefGoogle Scholar
Elliott, C.G. 1951. Some notes on Athrotaxis. Proceedings of the Linnean Society of New South Wales 76: 3640.Google Scholar
Elton, C.S. 1996. The Pattern of Animal Communities. London: Chapman and Hall.Google Scholar
Elwes, H.J. & Henry, A. 1912. The Trees of Great Britain and Ireland. Edinburgh, privately published.Google Scholar
Emberger, L. 1954. Sur les Ginkgoales et quelques rapprochements avec d-autres groupes systematiques. Svensk Botaniusker Tiodskrift 48: 361367.Google Scholar
Endo, S. A Neogene species of Sequoia from Japan. Botanical Gazette 94: 605610.CrossRefGoogle Scholar
Endo, S. & Okutsu, H. 1936. Glyptostrobus cone from the Liriodendron beds near the Sendai. Proceedings of the Imperial Academy of Tokyo 12: 138140.CrossRefGoogle Scholar
Engelmann, G.F., Chure, D.J. & Fiorillo, A.R. 2004. The implications of a dry climate for the paleoecology of the fauna of the Upper Jurassic Morrison formation. Sedimentary Geology 167: 297308.CrossRefGoogle Scholar
Enoki, T., Kawaguchi, H. & Iwatsubo, G. 1996. Topographic variation of soil properties and stand structure in a Pinus thunbergii plantation. Ecological Research 11: 299309.CrossRefGoogle Scholar
Ennos, R.A., Sinclair, W.T., & Perks, M.P. 1997. Genetic insights into the evolution of Scots pine, Pinus sylvestris L., in Scotland. Botanical Journal of Scotland 49: 257265.CrossRefGoogle Scholar
Ennos, R.A., Sinclair, W.T., Hu, X.-S., & Langdon, A. 1999. Using organelle markers to elucidate the history, ecology and evolution of plant populations. Pp 119 in Hollingsworth, P.M., Bateman, R.M. & Gornall, R.J. (eds.), Molecular Systematics and Plant Evolution. London: Taylor & Francis.Google Scholar
Enright, N.J. 1982. Does Araucaria hunsteinii compete with its neighbours? Australian Journal of Ecology 7: 9799.CrossRefGoogle Scholar
Enright, N.J. 1982. The Araucariaceae forests of New Guinea. Pp 382399 in Gressitt, J.L. (ed.), Biogeography and Ecology of New Guinea. The Hague: W.Junk.Google Scholar
Enright, N., Bartlett, R.M. & De Fretes, C.R. 1993. Patterns of species composition, recruitment and growth within canopy gaps in two New Zealand kauri (Agathis australis) forests. New Zealand Journal of Botany 31: 361373.CrossRefGoogle Scholar
Enright, N.J., Ogden, J. & Rigg, L.S. 1999. Dynamics of forest with Araucariaceae in the western Pacific. Journal of Vegetation Science 10: 793804.CrossRefGoogle Scholar
Erdtman, G.E. 1957. Pollen and Spore Morphology/Plant Taxonomy: Gymnospermae. Pteridophyta. Bryophyta. New York: Ronald Press Co.Google Scholar
Erdtman, H. 1956. Organic chemistry and conifer taxonomy. Pp 453494 in Todd, A. (ed.), Perspectives in Organic Chemistry. New York: Interscience Publishers.Google Scholar
Escapa, I., Cúneo, N.R. & Cladera, G. 2008. New evidence for the age of the Jurassic Flora from Cañadón del Zanio, Sierra de Taquetrén, Chubut. Ameghiniana 45: 633637.Google Scholar
Escapa, I., Cúneo, R. & Axsmith, B. 2008. A new genus of Cupressaceae (sensu lato) from the Hurassi of Patagonia: implications for conifer megasporangiate cone homologies. Review of Palaeobatany & Palynology 151: 110122.CrossRefGoogle Scholar
Espelta, J.M., Retana, J. & Habrouck, A. 2003. An economic and multi-criteria evaluation of reforestation methods to recover burned Pinus nigra forests in NE Spain. Forest Ecology and Management 180: 185198.CrossRefGoogle Scholar
Espinosa, M., Acuña, E., Cancino, J. Munoz, F. & Perry, D.A. 2005. Carbon sink potentials of radiata pine plantations in Chile. Forestry 78: 1119.CrossRefGoogle Scholar
Ewers, B.E., Oren, R., Albaugh, T.J. & Dougherty, P.M. 1999. Carry-over effects of water and nutrient supply on water use of Pinus taeda. Ecological Applications 9: 513525.CrossRefGoogle Scholar
Faegri, K. & van der Pijl, L. 1979. The Principles of Pollination Ecology. Oxford: Pergamon Press.Google Scholar
Fahey, T.J., Battles, J.J. & Wilson, G.F. 1988. Responses of early successional northern hardwood forests to changes in nutrient availability. Ecological Monographs 68: 183212.CrossRefGoogle Scholar
Fajardo, A. & González, M.E. 2009. Replacement patterns and species coexistence in an Andean AraucariaNothofagus forest. Journal of Vegetation Science 20: 11761190.CrossRefGoogle Scholar
Falcon-Lang, H.J. 2003. Do tree-rings in fossil woods give a palaeoclimate signal? International Association of Wood Anatomists Journal 24: 316.Google Scholar
Falcon-Lang, H.J. & Cantrill, D.J. 2001. Gymnosperm woods from the Cretaceous (Mid Aptian) Cerro Negro Formation, Byers Peninsula, Livingstone Island, Antarctic: the arborescent vegetation of a volcanic arc. Cretaceous Research 22: 277293.CrossRefGoogle Scholar
Falcon-Lang, H.J. & Cantrill, D.J. 2001. Leaf phenology of some mid-Cretaceous polar forests, Alexander Island, Antarctica. Geological Magazzine 138: 3952.CrossRefGoogle Scholar
Falcon-Lang, H.J., Cantrill, D.J. & Nichols, G.J. 2001. Biodiversity and terrestrial ecology of a mid-Cretaceous, high-latitude floodplain, Alexander Island, Antarctica. Journal of the Geological Society of London 158: 709725.CrossRefGoogle Scholar
Falcon-Lang, H.J., Kvaček, J. & Uličný, D. 2001. Fire-prone plant communities and palaeoclimate of a Late Cretaceous fluvial to estuarine environment, Pecínov Quarry, Czech Republic. Geological Magazine 138: 563576.CrossRefGoogle Scholar
Fallour, D.B. & Lefevre, F. 2001. Evidence of variation in segregation patterns within a Cedrus population. Journal of Heredity 92: 260266.CrossRefGoogle ScholarPubMed
Farjon, A. 1989. Biodiversity of Pinus (Pinaceae) in Mexico: speciation and paleo-endemism. Botanical Journal of the Linnean Society 121: 365385.Google Scholar
Farjon, A. 1990. Pinaceae: Drawings and Descriptions of the Genera Abies, Cedrus, Pseudolarix, Keteleeria, Nothotsuga, Tsuga, Cathaya, Pseudotsuga, Larix and Picea. Konigstein: Koeltz Scientific Books.Google Scholar
Farjon, A. 1992. The taxonomy of multiseed Junipers (Juniperus sect. Sabina) in southwest Asia and east Africa. Edinburgh Journal of Botany 49: 251283.CrossRefGoogle Scholar
Farjon, A. 1993. Nomenclature of the Mexican cypress or Cedar of Goa, Cupressus lusitanica Mill. (Cupressaceae). Taxon 42: 8184.CrossRefGoogle Scholar
Farjon, A. 1999. Species accounts: bigcone pinyon pine (Pinus maximartinezii Rzed.). Pp 101102 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Farjon, A. 2002. Extensions to the natural range of Taiwania cryptomerioides. Fitzroya 5: 4.Google Scholar
Farjon, A. 2007. In defence of a conifer taxonomy that recognises evolution. Taxon 56: 639641.CrossRefGoogle Scholar
Farjon, A. & Mill, R.R. 1998. Proposals to conserve the current spelling of two generic names of conifers. Taxon 48: 151154.CrossRefGoogle Scholar
Farjon, A. & Oritz García, S. 2005. The early development of ovuliferous cones in Cupressaceae s.lat.: a survey of the genera. Pp 2746 in Farjon, A. (ed.), A Monograph of Cupressaceae and Sciadopitys. London: Royal Botanic Gardens, Kew.Google Scholar
Farjon, A. & Page, C.N. (eds.). 1999. Conifers: Status Survey and Conservation Action Plan. Gland: IUCN.Google Scholar
Farjon, A. & Page, C.N. 2003. An action plan for the world’s conifers. Plant Talk 22: 4347.Google Scholar
Farjon, A. & Page, C.N. 2003. Diversity in strategies for conifer conservation, the action plan and future developments. Pp 397403 in Mill, R.R. (ed.), Proceeding of the Fourth International Conifer Conference: Conifers for the Future. Brugge: International Society for Horticultural Science.Google Scholar
Farlow, J.O. 1987. Speculations about the diet and digestive physiology of herbivorous dinosaurs. Paleobiology 13: 6072.CrossRefGoogle Scholar
Felsenstein, J. 1985. Confidence limits on phylogenies: an approach using the bootstrap. Evolution 39: 783791.CrossRefGoogle ScholarPubMed
Feng, X. & Epstein, S. 1994. Climatic implications for an 8000-year hydrogen isotope time series from Bristlecone pine trees. Science 265: 10791081.CrossRefGoogle ScholarPubMed
Ferguson, D.K. 1971. The Miocene Flora of Kleuzau, Western Germany. 1. The Leaf Remains. Amsterdam: North Holland.Google Scholar
Ferguson, D.K. 1985. The origin of leaf assemblages: new light on an old problem. Review of Palaeobotany and Palynology 46: 117188.CrossRefGoogle Scholar
Fernández, P.M. & Rigolot, E. 2007. The fire ecology and management of maritime pine (Pinus pinaster Ait.). Forest Ecology and Management 241: 113.CrossRefGoogle Scholar
Ferré, M.Y. Rouane, M.L. & Woltz, P. 1975. Structure des plantules et systématique du genre Podocarpus. Travaux Laboratoire Forestiere Toulouse 2, 116.Google Scholar
Ferré, Y. & Gaussen, H. 1945. Le remeau phylettique: Pinus, Pseudolarix, Keteleeria. Bulletin de la Société d’Histoire Naturelle de Toulouse 80: 10118.Google Scholar
Ferréira, A.G. & Handro, W. 1979. Aspects of seed germination in Araucaria angustifolia (Bert.) O. Ktze. Revista Braziliera de Botanica 2: 713.Google Scholar
Field, T.S. & Arens, N.C. 2005. Form, function and environments of the early angiosperms: merging extant phylogeny and ecophysiology with fossils. New Phytologist 166: 383408.CrossRefGoogle Scholar
Fiorillo, A.R. 1998. Dental microwear patterns from the sauropod dinosaurs Camarasaurus and Diplodocus: evidence for resource partitioning in the Late Jurassic of North America. Historical Biology 13: 116.CrossRefGoogle Scholar
Fischer, T.C., Meller, B., Kustatscher, E. & Butzmann, R. 2010. Permian Ginkgophyta fossils from the Dolomites resemble extant O-ha-tsuki aberrant leaf-like fructifications of Ginkgo biloba L. BMG Evolutionary Biology 10: 317.Google ScholarPubMed
Fitzgerald, P.G., Sandiford, M., Barrett, P.J. & Gleadow, A.J.W. 1986. Asymmetric extension associated with uplift and subsidence in the Transantarctic Mountains and Ross Embayment. Earth and Planetary Science Letters 81: 6778.CrossRefGoogle Scholar
Fitzpatrick, H.M. 1929. Coniferae: keys to the genera and species, with economic notes. Science Proceedings of the Royal Society of Dublin 19: 189260.Google Scholar
Fjeldsa, J. & Lovett, J.C. 1997. Geographical patterns of old and young species in African forest biota: the significance of specific montane areas as evolutionary centres. Biodiversity and Conservation 6: 325346.CrossRefGoogle Scholar
Fleming, C.A. 1962. New Zealand biogeography. A paleontologist’s approach. Tuatara 10: 53108.Google Scholar
Florence, L.Z. & Hicks, R.R. Jr. 1980. Further evidence for introgression of Pinus taeda with P. echinata: electrophoretic variability and variation in resistance to Cronartium fusiforme. Silvae Genetica 29: 4143.Google Scholar
Florin, R. 1930. Pilgerodendron, eine neue Koniferen-gattung aus Sud-Chile. Svensk Botanisk Tidskrift 24: 132135.Google Scholar
Florin, R. 1931. Untersuchungen zur Stammesgeschchte der Conifereales und Cordaitale, 1. teil: Morphologie und Epidermisstructur der Assimilationsorgane bei rezenten Koniferen. Kungliga Svenska Vetaenskapsacademiens Handlingar 10: 1558.Google Scholar
Florin, R. 1940. The Tertiary fossil conifers of southern Chile and their phytogeographical significance. K. Svenska Vetenskaps Akademie Handl 19(2): 1107.Google Scholar
Florin, R. 1940. Die heutige und frühere Verbreitung der Koniferengattung Acmopyle Pilger. Svensk Botanisk Tidskrift 34: 117140.Google Scholar
Florin, R. 1944. Die Koniferen des Oberkarbon und unteren Perms. Morphologie der weiblichen reproduktionsorgane der fossilen Cordaitales, fossilen und recenten Coniferales und taxales. Alaeontographica 85: 457654.Google Scholar
Florin, R. 1948. On the morphology and relationship of the Taxaceae. Botanical Gazette 110: 3139.CrossRefGoogle Scholar
Florin, R. 1950. Upper Carboniferous and Lower Permian conifers. Botanical Review 16: 258282.CrossRefGoogle Scholar
Florin, R. 1951. Evolution of cordaites and conifers. Acta Horticultura Bergiani 15: 285388.Google Scholar
Florin, R. 1956. Nomenclatural notes on genera of living gymnosperms. Taxon 5: 188192.CrossRefGoogle Scholar
Florin, R. 1956. Nomenclatural notes on living gymnosperms. Taxon 5: 186192.CrossRefGoogle Scholar
Florin, R. 1957. Notes on the systematics of the Podoarpaceae. Acta Horti Bergiani 17: 403411.Google Scholar
Florin, R. & Boutelje, J.B. 1954. External morphology and epidermal structure of leaves in the genus Libocedrus s. lat. Acta Horti Bergiani 17: 737.Google Scholar
Florschutz, F. 1925. On Pseudolarix kaempferi Gord: from the clay of Reuver. Recueil des Travaux Botaniques Néerlandais 22: 269274.Google Scholar
Flous, F. 1936. Especes nouvelles de Keteleeria. Bulletin de la Société d’Histoire Naturelle de Toulouse 69: 399408.Google Scholar
Flous, F. 1936. Revision de genre Keteleeria. Bulletin de la Société d’Histoire Naturelle de Toulouse 70: 273346.Google Scholar
Flynn, J.J., Parrish, J.M., Rakotosamimanana, B., Simpson, W.F. & Wyss, A.R. A Middle Jurassic mammal from Madagascar. Nature 401: 5760.CrossRefGoogle Scholar
Foley, J.A., Kutzback, J.E., Coe, M.T. & Levis, S. 1994. Feedbacks between climate and boreal forests during the Holocene Epoch. Nature 371: 52054.CrossRefGoogle Scholar
Fonda, R.W. 2001. Burning characteristics of needles from eight pine species. Forest Science 27: 390396.CrossRefGoogle Scholar
Fonda, R.W. & Varner, J.M. III. 2005. Burning characteristics of cones from eight pine species. Northwest Science 78: 322333.Google Scholar
Foreman, D.B. 1971. A Checklist of the Vascular Plants of Bougainville with Descriptions of some Common Forest Trees. Lae, Papua New Guinea: Division of Botany, Department of Forests.Google Scholar
Forrest, G. 1924–1925. Field notebooks. (Original handwritten field notebooks, housed in the library of Royal Botanic Garden Edinburgh, UK.)Google Scholar
Forster, B.A. & Baker, D.E. 1997. Characterisation of the serpentine soils of central Queensland, Australia. Pp 2737 in Jaffré, T., Reeves, R.D. & Bacquer, T. (eds.), The Ecology of Ultramafic and Metalliferous Areas. Noumea, New Caledonia: Centre ORSTOM.Google Scholar
Fowler, A.M., Palmer, J. & Fenwick, P. 2008. An assessment of the potential for centennial-scale reconstruction of atmospheric circulation from selected New Zealand tree-ring chronologies. Palaeogeography, Palaeoclimatology, Palaeoecology 265: 238254.CrossRefGoogle Scholar
Fowler, D.P. & Heimburger, C. 1958. The hybrid Pinus peuce Griesb. x Pinus strobus L. Silvae Genetica 7: 8186.Google Scholar
Francis, J.E. 1986. Growth rings in Cretaceous and Tertiary wood from Antarctica and their palaeoclimatic implications. Palaeontology 29: 665684.Google Scholar
Francis, J.E. 1983. The dominant conifer of the Jurassic Purbeck Formation, England. Palaeontology 26: 277294.Google Scholar
Francis, J.E. 1988. A 50-million year old fossil forest from Strathcona Fiord, Ellesmere Island, Arctic Canada: evidence for a warm polar climate. Arctic 41: 314318.CrossRefGoogle Scholar
Francis, J.E. & Hill, R.S. 1996. Pliocene fossil plants from the Transantarctic Mountains: evidence for climate from growth rings and fossil leaves. Palaios 11: 389396.CrossRefGoogle Scholar
Francis, J.E., Pirrie, D. & Crame, J.A. (eds.). 2006. Cretaceous–Tertiary High Latitude Palaeoenvironments, James Ross Basin, Antarctica. London: Geological Society of London.CrossRefGoogle Scholar
Francis, W.D. 1970. Australian Rain-Forest Trees, 3rd edn. Canberra: Australian Government Publications Service.Google Scholar
Franco, M. & Silvertown, J. 2000. Comparative demography of plants based upon elasticities of vital rates. Ecology 85: 531538.CrossRefGoogle Scholar
Frankel, O.H. & Soule, M.E. 1981. Conservation and Evolution. Cambridge: Cambridge University Press.Google Scholar
Frankis, M.P. 1988. Generic interrelationships in Pinaceae. Notes from the Royal Botanic Garden, Edinburgh 45: 527548.Google Scholar
Frankis, M.P. 1993. Morphology and affinities of Pinus brutia. Pp 1118 in Secretariat of the Organising Committee (ed.), International Symposium on Pinus brutia. Ankara: Ministry of Forestry.Google Scholar
Frankis, M.P. 1993. Nootka cypress: Chamaecyparis or Cupressus? Conifer Society of Australia Newsletter 12: 910.Google Scholar
Friedman, W. & Goliber, T.E. 1986. Photosynthesis in the female gametophyte of Ginkgo biloba. American Journal of Botany 73: 12611266.CrossRefGoogle Scholar
Friis, E.M., Chaloner, W.G. & Crane, P.R. (eds.). 1987. The Origins of Angiosperms and their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Frith, A.C. 1966. Araucaria angustifolia in Argentina. Institute Forestal Latino Americano Boletin Téchnica 22.Google Scholar
Froyd, C.A. 2005. Fossil stomates reveal early pine presence in Scotland: implications for postglacial colonisation analyses. Ecology 86: 579586.CrossRefGoogle Scholar
Fu, D.Z. 1992. Nageiaceae: a new gymnosperm family. Acta Phytotaxonomica Sinica 30: 515528.Google Scholar
Fu, L.G. & Jin, J.M. 1992. China Plant Red Data Book. Beijing: Science Press.Google Scholar
Fu, L.-G., Yu, Y.-F. & Farjon, A. 1999. Cupressaceae. Pp 6277 in Flora of China, Vol 4. Beijing: Science Press and Missouri Botanical Garden.Google Scholar
Fu, L.-K., Mill, R.R. & Turland, N.J. 1999. Validation of the name Cephalotaxus latifolia (Cephalotaxaceae), a species from southeast China. Novon 9: 185186.Google Scholar
Fukarek, P. 1950. [The recent area of distribution of P. omorika and some notes on its 1950 stands]. Godišnjak. Biological Institue, Sarajevo 3: 140198. (in Croatian).Google Scholar
Fukimroi, T. 1977. Stem biomass and structure of a mature Sequoia sempervirens stand on the Pacific coast of northern California. Journal of the Japanese Forestry Society 59: 435441.Google Scholar
Furman, B.J., Garattapaglia, D., Dvorak, W.S., & O’Malley, D.M. 1997. Analysis of genetic relationships of Central American and Mexican pines using RAPD markers that distinguish species. Molecular Ecology 6: 321331.CrossRefGoogle Scholar
Furman, T.E. 1970. The nodular mycorrhizas of Podocarpus rospigliosii. American Journal of Botany 37: 910913.CrossRefGoogle Scholar
Furnier, G.R. & Adams, W.T. 1986. Geographic patterns of allozyme variation in Jeffrey pine. American Journal of Botany 73: 10091015.CrossRefGoogle Scholar
Galbraith, D.W., Harkins, K.R., Maddox, J.M., et al. 1983. Rapid flow cytometric analysis of the cell cycle in intact plant tissues. Science 220: 10491051.CrossRefGoogle ScholarPubMed
Galtier, J., Scott, A.C., Powell, J.H., Glover, B.W. & Waters, C.N. 1992. Anatomically preserved conifer-like stems from the Upper Carboniferous of England. Proceedings of the Royal Society of London B 247: 211214.Google Scholar
Gamerro, J.C. 1995. Morfología del pollen de Saxegothaea conspicua (Podocarpaceae). Darwiniana 33: 295300.Google Scholar
García, D. & Obeso, J.R. 2003. Facilitation by herbivore mediated nurse plants in a threatened tree, Taxus baccata: local effects and landscape level consistency. Ecogeography 26: 739750.CrossRefGoogle Scholar
García, D., Zamora, R., Gómez, J.M., Jordano, P. & Hodar, J.A. 2000. Geographical variation in seed production, predation and abortion in Juniperus communis throughout its range in Europe. Journal of Ecology 88: 439446.CrossRefGoogle Scholar
García, D., Zamora, R., Gómez, J.M. & Hodar, J.A. 2001. Frugivory at Juniperus communis depends more on population characteristics than on individual attributes. Journal of Ecology 89: 639646.CrossRefGoogle Scholar
García, S., Finch, D.M. & Leon, G.C. 1998. Patterns of forest use and endemism in resident bird communities of north-central Michoacan, Mexico. Forest Ecology and Management 110: 151171.CrossRefGoogle Scholar
Garden, J. 1956. A revision of the genus Callitris Vent. Contributions to the New South Wales National Herbarium 2: 363392.Google Scholar
Garden, J. & Johnson, A.S. 1950. Microstrobos, a new name for a podocarpaceous genus. Contributions to the New South Wales National Herbarium 1: 519520.Google Scholar
Gardes, M. & Bruns, T.D. 1996. Commmunity structure of ectomycorrhizal fungi in a Pinus muricata forest: above- and below-ground views. Canadian Journal of Botany 74: 15721583.CrossRefGoogle Scholar
Gardner, H.M. 1926. East African pencil cedar. Empire Forestry Journal 5: 3946.Google Scholar
Gardner, J.S. 1886. A Monograph of the British Eocene Flora. II. Gymnospermae. London: British Museum (Natural History).CrossRefGoogle Scholar
Gardner, M. F. 2019. Threatened Conifers of the World. https://threatenedconifers.rbge.org.uk/conifers/prumnopitys-andinaGoogle Scholar
Garnett, G.N., Chambers, C.L. & Mathiasen, R.L. 2006. Use of witches’ brooms by Albert squirrels in ponderosa pine forests. Wildlife Society Bulletin 34: 467472.CrossRefGoogle Scholar
Garwood, N.C., Janos, D.P. & Brokaw, N. 1979. Earthquake-caused landslides: a major disturbance to tropical forests. Science 205: 997999.CrossRefGoogle Scholar
Gaussen, H. & Lacassagne, M. 1930. Les epiceas du Kamtchatka. Société d’Histoire Naturelle Toulouse Bulletin 59: 190202.Google Scholar
Gauthier, S., Bergeron, Y. & Simon, J.P. 1993. Cone serotiny in jack pine: ontogenetic, positional and environmental effects. Canadian Journal of Forest Research 23: 394401.CrossRefGoogle Scholar
Gauthier, S., Bergeron, Y. & Simon, J.P. 1996. Effects of fire regime on the serotiny level of jack pine. Journal of Ecology 84: 539548.CrossRefGoogle Scholar
Gayo, E., Hinojosa, L.F. & Villagran, C. 2005. On the persistence of tropical paleofloras in central Chile during the early Eocene. Review of Palaeobotany and Palynology 137: 4150.CrossRefGoogle Scholar
Ge, X.J., Zhou, X.L., Li, Z.C., et al. 2005. Low genetic diversity and significant populations structuring in the relict Amentotaxus argotaenia complex (Taxaceae) based on ISSR fingerprinting. Journal of Plant Research 118: 415422.CrossRefGoogle ScholarPubMed
Gee, C.T. 1989. Permian Glossopteris and Elatocladus megafossil floras from the English coast, Ellsworth Land, Antarctica. Antarctic Science 1: 3544.CrossRefGoogle Scholar
Gee, C.T. 2009. Dietary options of the sauropods. Journal of Vertebrate Paleontology 29: 3456.Google Scholar
Gee, C.T. 2011. Dietary options for the sauropod dinosaurs from an integrated botanical and paleobotanical perspective. Pp 3456 in Klein, N., Remes, K., Ge, C.T. & Sander, P.M. (eds.), Biology of the Sauropod Dinosaurs: Understanding the Life of Giants. Bloomington, IN: Indiana University Press.Google Scholar
Gee, C.T. 2011. Sauropod herbivory during Late Jurassic times: new evidence for conifer-dominated vegetation in the Morrison Formation of the western interior of North America. Journal of Vertebrate Paleontology 31: 115.Google Scholar
Gee, C.T. & Tidwell, W.D. 2010. A revision of characters of a new whole-plant araucarian A. delevoryasii gen et sp. nov. from the Late Jurassic Morrison Formation of Wyoming, USA. Pp 6694 in Gee, C.T. (ed.), Plants of Mesozoic Time: Morphological Innovations, Phylogeny Ecosystems. Bloomington, IN: Indiana University Press.Google Scholar
Gernandt, D.S., Liston, A. & Piñero, D. 2001. Variation in the nrDNA ITS of Pinus subsection Cembroides: implications for molecular systematic studies of pine species complexes. Phylogenetics and Evolution 21: 450468.Google ScholarPubMed
Gernandt, D.S., Liston, A. & Piñero, D. 2003. Phylogenetics of Pinus subsections Cembroides and Nelsoniae inferred from cpDNA sequences. Systematic Botany 4: 657673.Google Scholar
Gervais, B.R. & MacDonald, G.M. 2000. A 403-year record of July temperatures and treeline dynamics of Pinus sylvestris from the Kola Peninsula, northwest Russia. Arctic, Antarctic and Alpine Research 32: 295302.CrossRefGoogle Scholar
Ghandi, S.G., Mahajan, V. & Bedi, Y.E. 2015. Changing trends in biotechnology of secondary mechanisms in medicinal and aromatic plants. Planta 241: 303317.Google Scholar
Ghazoul, J. 2005. Pollen and seed dispersal amongst dispersed plants. Biological Review 80: 413443.CrossRefGoogle Scholar
Gibbs, G. 2006. Ghosts of Gondwana: The History of Life in New Zealand. Nelson: Craig Potton Publishing.Google Scholar
Gibbs, L.S. 1912. On the development of the female strobilus in podocarps. Annals of Botany 26: 525572.Google Scholar
Gibson, N. & Brown, M.J. 1991. The ecology of Lagarostrobos franklinii (Hook. f.) Quinn (Podocarpaceae) in Tasmania. 2. Population structure and spatial pattern. Australian Journal of Ecology 6: 223229.CrossRefGoogle Scholar
Gibson, N., Brown, M.J., Williams, K. & Brown, A.V. 1992. Flora and vegetation of ultramafic areas in Tasmania. Australian Journal of Ecology 17: 297303.CrossRefGoogle Scholar
Gibson, Q.L. & Barnes, R.D. 1984. Status of the international provenance trials of Pinus keysia and alternatives for future development. Pp. 214218 in Barnes, R.D. & Gibson, Q.L. (eds.), Provenance and Genetic Improvement Strategies in Tropical Forest Trees. Oxford: Commonwealth Forestry Institute.Google Scholar
Gillespie, W.H., Rothwell, G.W. & Scheckler, S.E. 1981. The earliest seeds. Nature 293: 462464.CrossRefGoogle Scholar
Gilliam, F.S. 2007. The ecological significance of the herbaceous layer in temperate forest ecosystems. BioScience 57: 845858.CrossRefGoogle Scholar
Gimmingham, C.H. 1977. The status of pinewoods in British ecosystems. Pp 14 in Bunce, R.G.H. & Jeffers, J.N.R. (eds.), Native Pinewoods of Scotland. Cambridge: Institute of Terrestrial Ecology.Google Scholar
Givinish, T.J. 1980. Ecological constraints on the evolution of breeding systems in seed plants: dioecy and dispersal in gymnosperms. Evolution 34: 959972.CrossRefGoogle Scholar
Givulescu, R. 1972. On some remains of Cunninghamia from the Neogene of Romania. Bulletin de la Société Linnéenne de Lyon 41: 8385.Google Scholar
Givulesu, R. 1990. Flora fosilă a Miocenului superior de las Chiuzbaia, judetul Maramures. Bucharest: Editura Academiei Romana (in Romanian).Google Scholar
Givulesu, R. 1997. Istoria padurilor fosile din Tertairul Transilvaniei, Banatuli, Crisanei si Marmamuresului. Napoca: Editura Carpatica, Cluj (in Romanian).Google Scholar
Gleason, H.A. & Cronquist, A. 1964. The Natural Geography of Plants. New York: Columbia University Press.Google Scholar
Glenn-Lewis, D.C., Peet, R.K. & Veblen, T.T. 1992. Plant Succession: Theory and Prediction. London: Chapman & Hall.Google Scholar
Glitzenstein, J.S., Streng, D.R. & Wade, D.D. 2003. Fire frequency effects on longleaf pine (Pinus palustris P. Miller) vegetation in South Carolina and Northeast Florida. Natural Areas Journal 23: 2237.Google Scholar
Godefroid, S., Rucquoji, S. & Koedam, N. 2006. Spatial variability of summer microclimates and plant species response along transects with clearcuts in a beech forest. Plant Ecology 185: 107121.CrossRefGoogle Scholar
Goldblatt, P. & Manning, J.C. 2002. Plant diversity of the Cape Region of southern Africa. Annals of the Missouri Botanical Garden 89: 281302.CrossRefGoogle Scholar
Golte, W. 1978. Die südandine und die südbrasilianische araukarie. Ein ökologischer Vergleich. Erdkunde 32: 279296.Google Scholar
Gomankov, A.V. & Meyen, S.V. 1980. On relations between assemblages of plant micro- and megafossils in the Permian of Angaraland. Palaeontological Zhurnal 4: 114122 (in Russian).Google Scholar
Gómez, A., González-Martínez, S.C., Collada, C., Climent, J. & Gil, L. 2003. Complex population genetic structure in the endemic Canary Island pine revealed using chloroplast microsatellite markers. Theoretical and Applied Genetics 107: 11231131.CrossRefGoogle ScholarPubMed
Gómez Mendoza, L. & Arriaga, L. 2007. Modelling the effect of climate change on the distribution of oak and pine species in Mexico. Conservation Biology 21: 15451555.CrossRefGoogle ScholarPubMed
Good, R.D. 1948. The Geography of the Flowering Plants. London: Longmans.Google Scholar
Good, R.D. 1958. The biogeography of Australia. Nature, London 181: 17631765.CrossRefGoogle Scholar
Good, R.D. 1964. The Geography of the Flowering Plants, 3rd edn. London: Longman Green.Google Scholar
Gorbunov, M.G. 1957. Remains of the genus Metasequoia in Tertiary deposits of west Siberia. Pp 313326 in A.N. Kristofovich Memorial Collection. Leningrad: Komarov Botanical Institute, Academy of Sciences of the U.S.S.R. (in Russian).Google Scholar
Gorbunov, M.G. & Shikina, I.A. 1972. Metasequoia wood from the Tertiary sediments of west Siberia. Botanicheskii Zhurnal 57: 322330 (in Russian).Google Scholar
Gorelova, S.G. 1978. The flora and stratigraphy of the coal-bearing Carboniferous of middle Siberia. Palaeontographica B 165: 5377.Google Scholar
Goremykin, V., Bobrova, V., Pahnke, J., et al. 1996. Noncoding sequences of the slowly evolving chloroplast inverted repeat in addition to rbcL data do not support gnetalean affinities of angiosperms. Molecular Biology and Evolution 13: 383396.CrossRefGoogle Scholar
Gosling, P.G., Ives, L.M., Cunningham, V.J., et al. 2005. Preliminary advice on fruit handling, seed pre-treatment and germination of embryos of Prumnopitys andina. Sibbaldia 3: 4150.CrossRefGoogle Scholar
Gothan, W. 1936. Nochmals die ‘Graskohle’: Nadeln der Schirmtanne (Sciadopitys). Zeitschrift für Gewinnung und Verwertung der Braunkohle. 35: 736738.Google Scholar
Goubitz, S., Werger, M.J.A. & Ne’-Eman, G. 2003. Germination response to fire-related factors of seeds from non-serotinous and serotinous cones. Plant Ecology 169: 195204.CrossRefGoogle Scholar
Goubitz, S., Nathan, R., Rottenberg, R., Shmida, A. & Ne’-Eman, G. 2004. Canopy seed bank structure in relation to fire, tree size and density. Plant Ecology 173: 191201.CrossRefGoogle Scholar
Gould, R.E. 1968. Morphology of Equisetum laterale Phillips, 1829, and E. bryanii sp. Nov. from the Mesozoic of south-eastern Queensland. Australian Journal of Botany 16: 153176.CrossRefGoogle Scholar
Gould, R.E. & Delevoryas, T. 1977. The biology of Glossopteris: evidence from petrified seed-bearing and pollen-bearing organs. Alcheringa 1: 387399.CrossRefGoogle Scholar
Gould, S.J. & Vrba, E. 1981. Exaptated: a missing term in the science of form. Paleobiology 8: 415.CrossRefGoogle Scholar
Graham, A. (ed.). 1971. Floristics and Paleofloristics of Asian and Eastern North America. Amsterdam: Elsevier.Google Scholar
Graham, A. & Jarzen, D.M. 1969. Studies in neotropical paleobotany. I. The Oligocene communities of Puerto Rico. Annals of the Missouri Botanical Gardens 56: 308357.CrossRefGoogle Scholar
Grandcolaas, P.M., Murienne, J., Robillard, T., et al. 2008. New Caledonia: a very old Darwinian island? Philosophical Transactions of the Royal Society B: Biological Sciences 363: 33093317.CrossRefGoogle Scholar
Grauvogel-Stamm, L. 1978. La Flore du Grès a Volzia (Bundsandstein Supérieur) des Vosges du Nord (France): Morphologie, anatomie, interprétations, phylogénetique, et paléogéographique. Université Louis Pasteur de Strasbourg, Institute Géologique Mémoires 50: 1225.Google Scholar
Gray, B. 1973. Distribution of Araucaria in Papua New Guinea. Research Bulletin Department of Forestry, Papua New Guinea.Google Scholar
Gray, B. 1975. Size composition and regeneration of Araucaria stands in New Guinea. Journal of Ecology 63: 273289.CrossRefGoogle Scholar
Gray, J.B., Wentworth, T.R. & Brownie, C. 2003. Extinction, colonization, and persistence of rare vascular flora in the longleaf pine–wiregrass ecosystem: responses to fire frequency and population size. Natural Areas Journal 23: 210219.Google Scholar
Gray, N.E. 1953. A taxonomic revision of Podocarpus VII. Journal of the Arnold Arboretum 34: 6770.CrossRefGoogle Scholar
Gray, N.E. 1953. A taxonomic revision of Podocarpus VIII: the African species of section Eupodocarpus, subsection A and E. Journal of the Arnold Arboretum 34: 163175.CrossRefGoogle Scholar
Gray, N.E. 1955. A taxonomic revision of Podocarpus IX: the South Pacific species of Section Eupodocarpus, subsection F. Journal of the Arnold Arboretum 36: 199206.CrossRefGoogle Scholar
Gray, N.E. 1956. A taxonomic revision of Podocarpus X. Journal of the Arnold Arboretum 39: 424477.CrossRefGoogle Scholar
Gray, N.E. 1956. A taxonomic revision of Podocarpus X: the South Pacific species of Section Eupodocarpus, subsection D. Journal of the Arnold Arboretum 37: 160172.CrossRefGoogle Scholar
Gray, N.E. 1958. A taxonomic revision of Podocarpus XI. Journal of the Arnold Arboretum 34: 6770.CrossRefGoogle Scholar
Gray, N.E. 1958. A taxonomic revision of Podocarpus XI: the South Pacific species of Section Podocarpus, subsection B. Journal of the Arnold Arboretum 39: 424477.CrossRefGoogle Scholar
Gray, N.E. 1960. A taxonomic revision of Podocarpus XII, section Microcarpus. Journal of the Arnold Arboretum 40(1): 3639.CrossRefGoogle Scholar
Gray, N.E. 1962. A taxonomic revision of Podocarpus XIII, section Polypodiopsis in the South Pacific. Journal of the Arnold Arboretum 43(1): 6779.CrossRefGoogle Scholar
Gray, N.E. & Buchholz, J.T. 1951. A taxonomic revision of Podocarpus V. Journal of the Arnold Arboretum 32: 8292.CrossRefGoogle Scholar
Gray, N.E. & Buchholz, J.T. 1951. A taxonomic revision of Podocarpus VI. Journal of the Arnold Arboretum 32: 9397.CrossRefGoogle Scholar
Greenberg, C.H. 2003. Vegetation recovery and stand structure following a prescribed stand-replacement burn in sand pine scrub. Natural Areas Journal 23: 141151.Google Scholar
Gregor, H. 1979. Fruktifikation der Gattung Cephalotaxus Siebold & Zuccarini aus den Tertiar Europas und Japans. Feddes Rep. 90: 110.CrossRefGoogle Scholar
Greguss, P. 1955. Xylotomische Bestimmung der Heute Lebenden Gymnospermen. Budapest: Akademia Kiado.Google Scholar
Greguss, P. 1967. Fossil Gymnosperm Woods in Hungary, from the Permian to the Pliocene. Budapest: Akademiai Kiado.Google Scholar
Greguss, P. 1972. Xylotomy of the Living Conifers. Budapest: Akadémiai Kladó.Google Scholar
Grierson, A.J.C., Long, D.G. & Page, C.N. 1980. Pinus bhutanica: a new 5-needle pine from Bhutan and India. Notes from the Royal Botanic Garden, Edinburgh 38: 297310.Google Scholar
Griffin, J.R. 1965. Digger pine seedling response to serpentine and non-serpentine soil. Ecology 46: 801807.CrossRefGoogle Scholar
Griffiths, M.E. & Lawes, M.J. 2006. Biogeographic, environmental, and phylogenetic influences on reproductive traits in subtropical forest trees, South Africa. Ecography 29: 614622.CrossRefGoogle Scholar
Griffiths, M.M. 1957. Foliar ontogeny in Podocarpus macrophyllus, with special reference to transfusion tissue. American Journal of Botany 44: 705715.CrossRefGoogle Scholar
Grime, J.P. & Hunt, R. 1975. Relative growth rate: its range and adaptive significance in a local flora. Journal of Ecology 63: 393422.CrossRefGoogle Scholar
Groves, R.H. (ed.). Australian Vegetation. Cambridge: Cambridge University Press.Google Scholar
Grubb, P.J. 1977. The maintenance of species richness in plant communities: the importance of the regeneration niche. Biological Review 52: 107145.CrossRefGoogle Scholar
Grubb, P.J. 1977. Control of forest growth and distribution on wet tropical mountains, with special reference to mineral nutrition. Annual Review of Ecology and Systematics 8: 83107.CrossRefGoogle Scholar
Grubb, P.J. 1992. A positive distrust in simplicity: lessons from plant defences and from competition among plants and among animals. Journal of Ecology 80: 585610.CrossRefGoogle Scholar
Grubb, P.J. & Stevens, P.J. 1979. The Forests of the Fatima Basin and Mt. Kerigomna and a Review of Montane and Subalpine Forests Elsewhere in P.N.G. Canberra: Australian National University.Google Scholar
Grubb, P.J., Lloyd, J.R., Pennington, T.D. & Whitmore, T.C. 1963. A comparison of montane and low-land rain forest in Equador. I. The forest structure, physiognomy and floristics. Journal of Ecology 51: 567601.CrossRefGoogle Scholar
Guariguata, M.R. 1990. Landslide disturbance and forest regeneration in upper Luquillo Mountains of Puerto Rico. Journal of Ecology 78: 814832.CrossRefGoogle Scholar
Guedes, M. 1970. La morphologie du complexe séminifère de Cryptomeria japonica (L. fil.) Don. Flora 159, 7183.CrossRefGoogle Scholar
Guillaumin, A. 1952. Les caractères de la végétation Néo-calédonienne. Compte rendu des séances de la Société de biogéographie. 29: 8286.Google Scholar
Guilon, J.H. 1975. Les massifs peridotitiques de Nouvelle-Calédonie. Memoir ORSTOM 76: 1120.Google Scholar
Gulbranson, E.L., Isbell, J.L., Ryberg, P.E., Taylor, T.N. & Flaig, P.P. 2012. Permian polar forests: deciduousness and environmental variation. Geobiology 10: 479495.CrossRefGoogle ScholarPubMed
Gulline, H.F. 1952. The cytology of Athrotaxis. Papers and Proceedings of the Royal Society of Tasmania 86: 131136.CrossRefGoogle Scholar
Gunckel, J.E. & Wetmore, R.H. 1946. Studies of development in long shoots and short shoots of Ginkgo biloba l. American Journal of Botany, 33: 285295, 532–543.CrossRefGoogle Scholar
Guo, Z.T., Ruddiman, W.F. & Hao, Q.Z. 2002. Onset of Asian desertification by 22 Myr ago inferred from loess deposits in China. Nature 416: 159163.CrossRefGoogle Scholar
Guries, R.P. & Ledig, F.T. 1982. Genetic diversity and population structure in pitchpine (Pinus rigida Mill.). Evolution 36: 387402.Google Scholar
Hahm, W.J., Riebe, C.S., Lukens, C.E. & Araki, S. 2014. Bedrock composition regulates mountain ecosystems and landscape evolution. Proceedings of the National Academy of Sciences USA 111: 33363343.CrossRefGoogle ScholarPubMed
Hahs, A., Enright, N.J. & Thomas, I. 2002. Plant communities, species richness and their environmental correlates in the sandy heaths of Little Desert national park, Victoria. Australian Journal of Ecology 24: 249257.CrossRefGoogle Scholar
Hair, J.B. & Beuzenberg, E.J. 1958. Contributions to a chromosome atlas of the New Zealand flora – 1. New Zealand Journal of Science 1: 617628.Google Scholar
Hall, R. & Blundell, D. (eds.). Tecotonic Evolution of Southeast Asia. London: Geological Society.Google Scholar
Hallam, A. & Wignall, P.B. 1997. Mass Extinctions and their Aftermath. Oxford: Oxford University Press.CrossRefGoogle Scholar
Halle, T.G. 1915. Some Xerophytic Leaf-Structures in Mesozoic Plants. Stockholm: Geol. Forens Forhandl.CrossRefGoogle Scholar
Halling, R.E. 2001. Ectomycorrhizae: co-evolution, significance and biogeography. Annals of the Missouri Botanical Garden 88: 513.CrossRefGoogle Scholar
Halpern, C. & Spies, T. 1995. Plant species diversity in natural and managed forests of the Pacific Northwest. Ecological Applications 5: 913934.CrossRefGoogle Scholar
Hammel, H.T. 1967. Freezing of xylem sap without cavitation. Plant Physiology 42: 5566.CrossRefGoogle ScholarPubMed
Hamrick, J.L. & Godt, M.J.W. 1996. Effect of life-history on genetic diversity of plant species. Philosophical Transactions of the Royal Society of London B 351: 12911298.Google Scholar
Hamrick, J.L., Sachnabel, A.F. & Wells, P.V. 1994. Distribution of genetic diversity within and among populations of Great Basin conifers. Pp 147161 in Harper, K.T. (ed.), Natural History of the Colorado Plateau and Great Basin. Denver, CO: University Press of Colorado.Google Scholar
Handro, W. & Ferréira, A.G. 1979. Aspects of seed germination in Araucaria angustifolia. Revista Brasiliera de Botanica 2: 713.Google Scholar
Harcombe, P.A., Harmon, M.E. & Greene, S.E. 1990. Changes in biomass and production over 53 years in a coastal Picea sitchensisTsuga heterophylla forest approaching maturity. Canadian Journal of Forest Research 20: 16021610.CrossRefGoogle Scholar
Harden, G.J. & Thompson, J. (eds.). 1990. Cupressaceae: Flora of New South Wales, Vol 1. Sydney: University of New South Wales Press.Google Scholar
Hare, R.C. & Switzer, G.L. 1969. Introgression with Shortleaf Pine May Explain Rust Resistance in Western Loblolly Pine. Washington, DC: USDA.Google Scholar
Harley, J.L. & Harley, E.L. 1987. A check-list of mycorrhiza in the British flora. New Phytologist 105(Suppl. S1): 1102.CrossRefGoogle Scholar
Harlow, W.M. (1931). The Identification of the Pines of the United States, Native and Introduced, by Needle Structure. Syracuse, NY: Syracuse University.Google Scholar
Harmon, M.E., Braton, S.P. & White, P.S. 1983. Disturbance and vegetation response in relation to environmental gradients in the Great Smoky Mountains. Vegetatio 55: 129139.CrossRefGoogle Scholar
Harper, C.J., Taylor, T.N., Krigs, M. & Taylor, E.L. 2013. Mycorrhizal symbiosis in the Paleozoic seed fern Glossopteris from Antarctica. Review of Palaeobotany and Palynology 192: 2231.CrossRefGoogle Scholar
Harris, T.M. 1976. Two neglected aspects of fossil conifers. American Journal of Botany 63: 902910.CrossRefGoogle Scholar
Harris, T.M. 1979. The Yorkshire Jurassic Flora, V. Coniferales. London: British Museum (Natural History).Google Scholar
Harris, T.M., Millington, W. & Miller, J. 1974. The Yorkshire Jurassic Flora: Vol 4 parts 1 and 2. Ginkgoales and Czekanowskiales. London: British Museum (Natural History).Google Scholar
Harrison, S. 1999. Local regional diversity in a patchy landscape: native, alien and endemic herbs on serpentine. Ecology 80: 7080.CrossRefGoogle Scholar
Hart, J.A. & Price, R.A. 1990. The genera of Cupressaceae (including Taxodiaceae) in the southeastern United States. Journal of the Arnold Arboretum 71: 275322.CrossRefGoogle Scholar
Hartesveld, R.J., Harvey, H.T., Shellhammer, H.S. & Stecker, R.E. The Giant Sequoia of the Sierra Nevada. Washington, DC: US Department of the Interior, National Park Service.Google Scholar
Hartzell, H.R. Jr. 1991. The Yew Tree: A Thousand Whispers. Eugene, OR: Hulogosi.Google Scholar
Harvey, H.F. 1985. Population biology and the conservation of rare species. Pp 111123 in White, J. (ed.), Studies in Plant Demography. London: Academic Press.Google Scholar
Hasebe, M., Ito, M., Kofuji, R., Iwatsuki, K. & Ueda, K. 1992. Phylogenetic relationships in Gnetophyta deduced from rbcL gene sequences. Botanical Magazine Tokyo 105: 385391.CrossRefGoogle Scholar
Hasebe, M., Kofuji, R., Ito, M., Iwatsuki, K. & Ueda, K. 1992. Phylogeny of gymnosperms inferred from rbcL gene sequences. Botanical Magazine Tokyo 105: 673679.CrossRefGoogle Scholar
Hastings, P.A. 1990. Southern oscillation influences on tropical cyclone. International Journal of Climatology 10: 291298.CrossRefGoogle Scholar
Hathway, B. 2000. Continental rift to back-arc basin: Jurassic–Cretaceous stratigraphical and structural evolution of the Larsen Basin, Antarctic peninsula. Journal of the Geological Society of London 157: 417432.CrossRefGoogle Scholar
Hawkesworth, F.G. & Weins, D. 1972. Biology and Classification of Dwarf Mistletoes (Arceuthobium). Washington, DC: USDA.Google Scholar
Haworth, M., Heath, J. & McElwain, J.C. 2010. Differences in the response sensitivity of stomatal index to atmospheric CO2 amongst four genera of Cupressaceae conifers. Annals of Botany 105: 411418.CrossRefGoogle Scholar
Hay, W.W., DeConto, R., Wold, C.N., et al. 1999. Alternative global Cretaceous paleogeography. Geological Society of America Special Papers 332: 147.Google Scholar
Hayashida, M. 1989. Seed dispersal by red squirrels and subsequent establishment of Korean pine. Forest Ecology and Management 28: 115129.CrossRefGoogle Scholar
Hayata, B. 1915. Icones Plantaraum Formosarum nec non et Contributiones ad Floram Formosanum, Vol. 5. Taiwan: Formosa Government.Google Scholar
Hayata, B. 1931. The Sciadopityaceae represented by Sciadopitys verticillata Sieb. et Zucc., an endemic species of Japan. Botanical Magazine Tokyo 45: 567569.Google Scholar
Hayata, B. 1932. The Taxodiaceae should be divided into several distinct families; and further Tetraclinis should represent a distinct family, the Tetraclinaceae. Botanical Magazine Tokyo 46: 2427.Google Scholar
Heads, M. 1990. Mesozoic tectonics and the deconstruction of biogeography: a new model of Australasian biology. Journal of Biogeography 17: 223225.Google Scholar
Heads, M. 2008. Panbiogeography of New Caledonia, south-west Pacific: basal angiosperms on basement terranes, ultramafic endemics inherited from volcanic island arcs and old taxa endemic to young islands. Journal of Biogeography 35: 21532175.CrossRefGoogle Scholar
Heads, M. 2010. Biogeographical affinities of the New Caledonian biota: a puzzle with 24 pieces. Journal of Biogeography 37: 11791201.CrossRefGoogle Scholar
Heaney, L.R. 2000. Dynamic disequilibrium: a long-term, large-scale perspective on the equilibrium model of island biogeography. Global Ecology and Biogeography 9: 5974.CrossRefGoogle Scholar
Heckman, D.S., Geiser, D.M., Eidell, B.R., et al. 2001. Molecular evidence for the early colonisation of land by fungi and plants. Science 293: 11291133.CrossRefGoogle ScholarPubMed
Heer, O. 1876. Flora Fossilis Arctica. IV. Part 1. Beitrage zur fossilen Flora Spitzbergens. K. Svenska Vetenskaps Akad. Handl. 14: 1144.Google Scholar
Hendricks, D.R. & Søndergaard, P. 1998. Metaseqoia glyptostroboides 50 years out of China: observations from the United States and Denmark. Dansk Dendrologisk Årsskrift 16: 624.Google Scholar
Henry, A. & McIntyre, M. 1926. The swamp cypresses, Glyptostrobus of China and Taxodium of America, with notes on allied genera. Proceedings of the Royal Irish Academy 37: 90116.Google Scholar
Henslow, G. 1889. Yew trees in Berks. Nature 40: 621.CrossRefGoogle Scholar
Herbert, D.A. 1950. Present day distribution and the geological past. Victorian Naturalist 66: 227237.Google Scholar
Herbert, D.A. 1966. Ecological segregation and Australian phytogeographic elements. Proceedings of the Royal Society of Queensland 78: 101111.Google Scholar
Herbert, D.A., Fownes, J.H. & Vitousek, P.M. 1999. Hurricane damage to a Hawaiian forest: nutrient supply rate affects resistance and resilience. Ecology 80: 908920.CrossRefGoogle Scholar
Heusser, C.J. 1981. Palynology of the last interglacial–glacial cycle in mid-latitudes of southern Chile. Quaternary Research 16: 293321.CrossRefGoogle Scholar
Heusser, C.J. & Flint, R.F. 1977. Quaternary glaciations and environments of northern Isla de Chiloé, Chile. Geology 5: 305308.2.0.CO;2>CrossRefGoogle Scholar
Heusser, C.J., Rabassa, J., Brandani, A. & Stuckenrath, R. 1988. Late Holocene vegetation of the Andes Araucaria region, Province of Neuquen, Argentina. Mountain Research and Development 8: 5363.CrossRefGoogle Scholar
Heywood, V.H., Brummitt, R.K., Culham, A. & Sedberg, O. 2007. Flowering Plant Families of the World. London: Royal Botanic Gardens, Kew.Google Scholar
Hickel, R. 1926. A propos de la germination des Araucaria. Bulletin de la Société Botanique de France 73: 968970.CrossRefGoogle Scholar
Hida, M. 1956. The comparative study of Taxodiaceae from the standpoint of the development of the cone scale. Botanical Magazine (Tokyo) 70: 4451.CrossRefGoogle Scholar
Hida, M. 1962. The systematic position of Metasequoia. Botanical Magazine (Tokyo) 75: 316323.CrossRefGoogle Scholar
Hiebert, R.D. & Hamrick, J.L. 1983. Patterns and levels of genetic variation in Great Basin bristlecone pine, Pinus longaeva. Evolution 37: 302310.CrossRefGoogle Scholar
Hiep, N.T., Loc, P.K., Luu, N.D.T., et al. 2004. Vietnam Conifers Conservation Status Review. Hanoi: Labour and Society Publisher.Google Scholar
Hill, C.R. & Crane, P.R. 1982. Evolutionary cladistic and the origin of angiosperms. Pp 269361 in Joysey, K.A. & Friday, A.E. (eds.), Problems of Phylogenetic Reconstruction. New York: Academic Press.Google Scholar
Hill, K.D. 2003. The Wollemi Pine, another living fossil? Pp 157164 in Mill, R.R. (ed.), Conifers for the Future? Proceedings of the Fourth International Conifer Conference. Wye: Acta Horticulturae.Google Scholar
Hill, R.S. 1982. The Eocene megafossil flora of Nerriga, New South Wales, Australia. Paleontographica Abt. B 181: 4477.Google Scholar
Hill, R.S. 2001. Biogeography, evolution and palaeoecology of Nothofagus (Nothofagaceae): the contribution of the fossil record. Australian Journal of Botany 49: 321332.CrossRefGoogle Scholar
Hill, R.S. & Brodribb, T.J. 2001. Macrofossil evidence for the onset of xeromorphy in Australian Casuarinaceae and tribe Banksieae (Proteaceae). Journal of Mediterranean Ecology 2: 127136.Google Scholar
Hill, R.S. & Carpenter, R.J. 1999. Ginkgo leaves from Paleogene sediments in Tasmania. Australian Journal of Botany 47: 717724.Google Scholar
Hill, R.S. & Christophel, D.C. 2001. Two new species of Dacrydium (Podocarpaceae) based on vegetative fossils from Middle Eocene sediments at Nelly Creek, South Australia. Australian Systematic Botany 14: 193205.CrossRefGoogle Scholar
Hill, R.S. & Read, J. 1987. Endemism in Tasmanian cool temperate rainforest: alternative hypotheses. Botanical Journal of the Linnean Society 95: 113124.CrossRefGoogle Scholar
Hill, R.S. & Scriven, L.J. 1997. Palaeoclimate across an altitudinal gradient in the Oligo-Miocene of northern Tasmania: an investigation of nearest living relative analysis. Australian Journal of Botany 45: 493505.Google Scholar
Hill, R.J., Jordan, G.J. & Carpenter, R.J. 1993. Taxodiaceous macrofossils from tertiary sediments in Tasmania. Australian Systematic Botany 6: 237249.CrossRefGoogle Scholar
Hill, T.G. & de Fraine, E. 1908. On the seedling structure of gymnosperms. I. Annals of Botany 22: 689712.CrossRefGoogle Scholar
Hill, T.G. & de Fraine, E. 1909. On the seedling structure of gymnosperms. part II. Abietineae and Araucarieae. Annals of Botany 23: 189227.CrossRefGoogle Scholar
Hills, L.V. & Ogilvie, R.T. 1970. Picea banksii n. sp. Beaufort formation (Tertiary), northwestern Banks Island, Arctic Canada. Canadian Journal of Botany 48: 457464.CrossRefGoogle Scholar
Hinojosa, L.F., Armesto, J.J. & Villagran, C. 2006. Are Chilean coastal forests pre-Pleistocene relicts? Evidence from foliar physiognomy, palaeoclimate and phytogeography. Journal of Biogeography 33: 331341.CrossRefGoogle Scholar
Hirayama, K. & Sakimoto, M. 2003. Regeneration of Cryptomeria japonica on a sloping topography in a cool-temperate mixed forest in the snowy region of Japan. Canadian Journal of Forest Research 33: 543551.CrossRefGoogle Scholar
Hirmer, M. 1936. Die Blüten der Coniferen: entwicklungsgeschichte und vergleichende Morphlogie des weiblichen Blütenzapfens der Coniferen. Bibliotheca Botanica 114: 1100.Google Scholar
Hizume, M., Abe, K.K. & Tanaka, A. 1988. Fluorescent chromosome banding in the Taxodiaceae. Kromosomo II 50: 16091619.Google Scholar
Hizume, M., Shiraishi, H. & Tanaka, A. 1988. A cytological study of Podocarpus macrophyllus with special reference to sex chromosomes. Japan Journal of Genetics 63: 413423.Google Scholar
Hizume, M., Yamasaki, Y., Kondo, K., et al. 1994. Fluorescent chromosome bandings in two Chinese varieties of Larix gmelinii, Pinaceae. Kromosomo II 74: 25632570.Google Scholar
Ho, C.C. 1950. Leaf anatomy of Chinese species of Podocarpus. Botanical Bulletin of Academia Sinica 3(4): 146150.Google Scholar
Hocodent, E. 1964. Recherches cytologiques sur le developpement de la misrospore en grain de pollen chez Podocarpus nagi (R.Br.). Comtes Rendus, l’Académie des Sciences (Paris) 259: 11791182.Google Scholar
Hocodent, E. 1968. Chromosome count and morphology in the female prothallus of Araucaria araucana. Caryologia 21: 385388.Google Scholar
Hodge, A., Campbell, C.D. & Fitter, A.H. 2001. An arbuscular mycorrhizal fungi accelerates decomposition and acquires nitrogen directly from organic material. Nature 413: 297299.CrossRefGoogle ScholarPubMed
Hodge, G.R. & Dvorak, W.S. 2007. Variation in pitch canker resistance among provenances of Pinus patula and Pinus tecunumanii from Mexico and Central America. New Forests 33: 193206.CrossRefGoogle Scholar
Hoeppner, S.S. & Rose, K.A. Individual-based modelling of flooding and salinity effects on a coastal swamp forest. Ecological Modelling 222: 35413558.CrossRefGoogle Scholar
Hoffmann, A.E. 1982. Flora Silvestre de Chile. Zona Araucana, 2nd edn. Santiago de Chile: Fundación Claudio Gay.Google Scholar
Hofgaard, A. 1993. Fifty years of change in a Swedish boreal old-growth Picea abies forest. Journal of Vegetation Science 4: 773784.CrossRefGoogle Scholar
Holder, C.D. 2006. The hydrological significance of cloud forests in the Sierra de las Minas Biosphere Reserve, Guatemala. Geoforum 37: 8293.CrossRefGoogle Scholar
Holdridge, L.R., Grenke, W.C., Hatheway, W.H., Liang, T. & Tosi, J.A. 1971. Forest Environments in Tropical Life Zones, a Pilot Study. Oxford: Pergamon Press.Google Scholar
Hollander, J.L. & Vander Wall, S.B. 2004. Effectiveness of six species of rodents as dispersers of singleleaf pinion pine (Pinus monophylla). Oecologia 138: 5765.CrossRefGoogle Scholar
Holloway, J.T. 1937. Ovule anatomy and development and embryogeny in Phyllocladus alpinus (Hook) and in P. glaucus (Carr.). Transactions and Proceedings of the Royal Society of New Zealand 67: 149165.Google Scholar
Holloway, J.T. 1954. Forests and climates of South Island, New Zealand. Transactions of the Royal Society of New Zealand 82: 329410.Google Scholar
Holmgren, M., Scheffer, M. & Huston, M.A. 1997. The interplay of facultation and competition in plant communities. Ecology 78: 19661975.CrossRefGoogle Scholar
Hooghiemstra, H., Wijninga, V.M. & Cleef, A.M. 2006. The palaeobotanical record of Colombia: implications for diogeography and biodiversity. Annals of the Missouri Botanical Garden 93: 297325.CrossRefGoogle Scholar
Hooker, W.J. 1844. Description, with a figure, of a new species of Thuja, the Alerse of Chile. Journal of Botany (London) 3: 144149.Google Scholar
Hoorne, C., Guero, J. Sarmiento, G.A. & Lorente, M.A. 1995. Andean tectonics as a cause for changing drainage patterns in Miocene northern South America. Geology 23: 237240.2.3.CO;2>CrossRefGoogle Scholar
Hope, G.S. 1998. Early fire and forest change in the Baliem Valley, Irian Jaya, Indonesia. Journal of Biogeography 25: 453461.CrossRefGoogle Scholar
Hori, T., Ridge, R.W., Tulecke, W., et al. (eds.), Ginkgo biloba: A Global Treasure from Biology to Medicine. Tokyo: Springer.Google Scholar
Horrocks, M., Ogden, J., Nichol, S.L., Alloway, B.V. & Sutton, D.G. 2000. A Late Quaternary palynological and sedimentological record from two coastal swamps at southern Kaitoke, Great Barrier Island, New Zealand. Journal of the Royal Society of New Zealand 30: 4969.CrossRefGoogle Scholar
Horton, T.R. & Bruns, T.D. 1998. Multiple-host fungi are the most frequent and abundant ectomycorrhizal types in a mixed stand of Douglas fir (Pseudotsuga menziesii) and Bishop pine (Pinus muricata). New Phytologist 139: 331339.CrossRefGoogle Scholar
Howe, H.F. 1989. Ecology of seed dispersal by birds. Annual Review of Ecology and Systematics 13: 201228.CrossRefGoogle Scholar
Howe, H.F. 1989. Scatter- and clump-dispersal and seedling demography: hypothesis and implications. Oecologia 79: 417427.CrossRefGoogle ScholarPubMed
Hsieh, L. 1976. The living fossil: the past and present Metasequoia glyptostroboides Hu et Cheng. Acta Phytotaxonomica Sinica 14: 5154.Google Scholar
Hsu, J. 1974. Late Cretaceous and Cenozoic vegetation in China, emphasising their connections with North America. Annals of the Missouri Botanical Garden 70: 490508.CrossRefGoogle Scholar
Hu, H.H. 1950. Metasequoia and its history. Chinese Journal of Botany 5: 913 (in Chinese).Google Scholar
Hu, Y.S. & Yao, B.L. 1981. Transfusion tissue in gymnosperm leaves. Botanical Journal of the Linnean Society 83: 263272.CrossRefGoogle Scholar
Hu, Y.S., Guan, L.-Q. & Tang, Z.-X. 1985. Anatomy of the secondary phloem and the crystalliferous phloem fibres in the stem of Torreya grandis. Acta Botanica Sinica 27: 569574 (in Chinese with English summary).Google Scholar
Hu, Y.S., Lin, J.X., Wang, X.P. & Wei, L.B. 1995. The biology and conservation of Taiwania cryptomerioides. Chinese Biodiversity 3: 206212 (in Chinese with English abstract, abstract seen only).Google Scholar
Hu, Z.-A. 1988. Biochemical systematics of Gymnosperms (1): peroxidases of Pinaceae. Acta Phytotaxonomica Sinica 21: 423431.Google Scholar
Hu, Z.-A. 1989. Biochemical systematics of Gymnosperms (2): seed protein peptides of Pinaceae. Acta Phytotaxonomica Sinica 22: 360366.Google Scholar
Huai, W.-X., Guo, L.-D. & Wei, H. 2003. Genetic diversity of an ectomycorrhizal fungus Tricholoma terreum in a Larix principisrupprechtii stand assessed using random amplified polymorphic DNA. Mycorrhiza 13: 265270.CrossRefGoogle Scholar
Hubbell, S.P. 2001. The Unified Neutral Theory of Biodiversity and Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
Hubert, N. & Gehring, C. 2015. Neighboring trees affect ectomycorrhizal fungal community composition in a woodland-forest ecotone. Mycorrhiza 25: 112.Google Scholar
Huffman, J.M., Platt, W.J., Grissino-Mayer, H. & Boyce, C.J. 2004. Fire history of a barrier island slash pine (Pinus elliottii) savanna. Natural Areas Journal 24: 258268.Google Scholar
Hughes, C.E., Pennington, R.T. & Antonelli, A. 2013. Neotropical plant evolution: assembling the big picture. Botanical Journal of the Linnean Society 171: 118.CrossRefGoogle Scholar
Hughes, M.F. 1976. Palaeobiology of Angiosperm Origins. Cambridge: Cambridge University Press.Google Scholar
Hulten, E. & Fries, M. 1986. Atlas of North European Vascular Plants, North of the Tropic of Cancer. Konigenstein: Koeltz Scientific Books.Google Scholar
Hummel, J., Suedekum, K.-H. & Clauss, M. 2005. Experimental estimation of the nutritional value of potential food plants of herbivorous dinosaurs, with special emphasis on sauropods. Journal of Vertebrate Paleontology 25: 72A.Google Scholar
Hummel, J., Gee, C.T., Südenkum, K.-H., et al. 2008. In vitro digestibility of fern and gymnosperm foliage: implications for sauropod feeding ecology and diet selection. Proceedings of the Royal Society B 275: 10151021.CrossRefGoogle ScholarPubMed
Huneycutt, M. & Askew, G.R. 1989. Electrophoretic identification of loblolly pine–shortleaf pine hybrids. Silvae Genetica 38: 9596.Google Scholar
Hunter, J.C. & Barbour, M.G. 2001. Through-growth by Pseudotsuga menziesii: a mechanism for change in forest composition without canopy gaps. Journal of Vegetation Science 12: 445452.CrossRefGoogle Scholar
Hunter, M.L. 1999. Maintaining Biodiversity in Forest Ecosystems. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Huntley, B. 1993. Species-richness in north-temperate zone forests. Journal of Biogeography 20: 163180.CrossRefGoogle Scholar
Huson, D.H. & Bryant, D. 2006. Application of genetic networks in evolutionary studies. Molecular Biology and Evolution 23: 256267.CrossRefGoogle ScholarPubMed
Hutchinson, A.H. 1917. Morphology of Keteleeria fortunei. Botanical Gazette 63: 124134.CrossRefGoogle Scholar
Hwang, S.Y., Lin, T.P., Ma, C.S., et al. 2003. Post-glacial population growth of Cunninghamia konishii (Cupressaceae) inferred from phylogeographical and mismatch analysis of chloroplast DNA variation. Molecular Ecology 12: 26892695.CrossRefGoogle Scholar
Hyakawa, T., Tomaru, N. & Yamamoto, S.I. 2004. Stem distribution and clonal structure of Chamaecyparis pisifera growing in an old-growth beech–conifer forest. Ecological Research 19: 411420.CrossRefGoogle Scholar
Hyland, B.P.M. 1975. A revision of the genus Agathis (Araucariaceae) in Australia. Brunonia 1: 103115.CrossRefGoogle Scholar
Hyndman, D.C. & Menzies, J.I. 1990. Rainforests of the Ok Tedi headwaters, New Guinea: an ecological analysis. Journal of Biogeography 17: 241273.CrossRefGoogle Scholar
Ishizuka, K. 1961. A relict stand of Picea glehnii Masters on Mt Hayacine, Iwate Prefecture. Ecological Review 15: 155162.Google Scholar
Ivonen, S., Rikala, R., Ryyppö, A. & Vapaavuori, E. 1999. Responses of Scots pine (Pinus sylvestris) seedlings grown in different nutrient regimes to changing root-zone temperature in spring. Tree Physiology 19: 951958.CrossRefGoogle Scholar
Iwata, H., Ujino-Ihara, T., Yoshimura, K., et al. 2001. Cleaved amplified polymorphic sequence markers in sugi, Cryptomeria japonica D. Don., and their locations on a linkage map. Theoretical and Applied Genetics 103: 881895.CrossRefGoogle Scholar
Iwatsuki, Z. & Sharp, A.J. 1967. The bryogeographical relationships between Eastern Asia and North America, I. Journal of the Hattori Botanical Laboratory 30: 152170.Google Scholar
Jaehnichen, H. 1999. Zapfen und Blaetter von Taiwania Hayata (Taxodiaceae) im Tertiaer von Eurasia. Courier Forschungsinstitut Senkenberg 109: 241.Google Scholar
Jaffré, T. & Pelletier, B. 1992. Plantes de Nouvelle-Calédonie permettant de revégétaliser des sites miniers. Noumea: SLN.Google Scholar
Jaffré, T. & Veillon, J.-M. 1990. Études floristique et structurale de deux forêts denses humides sur les roches ultrabasiques en Nouvelle-Calédonie. Bulletin du Muséum National d’Histoire Naturelle, Paris 12: 243272.Google Scholar
Jaffré, T., Lee, J., Brooks, R.R., Reeves, R.D. & Boswell, C.R. 1977. Plant–soil relationships in a New Caledonian serpentine flora. Plant Soil 46: 681685.Google Scholar
Jaffré, T., Veillon, J.-M. & Cherrier, J.F. 1987. Sur la presence de deux Cupressaceae Neocallitropsis pancheri (Carr.) Laubenf. et Libocedrus austrocaledonica Brong. et Gris. dans le massif du Paéoua et localites nouvelle de Gymnospermes en Nouvelle-Calédonie. Adansonia 3: 273288.Google Scholar
Jaffré, T., Morat, P.H., Veillon, J.-M. & Mackee, H.S. 1987. Changements dans la vegetation de la Nouvelle-Calédonie au cours du tertaire: la vegetation et la flore des roches ultrabaiques. Adansonia 4: 365391.Google Scholar
Jaffré, T., Morat, P. & Veillon, J.M. 1994. The flora of New Caledonia: characteristics and floristic composition of the most important plant formations. Bois et Forêts de Tropiques 242: 730.Google Scholar
Jaffré, T., Reeves, R.D. & Bacquer, T. 1997. The Ecology of Ultramafic and Metalliferous Areas. Noumea, New Caledonia: Centre ORSTOM.Google Scholar
Jaffré, T., Bouchet, P. & Veillon, J.-M. 1998. Threatened plants of New Caledonia: is the system of protected areas adequate? Biodiversity and Conservation 7: 109135.CrossRefGoogle Scholar
Jagel, A. & Stützel, T. 2001. Untersuchen zur Morphologies und Morphogenese der samenzapfen von Platycladus orientalis (L.) France (= Thuja orientalis L.) und Microbiota decussata Kom. (Cupressaceae). Bot. Jahrb. Syst. 123: 377404.Google Scholar
Jagel, A. & Stützel, T. 2001. Zur Abgrenzung von Chamaecyparis Spach und Cupressus L. (Cupressaceae) und ide systematishe Stelung von Cupressus nootkatensis D.Don [=Chamaecyparis nootkatensis (D.Don) Spach]. Feddes Repertorium 112: 179229.CrossRefGoogle Scholar
Jagels, R., Visscher, G.E., Lucas, J. & Goodell, B. 2003. Palaeo-adaptive properties of the xylem of Metasequoia: mechanical/hydraulic compromises. Annals of Botany 92: 7988.CrossRefGoogle Scholar
Jain, K.K. 1976. Evolution of wood structure in Pinaceae. Israel Journal of Botany 25: 2833.Google Scholar
Jain, S.L., Kutty, T.S. & Roychowdhury, T. Sauropod dinosaur from the Lower Jurassic Kota formation of India. Proceedings of the Royal Society of London B 188: 221228.Google Scholar
Janssen, T. & Bremner, K. 2004. The age of major monocot groups inferred from 800+ rbcL sequences. Botanical Journal of the Linnean Society 146: 385398.CrossRefGoogle Scholar
Jansson, R. 2003. Global patterns in endemism explained by past climatic change. Proceedings of the Royal Society B 270: 583590.CrossRefGoogle ScholarPubMed
Jeffrey, E.C. 1905. Comparative anatomy and phylogeny of the Coniferales. Part II: The Abietineae. Memoirs of the Boston Society of Natural History. 4: 137.Google Scholar
Jeffrey, E.C. 1907. Araucariopitys, a new genus of araucarians. Botanical Gazette 44: 435444.CrossRefGoogle Scholar
Jeffrey, E.C. 1912. The history, comparative anatomy, and evolution of the araucarioxylon type. Proceedings of the American Academy of Arts and Sciences 48: 531571.CrossRefGoogle Scholar
Jeffrey, E.C. 1925. Resin canals in the evolution of the conifers. Proceedings of the National Academy of Sciences, USA 11: 101110.CrossRefGoogle Scholar
Jenkyns, H.C. 2003. Evidence for rapid climate change in the Mesozoic–Paleogene greenhouse world. Philosphical Transations of the Royal Society of London A 361: 18851916.CrossRefGoogle ScholarPubMed
Jensen, H. & Levan, A. 1941. Colchicine induced tetraploidy in Sequoia gigantea. Hereditas 27: 220224.CrossRefGoogle Scholar
Ji, L., Dong, B., Wei, C. & Wang, M. 2004. In-situ species diversity in Korean pine broad-leaved mixed forest in Changbai Mountains. Chinese Journal of Applied Ecology 15: 15271530.Google Scholar
Jia, Z, Ma, L., Xu, C., Zhang, H. & Sun, S. 2005. Soil moisture content and physical and chemical properties of most stands of young Platycladus orientalis in Beijing mountain area. Journal of Soil Water Conservation 19: 160164 (in Chinese).Google Scholar
Jiang, Z.P. & Wang, H.R. 1997. Taxonomy of the Cupressaceae: subfamilies, tribes and genera. Acta Phytotaxonomica Sinica 35: 236248. (in Chinese with English summary).Google Scholar
Jiménez, M.N., Navarro, F.B., Ripoll, M.A., Boscio, I. & de Simon, E. 2005. Effects of shelter tubes on establishment and growth of Juniperus thurifera L. (Cupressaceae) seedlings in Mediteranean semi-arid environments. Annals of Forest Science 62: 717725.CrossRefGoogle Scholar
Jiménez, J., Jurado, E., Aguirre, O. & Estrada, E. 2005. Effects of grazing on restoration of endemic dwarf pine (Pinus culminicola Andresen et Beaman) populations in northeastern Mexico. Restoration Ecology 13: 103107.CrossRefGoogle Scholar
John, S.M., Daniel, M. & Sabnis, S.D. 1989. Chemosystematics of some conifers of India. Proceedings of the Indian Academy of Science (Plant Sciences) 99: 253258.CrossRefGoogle Scholar
Johns, R.J. 1976. A classification of the montane forests of Papua New Guinea. Science New Guinea 4: 105127.Google Scholar
Johns, R.J. 1985. Altitudinal zonation of pteridophytes in Papuasia. Proceedings of the Royal Society of Edinburgh 86: 381389.Google Scholar
Johns, R.J. 1986. The instability of the tropical ecosystem in New Guinea. Blumea 31: 341371.Google Scholar
Johnson, L.A.S. & Briggs, B.G. 1975. On the Proteaceae: the evolution and classification of a southern family. Botanical Journal of the Linnean Society 70: 83182.CrossRefGoogle Scholar
Johnson, L.P.V. 1939. A descriptive list of natural and artificial interspecific hybrids in North American forest-tree genera. Canadian Journal of Forest Research 17: 411443.Google Scholar
Johnson, M., Vander Wall, S.B. & Borchert, M. 2003. A comparative analysis of cone characteristics and seed dispersal strategies of three pines in the subsection Sabinianae. Plant Ecology 168: 6984.CrossRefGoogle Scholar
Johnson, P.N. 1977. Mycorrhizal endogonaceae in a New Zealand forest. New Phytologist 78: 161170.CrossRefGoogle Scholar
Johnstone, M.H., Lowry, D.C. & Quilty, P.G. 1973. The geology of south-western Australia: a review. Journal of the Royal Society of Western Australia 56: 517.Google Scholar
Jones, G. & Peterson, E. 1970. Plant species diversity in a woodland–meadow ecotone near Regina, Saskatchewan. Canadian Journal of Botany 48: 591601.CrossRefGoogle Scholar
Jones, M.M., Tuomisto, P., Clark, D.B. & Olivas, P. 2006. Effects of mesoscale environmental heterogeneity and dispersal limitation on floristic variation in rainforest ferns. Journal of Ecology 94: 181185.CrossRefGoogle Scholar
Jordan, G.J. 1997. Evidence of Pleistocene plant extinction and diversity from Regatta Point, western Tasmania, Australia. Botanical Journal of the Linnean Society 123: 4571.CrossRefGoogle Scholar
Jordan, G.J., Carpenter, R.J. & Hill, R.S. 1998. The macrofossil record of Proteaceae in Tasmania: a review with new species. Australian Systematic Botany 11: 465501.CrossRefGoogle Scholar
Jordon-Barbolla, L., Delgado-Valero, P., Geada-López, G., Vázquez-Lobo, A. & Piñero, D. 2011. Phylogeography of Pinus subsection Australes in the Caribbean basin. Annals of Botany 107: 229241.CrossRefGoogle Scholar
Jorge, G.A. & Javier, G.C. 2005. Contendio de carbono en la biomass aérea de bosques natives en Chile. Bosque (Valdivia) 26: 3338.Google Scholar
Joysey, K.A. & Friday, A.E. (eds.). 1982. Problems of Phylogenetic Reconstruction. New York: Academic Press.Google Scholar
Juan, I.I., Emerson, B.C., Orom, I.I. & Hewitt, G.M. 2000. Colonization and diversification: towards a phylogeographic synthesis for the Canary Islands. Trends in Ecology and Evolution 15: 104109.CrossRefGoogle ScholarPubMed
Just, T.K. 1948. Gymnosperms and the origin of angiosperms. Botanical Gazette 110: 91103.CrossRefGoogle Scholar
Kaeiser, M. 1939. The pollen of certain Taxodiaceae. Transactions of the Illinois State Academy of Sciences 32: 9193.Google Scholar
Kaeiser, M. 1953. Microstructure of the wood of Juniperus. Botanical Gazette 115: 155162.CrossRefGoogle Scholar
Kaiser, M. 1950. Wood anatomy of podocarps. American Journal of Botany 37: 663.Google Scholar
Kajimoto, T., Matsuura, Y., Osawa, A., et al. 2003. Size-mass allometry and biomass allocation of two larch species growing on the continuous permafrost region in Siberia. Forest Ecology and Management 222: 314325.CrossRefGoogle Scholar
Kaku, S. & Salt, R.W. 1968. Relation between freezing temperature and length of conifer needles. Canadian Journal of Botany 46: 12111213.CrossRefGoogle Scholar
Kantvilas, G. & Minchin, P.R. 1989. An analysis of epiphytic lichen communities in Tasmanian cool temperate rainforest. Vegetatio 84: 99112.CrossRefGoogle Scholar
Kapscinski, S. 1947. Cis jako roslina zywicielska. Wsechswiat 9: 267272 (in Polish).Google Scholar
Karavayev, M.N. 1958. Tsuga longibracteata Cheng first found in a fossil condition in the territory of the USSR. Botanical Journal 43: 7376 (in Russian).Google Scholar
Karhu, A., Vogl, C., Moran, G.F., Bell, J.C. & Savolainen, O. 2006. Analysis of microsatellite variation in Pinus radiata reveals effects of genetic drift but no recent bottlenecks. Journal of Evolutionary Biology 19: 167175.CrossRefGoogle ScholarPubMed
Karjalainen, L & Kuuluvainen, T. 2002. Amount and diversity of coarse woody debris within a boreal forest landscape dominated by Pinus sylvestris in Vienansalo wilderness, eastern Fennoscandia. Silvae Fennica 36: 147167.Google Scholar
Katoka, R., Taniguchi, T., Ooshima, H. & Futai, K. 2008. Comparison of the bacterial communities established on the mycorrhizae formed on Pinus thunbergii root tips by eight species of fungus. Plant and Soil 304: 267275.CrossRefGoogle Scholar
Katz, C., Oren, R., Schulze, E.-D. & Milburn, J.A. 1989. Uptake of water and solutes through the twigs of Picea abies (L.) Karst. Trees 3: 3337.CrossRefGoogle Scholar
Kauffman, J.B. 1990. Ecological relationships of vegetation and fire in Pacific Northwest Forests. Pp 3952 in Walstad, J.D., Radosevich, S.R. & Sandberg, D.V. (eds.), Natural and Prescribed Fire in Pacific Northwest Forests. Corvalis, OR: Oregon State University Press.Google Scholar
Kazakou, E., Dimitrakopoulos, P.G., Baker, A.J.M., Reeves, R.D. & Troumbis, A.Y. 2008. Hypotheses, mechanisms and trade-offs of tolerance and adaptation to serpentine soils: from species to ecosystem level. Biological Reviews 83: 495508.CrossRefGoogle ScholarPubMed
Keast, J.A. 1970. Adaptive evolution and shifts in niche occupation in island birds. Biotropica 2: 6175.CrossRefGoogle Scholar
Keeler, D.K. 1984. Rock avalanches caused by earthquakes: source characteristics. Science 223: 12881290.Google Scholar
Keeler, D.K. 1984. Landslides caused by earthquakes. Geological Society of America Bulletin 95: 406421.Google Scholar
Keeley, J.E., Bond, W.J., Bradstock, R.A., Pausas, J.G. & Rundel, P.W. 2012. Fire in Mediterranean Ecosystems: Ecology, Evolution and Management. Cambridge: Cambridge University Press.Google Scholar
Kemp, E.M. & Barrett, P.J. 1975. Antarctic glaciation and early Tertiary vegetation. Nature 258(5535): 507508.CrossRefGoogle Scholar
Kemp, E.M. & Harris, W.K. 1975. The vegetation of Tertiary islands on the Ninetyeast Ridge. Nature 258: 303307.CrossRefGoogle Scholar
Kemp, E.M. & Harris, W.K. 1977. The palynology of early Tertiary sediments, Ninetyeast Ridge, Indian Ocean. Palaeontological Association of London Special Papers in Palaeontology 19: 169.Google Scholar
Kemp, M. 1959. Morphological and ontogenetic studies on Torreya californica. II. Development of the megasporangiate shoot prior to pollination. American Journal of Botany 46: 249261.CrossRefGoogle Scholar
Kendall, K.C. & Arno, S.F. 1990. Whitebark Pine: An Important But Endangered Wildlife Resource. Washington, DC: USDA.Google Scholar
Kendall, M.W. 1947. On five species of Brachyphyllum from the Jurassic of Yorkshire and Wiltshire. Annals and Magazine of Natural History 11(14): 225251.CrossRefGoogle Scholar
Kendall, M.W. 1948. On six species of Pagiophylum from the Jurassic of Yorkshire and southern England. Annals and Magazine of Natural History 12(1): 73108.CrossRefGoogle Scholar
Kendall, M.W. 1949. A Jurassic member of the Araucariaceae. Annals of Botany 8: 151161.CrossRefGoogle Scholar
Kendall, M.W. 1949. On Brachphyllum expansum (Sternberg) Seward, and its cone. Annals and Magazine of Natural History 12(2): 308320.CrossRefGoogle Scholar
Kendall, M.W. 1952. Some conifers from the Jurassic of England. Annals and Magazine of Natural History 12(5): 583594.CrossRefGoogle Scholar
Keng, H. 1963. Aspects of morphology of Phyllocladus hypophyllus. Annals of Botany n.s. 27: 6978.CrossRefGoogle Scholar
Keng, H. 1973. On the family Phyllocladacaeae. Taiwania 18: 142145.Google Scholar
Keng, H. 1979. The phylloclade of Phyllocladus and its possible bearing on the foliate organs of Coniferophytes. Botanical Bulletin, Academica Sinica 20: 917.Google Scholar
Kenton, A., Parakonny, A.S., Bennett, S.T & Bennett, M.D. 1993. Does genome organisation influence speciation? A reappraisal of karyotype studies in evolutionary biology. Pp 189256 in Lees, D.R. & Edwards, D. (eds.), Evolutionary Patterns and Processes. London: Academic Press.Google Scholar
Keppel, G. & Ghazanfar, S.A. 2006. Trees of Fiji: A Guide to 1000 Rainforest Trees. Suva, Fiji: Department of Forestry.Google Scholar
Keppel, G., Lowe, A.J. & Possingham, H.P. 2009. Changing perspectives on the biogeography of the tropical South Pacific: influences of dispersal, vicariance and extinction. Journal of Biogeography 36: 10351054.CrossRefGoogle Scholar
Kernaghan, G. & Harper, K.A. 2001. Commmunity structure of ectomycorrhizal fungi across an alpine/subalpine ecotone. Ecography 24: 181188.CrossRefGoogle Scholar
Kerp, J.A. & Clement-Westerhof, J.A.. 1991. Aspects of Permian palaeobotany and palynology. XII: the form-genus Walchiostrobus Florin reconsidered. Neues Jahrbuch für Geologie und Paläontologie, Abh 183: 257268.CrossRefGoogle Scholar
Kershaw, A.P. 1971. A pollen diagram from Quincan Crater, north-east Queensland, Australia. New Phytologist 20: 669681.CrossRefGoogle Scholar
Kershaw, A.P. & Sluiter, I.R.K. 1982. Late Cenozoic pollen spectra from the Atherton Tableland, north-east Queensland. Australian Journal of Botany 30: 279295.CrossRefGoogle Scholar
Kessler, M. 1999. Plant species richness and endemism during natural landslide succession in a perhumid montane forest in the Bolivian Andes. Ecotropica 4: 123136.Google Scholar
Kettle, C.J., Ennos, R.A., Jaffré, T. et al. 2006. Importance of demography and dispersal for the resilience and distribution of a critically endangered tropical conifer Araucaria nemorosa. Diversity and Distributions 18: 248259.CrossRefGoogle Scholar
Kettle, C.J., Ennos, R.A., Jaffré, T., Gardner, M. & Hollongsworth, P.M. 2008. Cryptic genetic bottlenecks during restoration of an endangered tropical conifer. Biological Conservation 141: 19531961.CrossRefGoogle Scholar
Khan, A.C. & Valder, P.G. 1971. The occurrence of root nodules in Ginkgoales, Taxales and Coniferales. Proceedings of the Linnean Society of New South Wales 97: 3541.Google Scholar
Khan, A.M. 1976. Palynology of tertiary sediments from Papua New Guinea. II. Gymnosperm pollen from Upper Tertiary sediments. Australian Journal of Botany 24: 783791.CrossRefGoogle Scholar
Khoshoo, T.N. 1969. Chromosome evolution in the cycads. Pp 236240 in Darlington, C.D. & Lewis, K.R. (eds.), Chromosomes Today. New York: Plenum Press.Google Scholar
Khoshoo, T.N. & Ahuja, M.R. 1963. The chromosomes and relationships of Welwitschia mirabilis. Chromosoma 14: 522533.CrossRefGoogle Scholar
Khuri, S. & Talhouk, S.N. 1999. Species accounts: cedar of Lebanon (Cedrus libani A. Rich). Pp 108111 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Kildahl, N.J. 1908. Affinities of Phyllocladus. Botanical Gazette 46: 464465.CrossRefGoogle Scholar
Kilgore, B.M. 1973. The ecological role of fire in Sierran conifer forests. Quaternary Research 3: 496513.CrossRefGoogle Scholar
Kilgore, B.M. & Taylor, D. 1979. Fire history of a Sequoia-mixed conifer forest. Ecology 60: 129142.CrossRefGoogle Scholar
Kimura, M. 1980. A simple method for estimating evolutionary rates of base substitution through comparative studies of nucleotide sequences. Journal of Molecular Evolution 16: 111120.CrossRefGoogle ScholarPubMed
Kingdon-Ward, F. 1929. The Mekong–Salween Divide as a geographical barrier. The Geographical Journal 58: 4956.CrossRefGoogle Scholar
Kingdon-Ward, F. 1946. Botanical explorations in North Burma. Journal of the Royal Horticultural Society 71: 318325.Google Scholar
Kirchhefer, A.J. 2001. Reconstructions of summer temperatures from tree-rings of Scots pine (Pinus sylvestris L.) in coastal southern Norway. Holocene 11: 4152.CrossRefGoogle Scholar
Kitamura, S. 1960. Flora of Afghanistan. Kyoto: Nissh Printing Co. Ltd.Google Scholar
Kitzberger, T., Veblen, T.T. & Villalba, R. 1995. Tectonic influences on tree growth in northern Patagonia, Argentina: the roles of substrate stability and climatic variation. Canadian Journal of Forest Research 25: 16841696.CrossRefGoogle Scholar
Kitzberger, T., Veblen, T.T. & Villalba, R. 1997. Climatic influences on fire regimes along a rain forest-to-xeric woodland gradient in northern Patagonia, Argentina. Journal of Biogeography 24: 3547.CrossRefGoogle Scholar
Klaus, W. 1980. Neue beobachtungen zur Morphologie des Zapfens von Pinus und ihre Bedeutung fur die Systematik, Fossilbestimmung, Arealgestaltung und Evolution der Gattung. Plant Systematics and Evolution 134: 137171 (in German).CrossRefGoogle Scholar
Knopf, P., Nimsch, H. & Stutzel, T. 2007. Dacrydioum x suprini, sp. Nova: a natural hybrid of Dacrydium araucarioides x D. guillauminii. Feddes Repertorium 118: 5159.CrossRefGoogle Scholar
Kobe, R.K., Pacala, S.W., Silander, J.A. & Canham, C.D. 1995. Juvenile tree survivorship as a component of shade-tolerance. Ecological Applications 5: 517532.CrossRefGoogle Scholar
Koele, N., & Hildebrand, E. 2011. Differences in growth and nutrient uptake from a coarse soil substrate by ectomycorrizal- and fungicide-treated Picea abies seedlings. European Journal of Forest Research 130: 317324.CrossRefGoogle Scholar
Kohn, M.J. 1996. Predicting animal δ18O: accounting for diet and physiological adaptation. Geochimica et Cosmochimica Acta 60: 48114829.CrossRefGoogle Scholar
Koidzumi, G. 1942. A new species of Taiwania. Acta Phytotaxonomica Geobotanica 11: 138 (in Japanese).Google Scholar
Kojo, Y. 1987. A dendrochronological study of Cryptomeria japonica in Japan. Tree-Ring Bulletin 47: 121.Google Scholar
Konar, R.N. & Banerjee, S.K. 1963. The morphology and embryology of Cupressus funebris Endl. Phytomorphology 13: 321337.Google Scholar
Konopka, B., Noguchi, K., Sakata, T., Takahashi, M. & Konopkova, Z. 2006. Fine root dynamics in a Japanese cedar (Cryptomeria japonica) plantation throughout the growing season. Forest Ecology and Management 225: 278286.CrossRefGoogle Scholar
Koop, H. 1989. Forest Dynamics: Silvistar – A Comprehensive Monitoring System. Berlin: Springer.CrossRefGoogle Scholar
Koppen, W.P. 1936. Das geographische system der klimate. In Koppen, W.P. & Geiger, R. (eds.), Handbuch der Klimatologie. Berlin: Gebruder Borntraeger.Google Scholar
Körner, C. 2000. Why are there global gradients in species richness? Mountains might hold the answer. Trends in Ecology and Evolution 15: 513514.CrossRefGoogle Scholar
Körner, C., Farquhar, G.D. & Wong, S.C. 1991. Carbon isotope discrimination by plants follows latitudinal and altitudinal trends. Oecologia 88: 3040.CrossRefGoogle ScholarPubMed
Kough, J.L., Molina, R. & Linderman, R.G. 1985. Mycorrhizal responsiveness of Thuja, Calocedrus, Sequoia and Sequoiadendron species of western North America. Canadian Journal of Forest Research 15: 10491054.CrossRefGoogle Scholar
Kovar-Eder, J., Kvaček, Z. & Meller, B. 2001. Comparing early to middle Miocene floras and probable vegetation types of Oberdorf N Voitsberg (Austria), Bohemia (Czech Republic) and Wackersdorf (Germany). Review of Palaeobotany and Palynology 114: 83125.CrossRefGoogle Scholar
Krämer, K.U. & Green, P.S. 1990. Pteridophytes and gymnosperms (overall ordinal classifications). Pp 282361 in Kubitzky, K. (ed.), The Families and Genera of Vascular Plants, Vol. 1. Berlin: Springer.Google Scholar
Krassilov, V.A. 1967. Early Cretaceous Flora of South Primorye and its Significance to Stratigraphy. Moscow: Nauka.Google Scholar
Krassilov, V.A. 1984. New palaeobotanical data on origin and early evolution of angiospermy. Annals of the Missouri Botanical Garden 71: 577592.CrossRefGoogle Scholar
Kraüsel, R. 1943. Die Ginkgophyten der Trias von Lunz in Nieder-Österreich und von neuewelt bei Basel. Palaeontographica, Abt. B. 87: 5993.Google Scholar
Kraüsel, R. & Jain, K. 1964. New fossil coniferous wood from the Rajmahal Hills, Bihar, India. Palaeobotanist 12: 5967.Google Scholar
Kritchkova, A.I. & Samylina, V.A. 1979. On the peculiarities of leaves in some Mesozoic Ginkgolaes and Czekanowskiales. Botanic Zhurnal 64: 15291538 (in Russian).Google Scholar
Krol, S. 1978. An outline of ecology. Pp 6586 in Bartkowiak, S., Bugala, W., Czartoryski, A., et al. (eds.) The Yew: Taxus baccata. Warsaw: Department of the National Center for Scientific and Technical, and Economic Information.Google Scholar
Kruckeberg, A.R. 1954. The ecology of serpentine soils: a symposium. III. Plant species in relation to serpentine soils. Ecology 35: 267274.Google Scholar
Kruckeberg, A.R. & Rabinowitz, D. 1985. Biological aspects of endemism in higher plants. Annual Review of Ecology and Systematics 16: 447479.CrossRefGoogle Scholar
Krussman, G. 1985. Manual of Cultivated Conifers (English translation by H.-D. Warda). Portland, OR: Timber Press.Google Scholar
Krystofofich, A.N. 1929. Evolution of the Tertiary flora in Asia. New Phytologist 28: 303312.CrossRefGoogle Scholar
Krystofovich, A. 1935. A final link between the Tertiary Floras of Asia and Europe (Contribution to the age of the Arcto-Tertiary Floras of the Northern Holarctic). New Phytologist 34: 339344.CrossRefGoogle Scholar
Kubitssky, K. (ed.). 1977. Flowering Plants, Evolution and Classification of Higher Categories. Vienna: Springer.Google Scholar
Kudo, Y. & Yamamoto, Y. 1931. Amentotaxaceae. Journal of the Society for Tropical Agriculture (Taihoku) 3: 110111.Google Scholar
Kumagi, H., Sweda, T., Hayashi, K., et al. 1995. Growth-ring analysis of Early Tertiary conifer woods from the Canadian High Arctic and its paleoclimatic interpretation. Palaeogeography, Palaeoclimatology, Palaeoecology 116: 247262.CrossRefGoogle Scholar
Kunzmann, L. 1999. Koniferten der Oberkreide und ihre Relikte im Tertiar Europas: ein Beitrag zur Kenntnis ausgestornener Taxodiaceae und Geinitziaceae fam nov. Abhandlungen des Staatlichen Museums fur Minerologie und Geologie zu Dresden 45: 1190.Google Scholar
Kunzmann, L. 2007. Neue Untersuchungen zu Araucaria Jussieu aus der europaiscchen Kreide. Palaeontogrphica Abteilung B. 276: 97131.CrossRefGoogle Scholar
Kuo, Y.H. & Chen, W.C. 1999. Chemical constituents of the pericarp of Platycladus orientalis. Journal of the Chinese Chemical Society (Taipei) 46: 819824.CrossRefGoogle Scholar
Kurmann, M.H. 1994. Pollen morphology and ultrastructure in the Cupressceae. Acta Botanica Gallica 141: 141147.CrossRefGoogle Scholar
Kurz, H. 1938. Torreya west of the Apalachicola river. Proceedings of the Florida Academy of Science 3: 6667.Google Scholar
Kurz, S. 1873. On a few new plants from Yunnan. Journal of Botany, British and Foreign 11: 193196.Google Scholar
Kuser, J.E. 1999. Metasequoia glyptostroboides: fifty years growth in North America. Arnoldia 58: 7679.CrossRefGoogle Scholar
Kuser, J.E., Sheely, D.L. & Hendricks, D.R. 1997. Genetic variation in two ex-situ collections of the rare Metasequoia glyptostroboides (Cupressaceae). Silvae Genetica 46: 258264.Google Scholar
Kush, J.S., Meldahl, R.S., McMahon, C.K. & Boyer, W.D. 2004. Longleaf pine: a sustainable approach for increasing terrestrial carbon in the southern United States. Environmental Management 33 (suppl. 1): 139147.CrossRefGoogle Scholar
Kusumi, J., Tsumura, Y., Yoshimaru, H. & Tachida, H. 2000. Phylogenetic relationships in Taxodiaceae and Cupressaceae sensu stricto based on matK gene, chlL gene, trn–-trnF IGS region and trnL intron sequences. American Journal of Botany 87: 14801488.CrossRefGoogle ScholarPubMed
Kutty, T.S., Chaterjee, S., Galton, P.M. & Upchurch, P. 2007. Basal sauropodomorphs (Dinosauria: Saurischia) from the Lower Jurassic of India: their anatomy and relationships. Journal of Paleonotology 81: 12181240.CrossRefGoogle Scholar
Kwei, Y.L. & Lee, C.L. 1963. Anatomical studies of the leaf structure of Chinese pines. Acta Botanica Sinica 11: 4458 (in Chinese).Google Scholar
La Mana, L. & Rajchenberg, M. 2004. Soil properties and Austrocedrus chilensis forest decline in central Patagonia, Argentina. Plant & Soil 263: 2941.CrossRefGoogle Scholar
Lacassagne, M. 1934. Etude morphologique, anatomique et systematique du genre Picea. Travaux Laboratoire Forestiere Toulouse 3: 1292.Google Scholar
Lacey, W.S. & Lucas, R.C. 1981. The Triassic flora of Livingston Island, South Shetland Islands. British Antarctic Survey Bulletin 53: 157173.Google Scholar
Ladiges, P.Y. 2008. Synthesizing biotic patterns and geology for New Caledonia. Journal of Biogeography 35: 21512152.CrossRefGoogle Scholar
Ladiges, P.Y. & Cantrill, D. 2007. New Caledonia–Australian connections: biogeographic patterns and geology. Australian Systematic Botany 20: 383389.CrossRefGoogle Scholar
Lagabrielle, Y., Godderis, Y. & Donnadieu, Y. 2009. The tectonic history of Drake Passage and its possible impacts on global climate. Earth and Planetary Science Letters 279: 197211.CrossRefGoogle Scholar
Lakhanpal, R.N., Gileria, J.S. & Awasthi, N. 1975. A podocarpaceous wood from the Pliocene of Kutch, India. Geophytology 5: 172177.Google Scholar
LaMarche, V.C. & Hirschboeck, K.K. 1984. Frost rings in trees as records of major volcanic eruptions. Nature 307: 121126.CrossRefGoogle Scholar
Lambers, H., Chapin, F.S. & Pons, T.L. 1998. Plant Physiological Ecology. New York: Springer.CrossRefGoogle Scholar
Lamont, B. 1982. Mechanisms for enhancing nutrient uptake in plants, with particular reference to Mediterranean, South Africa and Western Australia. Botanical Review 48: 597689.CrossRefGoogle Scholar
Land, W.J.G. 1902. A morphological study of Thuja. Botanical Gazette 34: 249259.CrossRefGoogle Scholar
Lande, R. 1988. Genetics and demography in biological conservation. Science 241: 14551460.CrossRefGoogle ScholarPubMed
Landis, C.A., Campbell, H.J., Begg, J.B., et al. 2008. The Waipounamu Erosion Surface: questioning the antiquity of the New Zealand land surface and terrestrial fauna and flora. Geological Magazine 145: 173197.CrossRefGoogle Scholar
Landis, M.R. & Peart, D.R. 2005. Early performance predicts canopy attainment across life histories in sub-alpine forest trees. Ecology 86: 6372.CrossRefGoogle Scholar
Landis, R.M., Gurevitch, J., Fox, G.A., Fang, W. & Taub, D.R. 2005. Variation in recruitment and early demography in Pinus rigida following crown fire in the pine barrens of Long Island, New York. Journal of Ecology 93: 607617.CrossRefGoogle Scholar
Langenheim, J.H. 2003. Plant Resins: Chemistry, Evolution, Ecology and Ethnobotany. Portland, OR: Timber Press.Google Scholar
Lanner, R.M. 1996. Made for Each Other: A Symbiosis of Birds and Pines. New York: Oxford Univerity Press.CrossRefGoogle Scholar
Lanner, R.M. 2003. Made for each other: Clark’s nutcracker and the whitebark pine. Acta Horticulturae 615: 121125.CrossRefGoogle Scholar
Lanner, R.M. & Phillips, A.M. III. 1992. Natural hybridization and introgression of pinyon pines in northwestern Arizona. International Journal of Plant Sciences 153: 250257.CrossRefGoogle Scholar
Lanner, R.M. & Vander Wall, S.B. 1980. Dispersal of limber pine seed by Clark’s nutcracker. Journal of Forestry 78: 637639.CrossRefGoogle Scholar
Lanner, R.M., Hutchins, H.E. & Lanner, H.A. 1988. Bristlecone pine and Clark’s nutcracker: probable interaction in the White Mountains, California. Great Basin Naturalist 44: 357360.Google Scholar
Lara, A. & Vilalba, R. 1993. A 3,622 year temperature reconstruction from Fitzroya cupressoides tree rings in southern South America. Science 260: 11041106.CrossRefGoogle Scholar
Larson, D.W., Matthes, U. & Kelly, P.E. 2000. Flora. In: Cliff Ecology: Pattern and Process in Cliff Ecosystems. Cambridge Studies in Ecology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Latham, M. 1974. Nouvelle observation de la coupe de Nepoui (Nouvelle-Calédonie); consequence sur la chronologie de l’etagement des nivaux curirasses sur les massifs de roches ultrabasiques. C.R. Acadamie Science Paris 279: 10551058.Google Scholar
Lawrance, W.F. & Bierregaard, R.O. (eds.). 1997. Tropical Forest Remnants: Ecology, Management and Conservation of Fragmented Communities. Chicago, IL: University of Chicago Press.Google Scholar
Lawson, A.A. 1907. The gametophyte and embryo of the Cupressineae with special reference to Libocedrus decurrens. Annals of Botany 21: 281301.CrossRefGoogle Scholar
Lawson, A.A. 1923. The life history of Microcachrys tetragona (Hook.). Proceedings of the Linnean Society of New South Wales. 48: 177195, 499–514.Google Scholar
Le Duc, A., Adams, R.P. & Zhong, M. 1999. Using random amplification of polymorphic DNA for a taxonomic reevaluation of Pfitzer junipers. Hort Science 34: 11231125.Google Scholar
Leavitt, S.W. 1994. Major wet interval in White Mountains Medieval Warm Period evidences in delta 13C of bristlecone pine tree rings. Climate Change 26: 29307.Google Scholar
Leckie, D.G., Cloney, E. & Joyce, S.P. 2005. Automated detection and mapping of crown discoloration caused by jack pine budworm, with 2.5m resolution multispectral imagery. International Journal of Applied Earth Observation and Geoinformation 7: 6177.CrossRefGoogle Scholar
Ledig, F.T. 1987. Genetic structure and the conservation of California’s endemic and near-endemic conifers. Pp 587594 in Ellis, T.S. (ed.), Conservation and Management of Rare and Endangered Plants. Sacremento, CA: California Native Plant Society.Google Scholar
Ledig, F.T. 2000. Founder effects and the genetic structure of Coulter pine. Journal of Heredity 91: 307315.CrossRefGoogle Scholar
Ledig, F.T., Conckle, B., Bermejo-Valaquez, B., et al. 1999. Evidence for an extreme bottleneck in a rare Mexican pinyon: genetic diversity, disequilibrium, and the mating system in Pinus maximartinezii. Evolution 53: 9199.CrossRefGoogle Scholar
Lee, D.E., Lee, W.G. & Mortimer, N. 2001. Depletion and turnover in the New Zealand Cenozoic angiosperm flora in relation to palaeogeography and climate. Australian Journal of Botany 49: 341356.CrossRefGoogle Scholar
Lees, D.R. & Edwards, D. (eds.). 1993. Evolutionary Patterns and Processes. London: Academic Press.Google Scholar
Leigh, E.G. 1975. Structure and climate in tropical rain forest. Annual Review of Ecology and Systematics 6: 6786.CrossRefGoogle Scholar
Leigh, J., Hodge, A. & Fitter, J.H. 2009. Arbuscular mycorrhizal fungi can transfer substantial amounts of nitrogen to their host plant from organic material. New Phytologist 181: 199207.CrossRefGoogle ScholarPubMed
Leistner, O.A. 1966. Podocarpaceae. Flora of South Africa 1: 3441.Google Scholar
Lemoine-Sebastian, O. 1967. Appareil reproducteur male des Juniperus. Travaux Laboratoire Forestiere Toulouse T1 6: 135.Google Scholar
Lemoine-Sebastian, O. 1968. L’inflorescence femelle des Juniperus. Travaux Laboratoire Forestiere Toulouse T1 7: 1456.Google Scholar
Lemoine-Sebastian, C. 1968. La vascularisation du comlexe bractee-ecaille chez les Taxodiacees. Travaux des Laboratoire Forestiere de Toulouse 7: 122.Google Scholar
Leng, Q., Yang, H., Yang, Q. & Zhou, J. 2001. Variation of cuticle micromorphology of Metasequoia glyptostroboides (Taxodiaceae). Botanical Journal of the Linnean Society 136: 207219.CrossRefGoogle Scholar
Leopold, E.B. & Liu, G.G. 1994. A long pollen sequence of Neogene age, Alaska Range. Quaternary International 22–23: 103140.CrossRefGoogle Scholar
LePage, B.A., Yang, H. & Matsumoto, M. 2005. The evolution and biogeographic history of Metasequoia. Pp 3114 in LePage, B.A., Williams, C.J. & Yang, H. (eds.), The Geobiology and Ecology of Metasequoia: With Contributions from the 1st International Metasequoia Symposium. Dordrecht: Springer.CrossRefGoogle Scholar
Leski, T. & Rudawska, M. 2012. Ectomycorrhizal fungal community of naturally regenerated European larch (Larix decidua) seedlings. Symbiosis 56: 4553.CrossRefGoogle Scholar
Lester, S.E., Ruttenberg, B.I., Gaines, S.D. & Kinlan, B.P. 2007. The relationship between dispersal ability and geographic range size. Ecology Letters 10: 745758.CrossRefGoogle ScholarPubMed
Levisohn, I. 1954. Occurrence of ectotrophic and endotrophic mycorrhizas in forest trees. Forestry 27: 145146.CrossRefGoogle Scholar
Lewandowski, A., Burczyk, J. & Mejnartowicz, L. 1995. Genetic structure of English yew (Taxus baccata L.) in the Wierzchlas Reserve: implications for genetic conservation. Forest Ecology and Management 73: 221227.CrossRefGoogle Scholar
Li, B.-D. & Liu, Z.-T. 1984. The distribution pattern of Pinus yunnanensis. Journal of Yunnan University 1: 3346 (in Chinese).Google Scholar
Li, H. 1967. Notes on the nomenclature and taxonomy of Pseudolarix. Taiwania 13: 147152.Google Scholar
Li, H. 1968. The golden larch, Pseudolarix amabilis. Morris Arboretum Bulletin 19: 1925.Google Scholar
Li, H.-L. 1950. The Coniferales of Taiwan. Taiwania 1: 285310.Google Scholar
Li, H.-L. 1953. A reclassification of Libocedrus and Cupressaceae. Journal of the Arnold Arboretum 34: 1736.CrossRefGoogle Scholar
Li, H.-L. & Keng, H. 1954. Icones Gymnospermae Formosarum. Taiwania 5: 2584.Google Scholar
Li, L.-C. 1986. Cytological investigation of Taiwania flousiana Gaussen (Taxodiaceae). Acta Phytotaxonomica Sinica 24: 376381 (in Chinese with English abstract).Google Scholar
Li, L.-C. 1986. Karyotype study of Metasequoia glyptostroboides. Journal of Wuhan Botanical Research 4: 15 (in Chinese with English abstract).Google Scholar
Li, L.C. 1988. The systematic position of Sciadopitys Sieb. & Zucc. based on cytological data. Guhaia 8: 135141.Google Scholar
Li, L.-C. 1992. Karyoptype analysis of Athrotaxis (Taxodiaceae) and its systematic position. Acta Phytotaxonomica Sinica 30: 331341.Google Scholar
Li, L.-C. 1998. A cytotaxonomic study of the family Cupressaceae. Acta Botanica Yunnanica 20: 197203.Google Scholar
Li, L.-C. & Hsu, P.S. 1984. Karyotype analyses in Platycladus orientalis and Fokienia hodginsii. Acta Botanica Yunnanica 6: 447451 (in Chinese, with English summary).Google Scholar
Li, L.-C., Cen, Y., Xu, P. & Wang, H. 1996. Studies on the karyotypes of Chamaecyparis and the cytotaxonomy of Cupressaceae. Acta Botanica Yunnanica 18: 7276.Google Scholar
Li, L.-C., Lui, Y.-Q., Wang, Y.-Q. & Lui, G. 1996. Studies on the karyotypes of three species and the cytotaxonomy of Thujoideae (Cupressaceae). Acta Botanica Yunnanica 18: 437444.Google Scholar
Li, L.-C., Jiang, J.-H., Wang, Y.-Q. & Wang, G. 1997. The karyotype analysis of three species in Cupressaceae. Acta Botanica Yunnanica 19: 391394.Google Scholar
Li, S.-J. 1975. Reproductive biology of Chamaecyparis: 2. Pollen development and pollination mechanism. Taiwania 20: 139146.Google Scholar
Li, X., Huang, H. & Li, J. 2003. Genetic diversity of the rare plant Metasequoia glyptostroboides. Biodiversity Science 11: 100108.Google Scholar
Li, X.H., He, S.A. & Sheng, N. 1999. Establishment of a natural population during the ex-situ conservation of Taxus chinensis. Journal of Plant Resources and Environment 8: 3841.Google Scholar
Li, X.K., Huang, Y.Q., & Su, Z.M. 2000. Distribution pattern and its dynamics of Taxus chinensis var. mairei population on Yunbaoshan Mountain. Chinese Journal of Applied Ecology 11: 169172.Google ScholarPubMed
Li, X.K., Su, Z.M. & Huang, Y.Q. 2001. Studies on Taxus chinensis var. mairei population and its community. Journal of Nanjing Forestry University 25: 2328.Google Scholar
Li, X.-W. 1992. A new series and a new species of Pinus from Yunan. Acta Botanica Yunnanica 14: 259260.Google Scholar
Li, X.W. & Li, J. 1992. On the validity of the Tanaka line and its significance viewed from the distribution of eastern Asiatic genera in Yunnan. Acta Botanica Yunnanica 14: 112.Google Scholar
Li, X.W. & Li, J. 1997. The Tanaka-Kaiyong line: an important floristic line for the study of the flora of east Asia. Annals of the Missouri Botanical Garden 84: 888892.Google Scholar
Li, Y.-J., Wang, Y.-F., & Li, C.-S. 2008. Studies on fossil Metasequoia from north-east China and their taxonomic implications. Botanical Journal of the Linnean Society 130: 267297.Google Scholar
Li, Z.F., Zhou, D.X. & An, P. 1995. A study on the variant types of karyotypes for Fokienia hodginsii (Dunn.) Henry et Thomas. Scientia Silvae Sinicae 31: 215219.Google Scholar
Liang, H., Chow, K.Y., & Au, C.N. 1948. Properties of a ‘living fossil’ wood (Metasequoia glyptostroboides). Wood Technology 1: 16.Google Scholar
Liang, H.-P. 1990. Karyotype analysis in Cupressus gigantea. Acta Botanica Sinica 32: 653656.Google Scholar
Libby, W.J., Bannister, M.H. & Linhart, Y.B. 1968. The pines of Cedros and Guadalupe islands. Journal of Forestry 66: 846853.Google Scholar
Liguo, F., Li, N. & Mill, R.R. 1999. Pinaceae. In Wu, Z. & Raven, P.H. (eds.), Flora of China, Vol 4. Beijing: Science Press.Google Scholar
Lin, T.P., Wang, C.-T. & Yang, J.-C. 1998. Comparison of genetic diversity between Cunninghamia konishii and C. lanceolata. Journal of Heredity 89: 370373.CrossRefGoogle Scholar
Lin, X. & Zhang, D.H. 2004. Analysis for the origin of Ginkgo population in Tianmu Mountains. Scientific Silviculture Sinensis 40: 2831 (in Chinese, with English abstract).Google Scholar
Lindberg, C.J., Shaffer, G.P., Wood, W.B. & Day, J.W. 2011. Growth rates of bald cypress (Taxodium distichum) seedlings in a treated effluent-associated marsh. Ecological Engineering 37: 549553.CrossRefGoogle Scholar
Linder, H.P. 2003. The radiation of the Cape flora, southern Africa. Biological Reviews 78: 597638.CrossRefGoogle ScholarPubMed
Linder, H.P. & Hardy, C.R. 2004. Evolution of the species-rich Cape flora. Philosophical Transactions of the Royal Society London B 359: 16231632.CrossRefGoogle ScholarPubMed
Lindquist, B.L. 1948. The new varieties of Picea abies in Europe. Acta Horti Bergiani 14: 249342.Google Scholar
Lindqvist, J.K. 1986. Teredinid-bored Araucariaceae logs preserved in shoreface sediments, Wangaloa Formation (Paleocene), Otago, New Zealand. New Zealand Journal of Geology and Geophysics, 29: 253261.CrossRefGoogle Scholar
Ling, Y., Lu, W., Wang, Y., Chen, J. & Zhang, W. 2006. PCR-RFLP and AP-PCR of rbcL and ITS of rDNA show that xTaxodiomeria peizhongii (Taxodium x Cryptomeria) is not an intergeneric hybrid. Journal of Integrated Plant Biology 48: 468472.CrossRefGoogle Scholar
Linhart, Y.B. 1978. Maintenance of variation in cone morphology in California closed-cone pines: the role of fire, squirrels and seed output. The Southwestern Naturalist 23: 2940.CrossRefGoogle Scholar
Linnard, W. 1966. A note on the wood of Metasequoia. Wood 31: 46.Google Scholar
Liston, A. 1997. Biogeographic relationships between the Mediterranean and North American floras: insights from molecular data. Lagascalia 19: 323330.Google Scholar
Liston, A., Robinson, W.A., Oliphant, J.M. & Álvarez-Bullya, E.R. 1996. Length variation in the nuclear ribosomal DNA internal transcriber spacer region on non-flowering seed plants. Systematic Botany 21: 109120.CrossRefGoogle Scholar
Liston, A., Robinson, W.A., Piñero, D. & Álvarez-Buylla, E.R. 1999. Phylogenies of Pinus (Pinaceae) based on nuclear ribosomal DNA internal transcribed spacer region sequences. Molecular Phylogenetics and Evolutions 11: 95109.CrossRefGoogle ScholarPubMed
Liston, A., Parker-Defeniks, M., Syring, J.V., Willyard, A., & Cronn, R. 2007. Interspecific phylogenetic analysis enhances intraspecific phylogeographical inference: a case study in Pinus lambertiana. Molecular Ecology 16: 39263937.CrossRefGoogle ScholarPubMed
Little, D.P. 2000. Phylogenetic relationships and monophyly of Cupressus and Chamaecyparis (Cupressaceae) molecular and organismal evidence. American Journal of Botany 87: 139.Google Scholar
Little, E.L. 1953. A natural hybrid spruce in Alaska. Journal of Forestry 51: 745747.Google Scholar
Little, E.L. 1966. Varietal transfers in Cupressus and Chamaecyparis. Madrõno 18: 161192.Google Scholar
Little, E.L. 1970. Names of New World cypresses (Cupressus). Phytologia 20: 429445.CrossRefGoogle Scholar
Liu, S., Chi, X.G., Li, C. & Yang, R.H. 2001. The summarising for the forming and uplift mechanism of Quinghai–Tibet Plateau. World Geology 20: 105112.Google Scholar
Liu, T.S. 1966. Study of the Phytogeography of the Conifers and Taxads of Taiwan. Taipei: Taiwan Forestry Research Institute.Google Scholar
Liu, Y.-J. 1999. Studies on fossil Metasequoia from north-east China and their taxonomical implications. Botanical Journal of the Linnean Society 130: 267297.Google Scholar
Liu, Y.S., Guo, S.X. & Ferguson, D.K. 1996. Catalogue of the Cenozoic megafossil plants in China. Palaeontographica Abt. B. 238: 141179.Google Scholar
Liu, Z.-L., Zhang, D., Wang, X.-Q., Ma, X.-F., & Wang, X.-R. 2003. Intergenomic and interspecific 5S rDNA sequence variation in five Asian pines. American Journal of Botany 90: 1724.CrossRefGoogle Scholar
Llorens, L., Llusia, J., Murchie, E.H., Penuelas, J. & Beerling, D.J. 2009. Monoterpenes emissions and photoinhibition of ‘living fossil’ trees under CO2 enrichment in a simulated Cretaceous polar environment. Journal of Geophysical Research 114.CrossRefGoogle Scholar
Llorens, L., Osborne, C.P. & Beerling, D.J. 2009. Water-use response of ‘living fossil’ conifers to CO2 enrichment in a simulated Cretaceous polar environment. Annals of Botany 104: 179188.CrossRefGoogle Scholar
Loehle, C. 1989. Tree life history strategies: the role of defenses. Canadian Journal of Forest Research 18: 209222.CrossRefGoogle Scholar
Looby, W.J. & Doyle, J. 1937. Fertilisation and proembryo formation in Sequoia. Scientific Proceedings of the Royal Dublin Society 21: 457476.Google Scholar
López, R., Climente, J. & Gill, L. 2008. From desert to cloud-forest: the non-trivial phenotypic variation of Canary Island pine needles. Trees 22: 843849.CrossRefGoogle Scholar
Lorenz, K. & Lai, R. 2010. Carbon Sequestration in Forest Ecosystems. New York: Springer.CrossRefGoogle Scholar
Losos, J.B. 2008. Phylogenetic niche conservatism, phylogenetic signal and the relationship between phylogenetic relatedness and ecological similarity among species. Ecology Letters 11: 9951003.CrossRefGoogle ScholarPubMed
Lotova, L.L. 1975. On the correlation of the anatomical features of the wood and phloem in the Pinaceae. Journal Moscow University 1: 4151 (in Russian).Google Scholar
Loustau, D., Ben Brahim, M., Gaudillère, J.P. & Dreyer, E. 1999. Photosynthetic responses to phosphorous nutrition in two-year-old maritime pine seedlings. Tree Physiology 19: 707715.CrossRefGoogle Scholar
Love, A. 1951. Taxonomical evaluation of polyploids. Caryologia 3: 263284.CrossRefGoogle Scholar
Loveless, A.R. 1961. A nutritional interpretation of sclerophylly based on differences in the chemical composition of sclerophyllous and mesophytic leaves. Annals of Botany 25: 168184.CrossRefGoogle Scholar
Loveless, A.R. 1962. Further evidence to support a nutritional interpretation of sclerophylly. Annals of Botany 26: 551561.CrossRefGoogle Scholar
Lovett, J., Page, C.N. & Whitmore, T.C. 1986. Araucariaceae. The European Garden Flora. 1: 7273.Google Scholar
Lowry, J.B. 1972. Anthocyanins of the Podocarpaceae. Phytochemistry 11(2): 725731.CrossRefGoogle Scholar
Lowry, P.P. II,, Munzinger, J., Bouchet, P., et al. 2004. New Caledonia. Pp 193197 in Mittermeier, R.A. (ed.), Hotspots Revisited: Earth’s Biologically Richest and Most Threatened Terrestrial Ecosystems. Mexico: CEMEX.Google Scholar
Lu, S.-Y., Peng, C.I., Cheng, Y.-P., Hong, K.-H. & Chiang, T.-Y. 2001. Chloroplast phylogeography of Cunninghamia konishii (Cupressaceae), and an endemic conifer of Taiwan. Genome 44: 797807.CrossRefGoogle Scholar
Lusk, C.H. 1996. Stand dynamics of the shade-tolerant conifers Saxegothaea conspicua and Podocarpus nubigena in Chilean temperate rainforest. Journal of Vegetation Science 7: 549558.CrossRefGoogle Scholar
Lusk, C.H. 2001. Leaf life spans of some conifers of the temperate forests of South America. Revista Chilena Historia Natural 74: 711718.CrossRefGoogle Scholar
Lusk, C.H. 2008. Constraints on the evolution and geographic range of Pinus. New Phytologist 178: 13.CrossRefGoogle Scholar
Lusk, C.H., Duncan, R.P. & Bellingham, P.J. 2009. Light environments occupied by conifer and angiosperm seedlings in a New Zealand podocarp-broadleaved forest. New Zealand Journal of Ecology 33: 8389.Google Scholar
Lyell, C. 1830–1832. Principles of Geology, Being an Attempt to Explain the Former Changes of the Earth’s Surface, by Reference to Causes of Operation, Vols 1–3. London: John Murray.CrossRefGoogle Scholar
Ma, J. & Shao, G. 2003. Rediscovery of the first collection of the ‘Living Fossil’ Metasequoia glyptostroboides. Taxon 52: 585588.CrossRefGoogle Scholar
Ma, J.-S. 2003. The chronology of the ‘living fossil’ Metasequoia glyptostroboides (Taxodiaceae): a review (1943–2003). Harvard Papers in Botany 8: 918.Google Scholar
Ma, Q., Li, C.S., Li, F.L. & Vickulin, S.V. 2004. Epidermal structures and stomatal parameters of Chinese endemic Glyptostrobus pensilis K. Koch (Taxodiaceae). Botanical Journal of the Linnean Society 146: 153162.CrossRefGoogle Scholar
Ma, Q.-W. & Li, C.-S. 2002. Epidermal structures of Sequoia sempervirens (D.Don) Endl. (Taxodiaceae). Taiwania 47: 194202.Google Scholar
Ma, Q.-W. & Li, C.-S. 2002. Characteristics of reference in classifying fossil plants of genera Metasequoia and Sequoia by epidermal features. Journal of Wuhan Botanical Research 20(6): 413416 (in Chinese with English abstract).Google Scholar
Ma, Q.W. & Zhang, J.-B. 2003. Epidermal structure of Metasequoia glyptostroboides Hu et Cheng (Taxodiaceae). Bulletin of Botanical Research 23: 3235 (in Chinese with English abstract).Google Scholar
Ma, Q.W., Li, F.L. & Li, C.S, 2005. Leaf epidermal structure of stomatal parameters of the genus Taxodium (Taxodiaceae). Acta Phytotaxonomica Sinica 43: 512525 (in Chinese, with English abstract).CrossRefGoogle Scholar
Macario-Mendoza, P.A., Torres-Petch, S.A. & Cabrero-Cano, E.F. 1998. Estructura y composicion de una comunidad con Pinus caribaea var. hondurensis (Senecl.) Barr. y Golf., en el Estado de Quintana Roo, Mexico. Caribbean Journal of Science 34: 5057.Google Scholar
MacArthur, R.H. & Wilson, E.O. 1967. The Theory of Island Biogeography. Princeton, NJ: Princeton University Press.Google Scholar
MacDougall, A.S., Wilson, S.D. & Bakker, J.D. 2008. Climatic variability alters the outcome of long-term community assembly. Journal of Ecology 96: 346354.CrossRefGoogle Scholar
MacKinnon, K., Hatta, G., Halim, H. & Mangalik, A. 1996. The Ecology of Kalimantan. Singapore: Perilus Editions.Google Scholar
Macnair, M.R. 1987. Heavy metal tolerance in plants: a model evolutionary system. Trends in Ecology and Evolution 2: 354359.CrossRefGoogle ScholarPubMed
Macnair, M.R. 1989. The potential for rapid speciation in plants. Genome 31: 203210.CrossRefGoogle Scholar
Macovei, G. & Givulescu, R. 2006. The present stage in the knowledge of the fossil flora at Chiuzbaia, Maramures, Romania. Carpathian Journal of Earth and Environmantal Sciences 1: 4152.Google Scholar
Macphail, M.K. 1979. Vegetation and climates of southern Tasmania since the last glaciation. Quaternary Research 11: 306341.CrossRefGoogle Scholar
Macphail, M.K., Hill, R.S. & Wells, P.M. 1991. A Late Oligocene–Early Miocene cool climate flora in Tasmania. Alcheringa 15: 87106.CrossRefGoogle Scholar
Macphail, M.K., Jordan, G.J. & Hill, R.S. 1993. Key periods in the evolution of the flora and vegetation in western Tasmania I. The Early Middle Pleistocene. Australian Journal of Botany 41: 673707.CrossRefGoogle Scholar
Macphail, M., Hill, K., Partridge, A., Truswell, E. & Foster, C. 1995. Australia: Wollemi pine – old pollen records for a newly discovered genus of gymnosperm. Geology Today March–April: 42–44.Google Scholar
MacPhail, M.K., Colhoun, E.A. & Fitzsimons, S.J. 1995. Key periods in the Cenozoic vegetation and flora in western Tasmania. Australian Journal of Botany 43: 505526.CrossRefGoogle Scholar
Maekawa, F. 1954. Phylogenetic considerations in conifer taxonomy: Taxodiaceae. Journal of Japanese Botany 29: 307313 (in Japanese).Google Scholar
Maekhofer, L., Wilf, P., Kooyman, R.M., et al. 2015. Resolving analogs for an Eocene Patagonian paleorainforest using leaf size and floristics. American Journal of Botany 102: 11601173.CrossRefGoogle Scholar
Mägdefrau, K. 1956. Paläobiologie der Pflanzen. Jana: Gustav Fischer.Google Scholar
Maheshwari, H. 1972. Permian wood from Antarctica and revision of some Lower Gondwana wood taxa. Palaeontorraphia Bohemich 203: 1120.Google Scholar
Mai, D.H. 1989. Development of regional differentiation of the European vegetation during the Tertiary. Plant Systematics and Evolution 162: 7981.CrossRefGoogle Scholar
Mai, D.H. 1999. The Middle and Upper Miocene floras of the Meuro and Rauno sequences in the Lusatica region. I. Waterferns, conifers and monocotyledons. Palaeontographica B 256: 168.Google Scholar
Major, J. 1977. California climate in relation to vegetation. Pp 1174 in Barbour, M. & Major, J. (eds.), Terrestrial Vegetation of California. New York: Wiley.Google Scholar
Malusa, J. 1992. Phylogeny and biogeography of the pinyon pines (Pinus subsect. cembroides). Systematic Botany 17: 4266.CrossRefGoogle Scholar
Manders, P.T. 1987. Application of a transition matrix model to the population dynamics of the Clanwilliam cedar (Widdringtonia cedarbergensis) in natural stands subject to fire. Forest Ecology and Management 120: 171186.CrossRefGoogle Scholar
Manum, S. 1987. Mesozoic Sciadopitys-like leaves with observations on four species from the Jurassic of Andoya, Northern Norway, an amendation of Sciadopityoides sveshnikova. Reviews in Palaeobotany and Palynology 51: 145168.CrossRefGoogle Scholar
Manum, S.B. 1962. Studies in the Tertiary Flora of Spitsbergen. Oslo: Norske Polarinstitutt.Google Scholar
Mapes, G. 1987. Ovule inversion in the earliest conifers. American Journal of Botany 74: 12051210.CrossRefGoogle Scholar
Mapes, G. & Rothwell, G.W. 1984. Permineralised ovulate cones of Lebachia from Late Paleozoic limestones of Kansas. Palaeontology 27: 6994.Google Scholar
Mapes, G. & Rothwell, G.W. 1988. Diversity among Hamilton conifers. Pp 225244 in Mapes, G. & Mapes, R.H. (eds.), Regional Geology and Paleontology of Upper Paleozoic Hamilton Quarry Area in Southeastern Kansas. Lawrence, KS: Kansas Geological Survey.Google Scholar
Mapes, G. & Rothwell, G.W. 1991. Structure and relationships of primitive conifers. Neues Jahrbuch für Geologie und Paläontologie, Abh 183: 269287.CrossRefGoogle Scholar
Marion, A.F. 1884. Dolisotrobus sternbergi, nouveau genre de conifers fossiles Tertaires. Annales des Sciences Géologiques Paris 20(3): 120.Google Scholar
Markewitz, D., Sartori, F. & Craft, C. 2002. Soil change and carbon storage in longleaf pine stands on marginal agricultural lands. Ecological Applications 12: 12761285.CrossRefGoogle Scholar
Markwick, P.J. 1998. Fossil crocodilians as indicators of late Cretaceous and Cenozoic climates: implications for using paleontological data in reconstructing microclimates. Palaeogeography, Palaeoclimatology, Palaeoecology 137: 205271.CrossRefGoogle Scholar
Marrs, R.H., Proctor, J., Heaney, A. & Mountford, M.D. 1988. Changes in soil nitrogen-mineralisation and nitrification along an altitudinal transect in tropical rain forest in Costa Rica. Journal of Ecology 76: 466482.CrossRefGoogle Scholar
Marshall, J.D. & Zhang, J. 1994. Carbon isotope discrimination and water-use efficiency in native plants of the north-central Rockies. Ecology 75: 18871895.CrossRefGoogle Scholar
Marticorena, C. & Rodríguez, R. 1995. Flora de Chile: Vol I. Pteridophyta-Gymnospermae. Concepción: Universidad de Concepción.Google Scholar
Martin, H.A. 1973. Upper Tertiary palynology in southern New South Wales. Geological Society of Australia Special Publications 4: 3554.Google Scholar
Martin, H.A. 1982. Changing Cenozoic barriers and the Australian palaeobotanical record. Annals of the Missouri Botanical Garden 69: 625667.CrossRefGoogle Scholar
Martin, H.A. 1992. The Tertiary of southeastern Australia: was it tropical? Palaeobotanist 39: 270280.Google Scholar
Martin, H.A. 2006. Cenozoic climatic change and the development of the arid vegetation in Australia. Journal of Arid Environments 66: 533563.CrossRefGoogle Scholar
Martínez, M. 1942. Una nueva Pinacea Mexicana. Anales del instituto de Biología de la Universidad Nacional de México 13: 3134.Google Scholar
Martínez, M. 1946. Los Juniperus Mexicanos. Anales del Instituto de Biológia de la Universidad Nacional de México 17: 3128.Google Scholar
Martínez, O. 1981. Flora y fitosociologia de un relicto de Pilgerodendron uviferum (D.Don) Florin en el fundo San Pablo de Tregua (Valdivia, Chile). Bosque 4: 311.CrossRefGoogle Scholar
Martínez, O. & Munoz, A.M. 1988. Aspectos conservatives de las coniferas Chilienas. Bosque 9: 7782.CrossRefGoogle Scholar
Mason, H.L. 1930. The Santa Cruz Island pine. Madrõno 2: 10.Google Scholar
Mason, H.L. 1932. A phylogenetic series of the California closed-cone pines suggested by the fossil record. Madrõno 2: 4956.Google Scholar
Mastroberti, A.A. & Araujo Mariath, J.E. 2003. Compartmented cells in the mesophyll of Araucaria angustifolia (Araucariaceae). Australian Journal of Botany 51: 267274.CrossRefGoogle Scholar
Mastroberti, A.A. & Araujo Mariath, J.E. 2008. Development of mucilage cells of Araucaria angustifolia (Araucariaceae). Protoplasma 232: 233245.CrossRefGoogle Scholar
Mathews, S. 2009. Phylogenetic relationships among seed plants: persistent questions and the limits of molecular data. American Journal of Botany 96: 228236.CrossRefGoogle Scholar
Mathiasen, R.L. & Hawksworth, F.G. 1988. Dwarf mistletoes on western white pine in northern California and southern Oregon. Forest Science 34: 429440.CrossRefGoogle Scholar
Matos, J.A. & Schaal, B.A. 2000. Chloroplast evolution in the Pinus montezumae complex: a coalescent approach to hybridisation. Evolution 54: 12181233.Google Scholar
Matsui, T., Wilson, J.B. & Wst, C.J. 2004. Can Podocarpus totara and P. cunninghamii be distinguished by bark thickness? A study on the southern coast of Southland Otago, New Zealand. New Zealand Journal of Botany 42: 313320.CrossRefGoogle Scholar
Matsumoto, M., Ohsawa, T.A., Nishida, M. & Nishida, H. 1997. Glyptostrobus rubensawaensis sp. nov., a new permineralised conifer species from the Middle Miocene, Central Hokkaido, Japan. Paleontological Research 1: 8199.Google Scholar
Matthews, J.A. 1992. The Ecology of Recently Deglaciated Terrain: A Geoecological Approach to Glacier Forelands and Primary Succession. Cambridge: Cambridge University Press.Google Scholar
Mattson, D.J. & Jonkel, C. 1990. Stone Pines and Bears. Washington, DC: USDA.Google Scholar
Mattson, W.J. 1980. Herbivory in relation to plant nitrogen content. Annual Review of Ecology and Systematics 11: 119161.CrossRefGoogle Scholar
Matziris, D.I. 1982. Variation in growth and quality characters in Pinus pinaster provenances grown on seven sites in Greece. Silvae Genetica 31: 168173.Google Scholar
Mayr, E. 1963. Animal Species and Evolution. London: Oxford University Press.CrossRefGoogle Scholar
Mazus, H. 2000. Clues on the history of Podocarpus forest in Maputaland, South Africa, during the Quaternary, based on pollen analysis. African Geosciences Review 7: 7582.Google Scholar
Mazzetta, G.A. 2004. Mesozoic terrestrial giants. Journal of Morphology 260: 311.Google Scholar
Mazzetta, G.A., Christiansen, P. & Farina, R. 2004. Giants and bizarre body size of some southern South American Cretaceous dinosaurs. Historical Biology 16: 7183.CrossRefGoogle Scholar
McDowall, R.M. 2008. Process and pattern in the biogeography of New Zealand: a global microcosm? Journal of Biogeography 35: 197212.CrossRefGoogle Scholar
McEwen, W.M. 1988. Cone and seed phenology in several New Zealand conifer species. Tuatara 30: 6676.Google Scholar
McGee, P.A., Bullock, S. & Summerell, B.A. 1999. Structure of mycorrhizas of the Wollemi pine (Wollemia nobilis) and related Araucariaceae. Australian Journal of Botany 47: 8595.CrossRefGoogle Scholar
McGlone, M.S., Neall, V.E. & Clarkson, B.D. 1988. The effect of recent volcanic events and climatic changes on the vegetation of Mt. Egmont (Mt. Taranaki), New Zealand. New Zealand Journal of Botany 26: 123144.CrossRefGoogle Scholar
McIver, E.E. 1992. Fossil Fokienia (Cupressaceae) from the Paleocene of Alberta, Canada. Canadian Journal of Botany 70: 742749.CrossRefGoogle Scholar
McIver, E.E. & Basinger, J.F. 1993. Flora of the Ravenscrag Formation (Paleocene), south-western Saskatchewan, Canada. Palaeontographic Canada 10: 1167.Google Scholar
McLeod, K.W. & Ciravolo, T.G. 1997. Differential sensitivity of Nyssa aquatica and Taxodium distichum seedlings grown in fly ash amended sand. Wetlands 17: 330335.CrossRefGoogle Scholar
McLoughlin, S. 2001. The breakup of Gondwana and its impact on pre-Cenozoic floristic provincialism. Australian Journal of Botany 49: 271300.CrossRefGoogle Scholar
McLoughlin, S., Drinnnan, A.N. & Rozefelds, A.C. 1995. A Cenomanian flora from the Winton Formation, Eromanga basin, Queensland, Australia. Memoirs of the Queensland Museum 38: 273313.Google Scholar
McLoughlin, S., Carpenteer, R.J., Jordan, G.J. & Hill, R.S. 2008. Seed ferns survived the end-Cretaceous mass extinction in Tasmania. American Journal of Botany 95: 465471.CrossRefGoogle ScholarPubMed
McMillan, C. 1956. The edaphic restriction of Cupressus and Pinus in the Coast Ranges of central California. Ecological Monographs 26: 177212.CrossRefGoogle Scholar
McNab, W.H. 1993. A topographic index to quantify the effect of meso-scale landform on site productivity. Canadian Journal of Forest Research 23: 11001107.CrossRefGoogle Scholar
McQuillan, P.B., Watson, J.E.M., Fitzgerald, N.B., Leaman, D. & Obendorf, D. 2009. The importance of ecological process for terrestrial biodiversity conservation in Tasmania: a review. Pacific Conservation Biology 15: 171196.CrossRefGoogle Scholar
Meacham, N.M. 2002. Fire and the natural history of Giant Sequoias. The American Biology Teacher 64: 573578.CrossRefGoogle Scholar
Medail, F. & Quezel, P. 1997. Hot-spots analysis for conservation of plant diversity in the Mediterranean basin. Annals of the Missouri Botanical Garden 84: 112127.CrossRefGoogle Scholar
Meijer Drees, E. 1940. The genus Agathis in Malaysia. Bulletin Jardin Botanique Buitenzorg 3(16): 455474.Google Scholar
Meinzer, F., Goldstein, G. & Marisol, J. 1984. The effect of atmospheric humidity on stomatal control of gas exchange in two tropical coniferous species. Canadian Journal of Botany 62: 591595.CrossRefGoogle Scholar
Melville, R. 1954. The podocarps of East Africa. Kew Bulletin 1954: 563574.CrossRefGoogle Scholar
Melville, R. 1969. Leaf venation patterns and the origin of the angiosperms. Nature 224: 121125.CrossRefGoogle Scholar
Melville, R. 1975. The distribution of Australian relict plants and its bearing on angiosperm evolution. Botanical Journal of the Linnean Society 70: 6788.CrossRefGoogle Scholar
Mendoza, C.F.V. & Rodríguez-Banderas, A. 2006. Evolutionary analysis of Pinus leiophylla: a study using an intron II sequence fragment of mitochondrial nad1. Canadian Journal of Botany 84: 172177.CrossRefGoogle Scholar
Menendez, C.A. & Caccavari, M.A. 1966. Estructurea epidermica de Araucaria nathorsti Dus: Det Terciaro de Pico Quemado, Rio Negro. Ameghiniana 4: 195199.Google Scholar
Mercer, J.H. 1984. Late Cenozoic glacier variation in South America south of the equator. Pp 4558 in Vogel, J.G. (ed.), Late Cainozoic Palaeoclimates of the Southern Hemisphere. Balkema: Rotterdam.Google Scholar
Merrill, E.D. 1948. A living Metasequoia in China. Science 107: 140.CrossRefGoogle ScholarPubMed
Merrill, E.D. 1948. Metasequoia, another ‘living fossil’. Arnoldia 8: 18.CrossRefGoogle Scholar
Merrill, E.D. & Chaney, R.W. 1949. Metasequoia research from the standpoint of both botany and paleobotany. American Philosophical Society Yearbook 1948: 150151.Google Scholar
Merrill, E.D. 1948. Metasequoia, a living relict of a fossil genus. Journal of the Royal Horticultural Society 73: 211216.Google Scholar
Metsker, J.P. 2000. Tree functional group richness and landscape structure in a Brazilian tropical fragmented landscape. Ecological Applications 10: 11471161.CrossRefGoogle Scholar
Mew, G. & Lee, R. 1981. Investigation of properties and genesis of west Coast wetland soils, South Island, New Zealand. 1. Type localities, profile morphology, and soil chemistry. New Zealand Journal of Science 24: 124.Google Scholar
Meyen, S.V. 1976. Permian conifers of the west Angaraland and new puzzles in the coniferalean phylogeny. Palaeobotanist 25: 298313.Google Scholar
Meyen, S.V. 1977. Cardiolepidaceae: a new coniferalean family from the Upper Permian of north Eurasia. Paleontological Zhurnal 3: 130140 (in Russian).Google Scholar
Meyen, S.V. 1982. The Carboniferous and Permian floras of Angaraland (synthesis). Biological Memoirs 7: 1109.Google Scholar
Meyen, S.V. 1987. Fundamentals of Palaeobotany. London: Chapman & Hall.CrossRefGoogle Scholar
Mezquida, E.T. & Benkman, C.W. 2005. The geographic selection mosaic for squirrels, crossbills and Aleppo pine. Journal of Evolutionary Biology 18: 348357.CrossRefGoogle Scholar
Mickle, J.E. 1993. Cuticular micromorphology of Pagiophyllum bladenensis, comb. nov., from the Late Cretaceous of the North Carolina Coastal Plain, U.S.A. Bulletin of the Torrey Botanical Club 120: 387391.CrossRefGoogle Scholar
Middleton, B.A. 2009. Effects of hurricane Katrina on the forest structure of Taxodium distichum swamps of the Gulf Coast, USA. Wetlands 29: 8087.CrossRefGoogle Scholar
Midgley, J.J. 1989. Pollen dispersal differences for a conifer canopy species in the Knysna Forest. South African Journal of Botany 55: 662663.CrossRefGoogle Scholar
Midgley, J.J., Cameron, M.C. & Bond, W.J. 1995. Gap characteristics and replacement patterns in the Knysa Forest, South Africa. Journal of Vegetation Science 6: 2936.CrossRefGoogle Scholar
Midgley, J.J., Everard, D.A. & van Wyk, G. 1995. Relative lack of regeneration of shade-intolerant species in some South African forests. South African Journal of Science 91: 78.Google Scholar
Midgley, J.J., Midgley, G. & Bond, W.J. 2002. Why were dinosaurs so large? A food quality hypothesis. Evolutionary and Ecological Research 4: 10931095.Google Scholar
Miki, S. 1938. On the change of flora of Japan since the Upper Pliocene and the floral composition of the present. Japanese Journal of Botany 8.Google Scholar
Miki, S. 1954. The occurrence of the remains of Taiwania and Palaeotsuga (n. subgen.) from Pliocene beds in Japan. Proceedings of the Japanese Academy 30(10): 976981.CrossRefGoogle Scholar
Miki, S. & Hikita, S. 1950. The probable chromosome number on the remains of Metasequoia and Sequoia in Japan. Botanical Magazine (Tokyo) 63: 119123.CrossRefGoogle Scholar
Mildenhall, D.C. & Johnston, M.R. 1971. A megastrobilus belonging to the genus Araucarites from the Upper Moutan (Upper Albian), Walparapa, North Island, New Zealand. New Zealand Journal of Botany 9: 6779.CrossRefGoogle Scholar
Mildenhall, D.C. & Pocknall, D.T. 1989. Miocene–Pliocene spores and pollen from Central Otago, South island, New Zealand. New Zealand Geological Survey Paleontological Bulletin 59.Google Scholar
Miles, A.C., Castleberry, S.B., Miller, D.A. & Conner, L.M. 2006. Multi-scale roost site selection by evening bats on pine-dominated landscapes in southwest Georgia. Journal of Wildlife Management 70: 11911199.CrossRefGoogle Scholar
Mill, R.R. & Quinn, C.J. 2001. Prumnopitys andina reinstated as the correct name for Lleuque, the Chilean conifer recently renamed P. spicata (Podocarpaceae). Taxon 50: 11431154.CrossRefGoogle Scholar
Millar, C.I. 1986. Bishop pine (Pinus muricata) of inland Marin County, California. Madrõno 33: 123129.Google Scholar
Millar, C.I. 1986. The Californian closed cone pines (subsection Oocarpae Little and Crichfield): a taxonomic history and review. Taxon 35: 657670.CrossRefGoogle Scholar
Millar, C.I., Westfall, R.D., Delany, D.L., King, J.C. & Graumlich, L.J. 2004. Response of subalpine conifers in the Sierra Nevada, California, U.S.A., to 20th-century warming and decadal climate variability. Arctic, Antarctic and Alpine Research 36: 181200.CrossRefGoogle Scholar
Millay, M.A. & Taylor, T.N. 1976. Evolutionary trends in fossil gymnosperms pollen. Review of Palaeobotany and Palynology 21: 6591.CrossRefGoogle Scholar
Miller, C.N. 1971. Evolution of the fern family Osmundaceae, based on anatomical features. Contributions of the University of Michigan Museum of Paleontology 23: 105169.Google Scholar
Miller, C.N. 1976. Early evolution of the Pinaceae. Review of Palaeobotany and Palynology 21: 101117.CrossRefGoogle Scholar
Miller, C.N. 1977. Mesozoic conifers. Botanical Review 43: 217280.CrossRefGoogle Scholar
Miller, C.N. 1990. Stems and leaves of Cunninghamiostrobus goedertii from the Oligocene of Washington. American Journal of Botany 77: 693971.CrossRefGoogle Scholar
Miller, C.N. & Brown, J.T. 1983. A new volzialean cone bearing seeds with embryos from the Permian of Texas. American Journal of Botany 60: 561569.CrossRefGoogle Scholar
Miller, C.N. & LaPasha, C.A. 1983. Structure and affinities of Athrotaxites berryi Bell, an early Cretaceous conifer. American Journal of Botany 70: 772779.CrossRefGoogle Scholar
Miller, C.N. Jr. 1974. Pityostrobus hallii, a new species of structurally preserved conifer cones from the Late Cretaceous of Maryland. American Journal of Botany 61: 798804.CrossRefGoogle Scholar
Miller, C.N. Jr. 1976. Two new pinaceous cones from the early Cretaceous of California. Journal of Paleontology 50: 821832.Google Scholar
Miller, C.N. Jr. 1977. Pityostrobus lynni (Berry) comb nov., a pinaceous seed cone from the Paleocene of Virginia. Bulletin of the Torrey Botanical Club 104: 59.CrossRefGoogle Scholar
Miller, C.N. Jr. 1978. Pinus burtii, a new species of petrified cones from the Miocene of Martha’s Vineyard. Bulletin of the Torrey Botanical Club 105: 9397.CrossRefGoogle Scholar
Miller, C.N. Jr. 1978. Pityostrobus cliffwoodensis (Berry) comb. Nov., a pinaceous seed cone from the Late Cretaceous of New Jersey. Botanical Gazette 139: 284287.CrossRefGoogle Scholar
Miller, C.N. Jr. & Malinky, J.M. 1986. Seed cones of Pinus from the late Cretaceous of New Jersey, U.S.A. Review of Palaeobotany and Palynology 46: 257272.CrossRefGoogle Scholar
Miller, C.N. Jr. & Robison, C.R. 1975. Two new species of structurally preserved pinaceous cones from the Late Cretaceous of Martha’s Vineyard Island, Massachusetts. Journal of Paleontology 49: 138150.Google Scholar
Ministry of Science, Technology and Environment (MSTE) 1996. Red Data Book of Vietnam, Vol 2 – Plant. Hanoi: Science and Technology Publisher.Google Scholar
Mitchell, A.F. 1994. Champion Trees of the British Isles. London: HMSO.Google Scholar
Mitra, A.K. 1927. On the occurrence of two ovules on araucarian cone scales. Annals of Botany 41: 461471.CrossRefGoogle Scholar
Mitton, J.B. & Latta, R.G. 1997. A comparison of population differentiation across four classes of gene markers in limber pine (Pinus flexilis James). Genetics 146: 11531163.Google Scholar
Mitton, J.B., Kreiser, B.R. & Latta, R.G. 2000. Glacial refugia of limber pine (Pinus flexilis James) inferred from the population structure of mitochondrial DNA. Molecular Ecology 9: 9197.CrossRefGoogle ScholarPubMed
Miyaki, M. 1987. Seeed dispersal of the Korean pine, Pinus koraiensis, by the red squirrel, Sciurus vulgaris. Ecological Research 2: 147157.CrossRefGoogle Scholar
Miyamoto, K., Suzuki, E., Kohyama, T., et al. 2003. Habitat differentiation among tree species with small-scale variation of humus depth and topography in a tropical heath forest of central Kalimantan, Indonesia. Journal of Tropical Ecology 19: 4354.CrossRefGoogle Scholar
Miyazaki, Y. Effect of hiba (Thujopsis dolabrata variety hondae) wood oil on the house dust mite (Dermatophagoides pteronyssinus). Mokuzai Gakkaishi 42: 624626.Google Scholar
Molina, A. 1964. Coniferas de Honduras. Ceiba 10: 521.Google Scholar
Molinari, A., Wauters, L.A., Airoldi, G., et al. 2006. Cone selection by Eurasian red squirrels in mixed conifer forests in the Italian Alps. Acta Oecologica 30: 110.CrossRefGoogle Scholar
Moll, E.J. 1972. The current status of mistbelt mixed Podocarpus forest in Natal. Bothalia 10: 595598.CrossRefGoogle Scholar
Möller, M., Mill, R.R., Glidewell, S.M., Masson, D. & Williamson, B. 2000. Comparative biology of the pollination mechanisms in Acmopyle pancheri and Phyllocladus hypophyllus (Podocarpaceae s.l.) and their taxonomic significance. Annals of Botany n.s. 86: 149158.CrossRefGoogle Scholar
Molloy, B.P.J. & Muñoz-Schick, M. 1999. The correct name for the Chilean conifer Lleuque (Podocarpaceae). New Zealand Journal of Botany 37: 189193.CrossRefGoogle Scholar
Momohara, A. 1994. Floral and paleoenvironmental history from the late Pliocene to Middle Pleistocene in and around central Japan. Palaeogeography, Palaeoclimatology, Palaeoecology 108: 281293.CrossRefGoogle Scholar
Momohara, A., Mizuno, K., Tsuji, S. & Kowaha, S. 1990. Early Pleistocene plant biostratigraphy of the Shobudani Formation, southwest Japan, with reference to extinction of plants. Daiyonki-kenkyu (Quaternary Research): 29: 115.CrossRefGoogle Scholar
Mong, C.E. & Vetaas, O.R. 2006. Establishment of Pinus wallichiana on a Himalayan glacier foreland: stochastic distribution or safe sites? Arctic and Alpine Research 38: 584592.CrossRefGoogle Scholar
Monk, C.D. 1965. Southern mixed hardwood forest of north-central Florida. Ecological Monographs 35: 335354.CrossRefGoogle Scholar
Monk, C.D. 1966. An ecological significance of evergreen-ness. Ecology 47: 504505.CrossRefGoogle Scholar
Monk, C.D. & Brown, T.W. 1965. Ecological consideration of cypress heads in north-central Florida. American Midland Naturalist 74: 126140.CrossRefGoogle Scholar
Moore, D.M. 1983. Flora of Tierra del Fuego. Missouri: Nelson.Google Scholar
Moore, L.B. & Edgar, E. 1970. Flora of New Zealand, Vol. 2. Wellington: Government Printer.Google Scholar
Mopper, S., Mitton, J.B., Whitham, T.G., Cobb, N.S. & Christensen, K.M. 1991. Genetic differentiation and heterozygosity in pinyon pine associated with resistance to herbivory and environmental stress. Evolution 45: 989990.CrossRefGoogle ScholarPubMed
Morat, P. 1993. Our knowledge of the flora of New Caledonia: endemism and diversity in relation to vegetation types and substrate. Biodiversity Letters 1: 7281.CrossRefGoogle Scholar
Morat, P., Veillon, J.-M. & MacKee, H.S. 1986. Floristic relationships of the New Caledonian phanerogams. Telopea 2: 631679.CrossRefGoogle Scholar
Morat, P., Jaffré, T. & Veillon, J.-M. 1999. Menaces sur les taxons rares de la Nouvelle-Calédonie. Actes du Colloque sur les especes vegetaux manacees de France. Bulletin Société Botanique du Sud-ost 19: 129144.Google Scholar
Morgenstern, E.K. & Farrar, J.L. 1964. Introgressive Hybridisation in Red Spruce and Black Spruce. Toronto: University of Toronto Faculty of Science.Google Scholar
Morley, R.J. 2000. Origin and Evolution of Tropical Rainforests. New York: Wiley.Google Scholar
Morrison, T.M. & English, D.A. 1967. The significance of mycorrhizal root nodules of Agathis australis. New Phytologist 66: 245250.CrossRefGoogle Scholar
Morvan, J. 1990. Comparative development, structure and phylogeny of the female strobile of Saxegothaea compared with Microcachrys tetragona Hook (Podocarpaceae). Compte Rendu de l’Académie des Sciences Paris, sér. 3, Sciences 310: 651656.Google Scholar
Moslandl, R. & Kleinert, A. 1998. Development of oaks (Quercus petraea (Matt.) Liebl.) emerged from bird-dispersed seeds under old-growth pine (Pinus sylvestris L.) stands. Forest Ecology and Management 106: 3544.CrossRefGoogle Scholar
Mosle, B., Collinson, M.E., Finch, P., et al. 1998. Factors influencing the preservation of plant cuticles: a comparison of morphology and chemical composition of modern and fossil examples. Organic Geochemistry 29: 13691380.CrossRefGoogle Scholar
Mosseler, A., Eggler, K.N. & Hughes, G.A. 1992. Low levels of genetic diversity in red pine confirmed by random amplified polymorphic DNA markers. Canadian Journal of Forest Research 22: 13321337.CrossRefGoogle Scholar
Motzkin, G., Patterson, W.A. & Drake, N.E.R. 1993. Fire history and vegetation dynamics of Chamaecyparis thyoides wetland on Cape Cod, Massachusetts. Journal of Ecology 81: 391402.CrossRefGoogle Scholar
Mueller-Dombois, D. & Fosber, F.R. 1998. Vegetation of the Tropical Pacific Islands. New York: Springer.CrossRefGoogle Scholar
Muir, M.D. & Van Konijnenburgh-Van Citert, J.H.A. 1970. A Rhaeto-Liassic flora from Airel, northern France. Palaeontology 13: 433442.Google Scholar
Muir, P.S. & Lotan, J.E. 1985. Serotiny and life history of Pinus contorta var. latifolia. Canadian Journal of Botany 63: 938945.CrossRefGoogle Scholar
Muir, P.S. & Lotan, J.E. 1985. Disturbance history and serotiny of Pinus contorta in western Montana. Ecology 66: 16581668.CrossRefGoogle Scholar
Muller, J. 1984. Significance of fossil pollen for angiosperm history. Annals of the Missouri Botanical Garden 71: 419443.CrossRefGoogle Scholar
Murphy, T.M. 1981. Immunochemical comparisons of seed proteins from populations of Pinus radiata (Pinaceae). American Journal of Botany 68: 254259.CrossRefGoogle Scholar
Murray, B.G. & Davies, B.J. 1996. An improved method of preparing the chromosomes of pines and other gymnosperms. Biotechnic & Histochemistry 71: 115117.CrossRefGoogle ScholarPubMed
Musoko, M., Last, F.T. & Mason, P.A. 1994. Populations of spores of vesicular-arbuscular mycorrhizal fungi in undisturbed soils of secondary deciduous moist tropical forests in Cameroon. Forest Ecology and Management 63: 359377.CrossRefGoogle Scholar
Mustoe, G.E. 2002. Eocene Ginkgo leaf fossils from the Pacific Northwest. Canadian Journal of Botany 80: 10781087.CrossRefGoogle Scholar
Myers, T.S. & Fiorilo, A.R. 2009. Evidence for gregarious behaviour and age segregation in sauropod dinosaurs. Palaeogeography, Palaeoclimatology, Palaeoecology 274: 96104.CrossRefGoogle Scholar
Nadkarni, N. & Wheelwright, N.T. (eds.). 2000. Monteverde: Ecology and Conservation of a Tropical Cloud Forest. Oxford: Oxford University Press.CrossRefGoogle Scholar
Nagakura, J., Kaneko, S., Takahashii, M & Tange, T. 2008. Nitrogen promotes water consumption in seedings of Cryptomeria japonica but not in Chamaecyparis obtusa. Forest Ecology and Management 255: 25332541.CrossRefGoogle Scholar
Nagalingum, N.S., Drinnan, A.N., Upia, R. & McLoughlin, S. 2002. Fern spore diversity and abundance in Australia during the Cretaceous. Review of Palaeobotany and Palynology 119: 6992.CrossRefGoogle Scholar
Nakagawa, T., Yasuda, Y. & Tabata, H. 1996. Pollen morphology of Himalayan Pinus and Quercus and its importance in palynological studies in Himalayan area. Review of Palaeobotany and Palynology 91: 317329.CrossRefGoogle Scholar
Nakai, T. 1938. Indigenous species of conifers and taxads of Korea and Manchuria and their distribution. Botanical Magazine Tokyo 158: 129.Google Scholar
Nakamura, Y. & Kano, T. 1943. Cytological and genetical studies on forest plants. II. Chromosome conjugation in pollen mother cells and pollen fertility of the intermediate black pine (Pinus thunbergii Uyeki). Seiken Ziho 2: 9096 (in Japanese).Google Scholar
Nakano, T. 1985. Ecologica studies of Abies and Tsuga forests in Kyishu, Japan. Bulletin Miyazaki University Forest 11: 1165 (in Japanese with English summary).Google Scholar
Nakashizuka, T., Iida, S., Suzuki, W. & Tanimoto, T. 1993. Seed dispersal and vegetation development on a debris avalanche on the Ontake volcano, Central Japan. Journal of Vegetation Science 4: 537542.CrossRefGoogle Scholar
Namboodiri, K.K. & Beck, C.B. 1968. A comparative study of the primary vascular system of conifers. 1. Genera with helical phyllotaxis. American Journal of Botany 55: 447457.CrossRefGoogle Scholar
Nanami, S., Kawaguci, H. & Yamakura, T. 1999. Doecy-induced spatial patterns of two codominant tree species, Podocarpus nagi and Neolitsea aciculate. Journal of Ecology 87: 678687.CrossRefGoogle Scholar
Nanami, S., Kawaguchi, H. & Yamakura, T. 2005. Sex ratio and gender-dependent neighboring effects in Podocarpus nagi, a dioecious tree. Plant Ecology 177: 209222.CrossRefGoogle Scholar
Nanko, K., Hotta, N. & Suzuki, M. 2006. Evaluating the influence of canopy species and meteorological factors on throughfall drop size distribution. Journal of Hydrology 329: 422431.CrossRefGoogle Scholar
Napp-Zinn, K. 1996. Anatomie des Blattes. I. Gymnospermen. Handbuch der Pflanzenanatomie 8: 1369.Google Scholar
Naranjo, C.A., Poggio, L. & Brandham, P.E. 1983. A practical method of chromosome classification on the basis of centromere position. Genetics 62: 5153.Google Scholar
Nasi, R., Jaffré, T. & Sarrailh, J.M. 2002. Les forêts de montagnes de la Nouvelle-Caledonia. Bois et Forets Tropiques 274: 517.Google Scholar
Nasir, E., Siddiqi, M.A. & Ali, Z. 1969. Gymnosperms of West Pakistan. Rawlpindi: Ferozsons Press.Google Scholar
Natarajan, S., Murti, V.V.S. & Seshadri, T.R. 1970. Biflavones of some Cupressaceae plants. Phytochemistry 9: 575579.CrossRefGoogle Scholar
Nathan, R. & Ne’eman, G. 2004. Spatiotemporal dynamics of recruitment in Aleppo pine (Pinus halepensis Miller). Plant Ecology 171: 123137.CrossRefGoogle Scholar
Nathan, R., Safriel, U.N., Noy-Meir, I. & Schiller, G. 1999. Seed release without fire in Pinus halepensis, a Mediterranean serotinous wind-dispersed tree. Journal of Ecology 87: 659669.CrossRefGoogle Scholar
Nathan, R., Safrielm, U.N. & Noy-Meir, I. 2001. Field evaluation and sensitivity analysis of a mechanistic model for tree seed dispersal by wind. Ecology 82: 374388.CrossRefGoogle Scholar
Ne’eman, G. & Izhaki, I. 1998. Stability of pre- and post-fire spatial structure of pine trees in Aleppo pine forest. Ecography 21: 535–524.Google Scholar
Ne’eman, G., Fotheringham, C.J. & Keeley, J.E. 1999. Patch to landscape patterns in post-fire recruitment of a serotinous conifer. Plant Ecology 145: 235242.CrossRefGoogle Scholar
Ne’eman, G., Goubitz, S. & Nathan, R. 2004. Reproductive traits of Pinus halepensis in the light of fire: a critical review. Plant Ecology 171: 6979.CrossRefGoogle Scholar
Nelson, E.C. 1974. Disjunct plant distributions on the southwestern Nullabor Plain, Western Australia. Journal of the Royal Society of Western Australia 57: 105116.Google Scholar
Newberry, D.M., Renshaw, E. & Brunig, E.F. 1986. Spatial pattern of trees in kerangas forest, Sarawak. Vegetatio 65: 7789.CrossRefGoogle Scholar
Newcomer, E.H. 1945. Induced parthenocarpy in Ginkgo. American Naturalist, 79: 186187.CrossRefGoogle Scholar
Newcomer, E.H. 1954. The karyotype and possible sex chromosomes of Ginkgo biloba. American Journal of Botany 41: 542545.CrossRefGoogle Scholar
Newman, E.I. 1973. Competition and diversity in herbaceaous vegetation. Nature 244: 310.CrossRefGoogle Scholar
Newman, E.I. 1995. Phosphorous inputs to terrestrial ecosystems. Journal of Ecology 83: 713726.CrossRefGoogle Scholar
Newnham, R. 1999. Environmental change in Northland, New Zealand, during the last glacial and Holocene. Quaternary International 57: 6170.CrossRefGoogle Scholar
Newsome, J. & Flenley, J.R. 1988. Late Quaternary vegetational history of the Central Highlands of Sumatra. II. Palaeo-palynology and vegetational history. Journal of Biogeography 15: 555578.CrossRefGoogle Scholar
Newton, A.C., Allnutt, T.R., Gillies, A.C.M., Lowe, A.J. & Ennos, R.A. 1999. Molecular phylogeography, intraspecific variation and the conservation of tree species. Trends in Ecology and Evolution 14: 140145.CrossRefGoogle ScholarPubMed
Newton, A.C., Allnutt, T.R., Dvorak, W.S., Del-Castillo, R.F. & Ennos, R.A. 2002. Patterns of genetic variation in Pinus chiapensis, a threatened Mexican pine, detected by RAPD and mitochondrial DNA RFLP markers. Heredity 89: 191198.CrossRefGoogle ScholarPubMed
Nghia, N.H. 2000. Some Threatened Tree Species of Vietnam. Hanoi: Agriculture Publisher.Google Scholar
Nguyễn, T.H. & Vidal, J.E. 1996. Flora du Camboge, du Laos et du Viêtnam. Paris: Museum National d’Histoire Naturelle.Google Scholar
Nguyễn, T.H., Do, T.D. & Phan, K.L. 2002. The diversity of the flora of Vietnam. 9. Taiwania Hayata and T. cryptomerioides Hayata (Taxodiaceae): new genus and species for the flora. Journal of Genetics and Applications 1: 3240 (in Vietnamese with English summary).Google Scholar
Nguyễn, T.H., Loc, P.K., Nguyễn, D.T.L., et al. 2004. Vietnam Conifers Conservation Status Review 2004. Fauna & Flora International, Vietnam Programme. Hanoi: Labour and Society Publisher.Google Scholar
Nicholson, R. & García-Biao, B. 1999. Observations on the propagation of Cupressus dupreziana, an endemic conifer of the Sahara Desert. Botanic Garden Conservation News 3: 4950.Google Scholar
Niemann, G.J. & Van Genderen, H.H. 1980. Chemical relationships between Pinaceae. Biochemical Systematics and Ecology 2: 237240.CrossRefGoogle Scholar
Nikitenko, B.L., Shurygin, B.N., Knyazev, V.G., et al. 2013. Jurassic and Cretaceous stratigraphy of the Anabar area (Arctic Siberia), Laptev Sea Coast and the Boreal zonal standard. Russian Geology and Geophysics 54: 808837.CrossRefGoogle Scholar
Niklas, K.J. 1981. Simulated wind pollination and airflow around ovules on some early seed plants. Science 211: 275277.CrossRefGoogle ScholarPubMed
Niklas, K.J. 1981. Airflow patterns around some early seed plant ovules and cupules: implications concerning efficiency in wind pollination. American Journal of Botany 68: 635650.CrossRefGoogle Scholar
Niklas, K.J. 1983. The influence of paleozoic ovule and cupule morphologies of wind pollination. Evolution 37: 968986.CrossRefGoogle ScholarPubMed
Niklas, K.J. 1983. Early seed plant wind pollination studies: a reply. Taxon 32: 99100.CrossRefGoogle Scholar
Niklas, K.J. 1985. The aerodynamics of wind pollination. Botanical Review 51: 328386.CrossRefGoogle Scholar
Niklas, K.J. & Norstog, K. 1984. Aerodynamics and pollen grain depositional patterns on cycad megastrobili: implications of the reproduction of three cycad genera (Cycas, Dioon and Zamia). Botanical Gazette 145: 92104.CrossRefGoogle Scholar
Nilsson, T. 1983. The Pleistocene: Geology and Life in the Quaternary Age. Stuttgart: Ferdinand Enke.Google Scholar
Ninemets, Ü., Sparrow, A. & Cescatti, A. 2005. Light capture efficiency decreases with increasing tree age and size in the southern hemisphere gymnosperm Agathis australis. Trees 19: 177190.CrossRefGoogle Scholar
Ning, S.J., Zhao, T.L., Tang, R.Q., et al. 1997. Preliminary studies on the phytocoenological features of the Calocedrus macrolepis community in Mulun, Huanjiang county, Guangxi. Guihaia 17: 321330.Google Scholar
Nishida, H. 1991. Diversity and significance of Late Cretaceous permineralized plant remains from Hokkaido, Japan. Botanical Magazine Tokyo 104: 253273.CrossRefGoogle Scholar
Nishida, M. & Nishida, H. 1986. Structure and affinities of the petrified plants from the Cretaceous of northern Japan and Saghalien. III. Petrified plants from the Upper Cretaceous of Saghalien. Botanical Magazine Tokyo 99: 191204.CrossRefGoogle Scholar
Nishida, M., Ohsawa, T. & Nishida, H. 1992. Structure and affinities of the petrified plants from the Cretaceous of Northern Japan and Saghalien VIII. Gen et sp. nov., a taxodiaceous cone from the Upper Cretaceous of Hokkaido. Journal of Japanese Botany 67: 19.Google Scholar
Nix, H.A. & Kalma, J.D. 1972. Climate as a dominant control in the biogeography of northern Australia and New Guinea. Pp 6191 in Walker, D. (ed.), Bridge and Barrier: The Natural and Cultural History of Torres Strait. Canberra: Australian National University.Google Scholar
Nixon, K.C., Crepet, W.L., Friiis, E.M. & Stevenson, D.W. 1994. A reevaluation of phylogenetic relationships of seed plants. Annals of the Missouri Botanical Garden 81: 484533.CrossRefGoogle Scholar
Nocleberg, W.J., Parfenov, L.M., Monger, J.W.H., et al. 1998. Phanerozoic tectonic evolution of the circum-North Pacific. US Geological Survey Open File Report 98-754.CrossRefGoogle Scholar
Norton, D.A., Herbert, J.W. & Beveridge, A.E. 1988. The ecology of Dacrydium cupressinum: a review. New Zealand Journal of Botany 26: 3762.CrossRefGoogle Scholar
Nosetto, M.D., Jobbagy, E.G. & Paruelo, J.M. 2006. Carbon sequestration in semi-arid rangelands: comparison of Pinus ponderosa plantations and grazing exclusion in NW Patagonia. Journal of Arid Environments 67: 142156.CrossRefGoogle Scholar
Nosova, N.V. 2014. Distribution of Miroviaceae (Mesozoic conifers) in Russia. In 9th European Palaeobotany-Palynology Conference, Padova, Italy.Google Scholar
Novak, F. 1927. Zur funfzigjahrenrigen Entdeckung der Picea omorika. Mitt. Deutsche Dendrol. Gesellschaft 38: 4756.Google Scholar
Numata, M., Miyakaki, A. & Itow, S. Natural and semi-natural vegetation in Japan. Blumea 20: 435496.Google Scholar
Nyland, R.D. 1998. Patterns of lodgepole pine regeneration following the 1988 Yellowstone fires. Forest Ecology and Management 111: 2333.CrossRefGoogle Scholar
O’Brien, I.E.O., Smith, D.R., Gardner, R.C. & Murray, J.C. 1996. Flow cytometric determination of genome size in Pinus. Plant Science 115: 9199.CrossRefGoogle Scholar
O’Gorman, E.J. & Hone, D.W.E. 2012. Body size distribution of the dinosaurs. PLoS One 7: e51925.CrossRefGoogle Scholar
Odum, H.T. & Pigeon, R.F. (eds.). 1970. A Tropical Rainforest: A Study of Irradiation and Ecology at El Verde, Puerto Rico. Washington, DC: US Atomic Energy Commission, Division of Technical Information.Google Scholar
Offler, C.E. 1984. Extant and fossil Coniferales of Australia and New Guinea. Part 1: a study of the external morphology of the vegetative shoots of the extant species. Palaeontographica B 193: 18120.Google Scholar
Offord, C.A., Porter, C.L., Meagher, P.F. & Errington, G. 1999. Sexual reproduction and early plant growth of the Wollemi pine (Wollemia nobilis), a rare and threatened Australian conifer. Annals of Botany 84: 19.CrossRefGoogle Scholar
Ogaya, R., Llorens, L. & Penuelas, J. 2011. Density and length of stomatal and epidermal cells in ‘living fossil’ trees grown under elevated CO2 and polar light regime. Acta Oecologia 37: 381385.CrossRefGoogle Scholar
Ogden, J. 1978. On the dendrochronological potential of Australian trees. Australian Journal of Ecology 3: 339356.CrossRefGoogle Scholar
Ogden, J., Wardle, G.M. & Ah, M. 1987. Population dynamics of the emergent conifer Agathis australis (D.Don) Lindl. (kauri) in New Zealand. II. Seedling population sizes and gap-phase regeneration. New Zealand Journal of Botany 25: 231242.CrossRefGoogle Scholar
Ogilvie, R.T. 1972. Speciation on the North American spruces and its relations to white spruce. Pp 17 in Mcminn, R.G. (ed.), White Spruce: The Ecology of a Northern Resource. Edmonton: Canadian Forest Service.Google Scholar
Ogilvie, R.T. & Rudloff, E. 1968. Chemosystematic studies on the genus Picea (Pinaceae) IV. The introgression of white and Engelmann spruce found along the Bow River. Canadian Journal of Botany 46: 901908.CrossRefGoogle Scholar
Ogura, Y. 1930. On the structure and affinities of some Cretaceous plants from Hokkaido. Journal of the Faculty of Science Tokyo Unversity, Sect. 3 Botany 2: 381412.Google Scholar
Ohana, T. & Kimura, T. 1995. Further observations of Cunninghamiostrobus yubariensis Stopes and Fuji from the Upper Yezo Group (Upper Cretaceous), Hokkaido, Japan. Transactions and Proceedings of the Palaeontological Society of Japan n.s. 178: 122141.Google Scholar
Ohsawa, T., Nihida, M. & Nishida, H. 1991. Structure and affinities of the petrified plants from the Cretaceous of Northern Japan and Saghalien IX. A petrified cone of Sciadopitys from the Upper Cretaceous of Hokkaido. Journal of Phytogeography and Taxonomy 39: 97105.Google Scholar
Oksbjerg, E. 1953. Om Picea omorika. Skovforester Tidsskrift 2: 179192.Google Scholar
Oladele, F.A. 1983. Inner surface sculpture patterns of cuticles in Cupressaceae. Canadian Journal of Botany 61: 12221231.CrossRefGoogle Scholar
Oladele, F.A. 1983. Scanning electron microscope study of stomatal-complex configuration in Cupressaceae. Canadian Journal of Botany 61: 12321240.CrossRefGoogle Scholar
Oldfield, S., Lusty, C. & MacKinven, A. 1998. The World List of Threatened Trees. Cambridge: World Conservation Press.Google Scholar
Oldham, T.C.B. 1976. Flora of the Wealden plant-debris beds in England. Palaeontology 19: 437502.Google Scholar
Oliveira, J.M., Roig, F.A. & Pillar, V.D. 2012. Climate signals in tree rings of Araucaria angustifolia in the southern Brazilian highlands. Austal Ecology 35: 134147.CrossRefGoogle Scholar
Oliver, F.W. 1903. The ovules of the older gymnosperms. Annals of Botany 17: 451476.CrossRefGoogle Scholar
Oostermeijer, J.G.B., Luijten, S.H. & den Niijs, J.C.M. 2003. Integrating demographic and genetic approaches in plant conservation. Biological Conservation 113: 389398.CrossRefGoogle Scholar
Oren, R. 2001. Soil fertility limits carbon sequestration by forest ecosystems in a CO2-enriched atmosphere. Nature: 411: 469472.CrossRefGoogle Scholar
Orman, H.R. & Reid, J.S. 1946. Wood anatomy of New Zealand Dacrydium species. New Zealand Journal of Forestry 5.Google Scholar
Orr, M.Y. 1933. Taiwania in Upper Burma, a new record. Notes from the Royal Botanical Gardens, Edinburgh 18: 6.Google Scholar
Osborne, C.P. & Beerling, D.J. 2002. A process-based model of conifer forest structure and function with special emphasis on leaf lifespans. Global Biogeochemical Cycles 16(4): 44-1–44-23.CrossRefGoogle Scholar
Osborne, C.P. & Beerling, D.J. 2003. The penalty of a long, hot summer: photosynthetic acclimation to high CO2 and continuous light in ‘living fossil’ conifers. Plant Physiology 133: 803812.CrossRefGoogle ScholarPubMed
Osborne, T.G.B. 1960. Some observations on the life history of Podocarpus falcatus. Australian Journal of Botany 8: 243255.CrossRefGoogle Scholar
Ota, H. 1998. Geographic patterns of endemism and speciation in amphibians and reptiles of the Ryukyu archipelago, Japan, with special reference to their paleogeographic implications. Research of Population Ecology 40: 189204.CrossRefGoogle Scholar
Otto, A., Simmoneit, B.R. & Wilde, V. 2007. Terpenoids as chemosystematic markers in selected fossil and extant species of pine (Pinus, Pinaceae). Botanical Journal of the Linnean Society 154: 129140.CrossRefGoogle Scholar
Otto-Bliesner, B.L. & Upchurch, G.R. Jr. 1997. Vegetation-induced warming of high latitude regions during the Late Cretaceous period. Nature 385: 804807.CrossRefGoogle Scholar
Owens, J.N. & Bladke, M.D. 1983. Pollen morphology and development of the pollination mechanism in Tsuga heterophylla and T. mertensiana. Canadian Journal of Botany 61: 30413048.CrossRefGoogle Scholar
Owens, J.N. & Molder, M. 1974. Cone initiation and development before dormancy in yellow cedar (Chamaecyparis nootkatensis). Canadian Journal of Botany 55: 20752084.CrossRefGoogle Scholar
Owens, J.N. & Molder, M. 1975. Pollination, female gametophyte and embryo and seed development in yellow cedar (Chamaecyparis nootkatensis). Canadian Journal of Botany 53: 186199.CrossRefGoogle Scholar
Owens, J.N. & Molder, M. 1980. Sexual reproduction in western red cedar (Thuja plicata). Canadian Journal of Botany 58: 13761393.CrossRefGoogle Scholar
Owino, F. 1977. Genetic divergence in selected populations of loblolly pine. Silvae Genetica 26: 6466.Google Scholar
Page, C.N. 1973. Two hybrids in Equisetum new to the British flora. Watsonia 9: 229237.Google Scholar
Page, C.N. 1979. The herbarium preservation of conifer specimens. Taxon 28: 375379.CrossRefGoogle Scholar
Page, C.N. 1979. The earliest find of living Taiwania (Taxodiaceae). Kew Bulletin 34: 527528.CrossRefGoogle Scholar
Page, C.N. 1985. Pteridophyte biology: the biology of the amphibians of the plant world. Proceedings of the Royal Society of Edinburgh 86B: 439442.Google Scholar
Page, C.N. 1989. Compression and slingshot megaspore ejection in Selaginella selagionoides: a new phenomenon in pteridophytes. Fern Gazette 13: 267275.Google Scholar
Page, C.N. 1990. Taxonomic evaluation of the fern genus Pteridium and its active evolutionary state. Pp 2334 in Thomson, J.A. & Smith, R.T. (eds.), Bracken Biology and Management. Canberra: Australian Institute of Agricultural Science.Google Scholar
Page, C.N. 1990. Hybrids in the genus Equisetum in Europe; an updated annotation. Pp 181186 in Rita, J. (ed.), Taxonomía, Biogeografía y Conservación de Pteridófitos. IME: Palma de Majorca.Google Scholar
Page, C.N. & Bronsey, P.J. 1986. Tree-fern skirts, a defence against climbers and large epiphytes. Journal of Ecology 74: 787796.CrossRefGoogle Scholar
Page, C.N. & McHaffie, H.S. 1991. Pteridophytes as indicators of landscape change in the British Isles in the last hundred years. Pp 2540 in Camus, J.M. (ed.), The History of British Pteridology. London: British Pteridological Society.Google Scholar
Page, V.M. 1973. A new conifer from the Upper Cretaceous of central California. American Journal of Botany 60: 570575.CrossRefGoogle Scholar
Palgrave, K.C. 1957. Trees of Central Africa. Glasgow: Glasgow University Press.Google Scholar
Palma, A.C., Winter, K., Aranda, J., et al. 2020. Why are all tropical conifers disadvantaged in fertile soils? Comparison of Podocarpus guatemalensis with pioneer Ficus insipida. Tree Physiology 40: 810821.CrossRefGoogle ScholarPubMed
Palmarev, E., Petkova, A. & Uzunova, K. 1978. Beitrage zur Entwicklungsgeschchter der Gattung Taiwania Hay. und Cunninghamia R.Br. in Holarktis. Fitologia 9: 316.Google Scholar
Palmer, E. & Pitman, N. Trees of Souhtern Africa. Cape Town: Balkema.Google Scholar
Pande, R.K. 1991. Hydrological properties of oak and pine forests in central Himalaya, India. Indonesian Journal of Geography 21: 4554.Google Scholar
Pant, D.D. 1982. The Lower Gondwana gymnosperms and their relationships. Review of Palaeobotany and Palynology 37: 5570.CrossRefGoogle Scholar
Pant, D.D. & Nautiyal, D.D. 1967. On the structure of Buriadia heterophylla (Feistmantel) Seward & Sanhi and its fructifications. Philosophical Transactions of the Royal Society of London B 252: 2748.Google Scholar
Panti, C., Cersari, S.N., Marenssi, S.A. & Olivero, E.B. 2007. A new araucarian fossil species from the Paleogene of southern Argentina. Ameghiniana 44: 215222.Google Scholar
Parc, G., Canaguier, A., Lamdre, P., et al. 2002. Production of taxoids with biological activity by plants and callus culture from selected Taxus genotypes. Phytochemistry 59: 725730.CrossRefGoogle ScholarPubMed
Parchmann, T.L. & Benkman, C.W. 2002. Diversifying coevolution between crossbills and black spruce on Newfoundland. Evolution 56: 16631672.Google Scholar
Parchmann, T.L., Benkman, C.W. & Mezquida, E.T. 2007. Coevolution between Hispaniolan crossbills and pine: does more time allow for greater phenotypic escalation at lower latitude? Evolution 61: 21422153.CrossRefGoogle Scholar
Parchmann, T.L., Benkman, C.W., Jenkins, B. & Buerkle, C.A. 2011. Low levels of population genetic structure in Pinus contorta (Panaceae) across a geographic mosaic of co-evolution. American Journal of Botany 98: 669679.CrossRefGoogle Scholar
Parker, A.J. 1982. The topographical relative moisture index: an approach to soil moisture assessment in mountain terrain. Physical Geography 3: 160168.CrossRefGoogle Scholar
Parker, A.J. 1986. Persistence of lodgepole pine forests in the central Sierra Nevada. Ecology 67: 15601567.CrossRefGoogle Scholar
Parker, A.J. 1994. Latitudinal gradients of coniferous tree species, vegetation, and climate in Sierra-Cascade axis of northern California. Vegetatio 115: 145155.CrossRefGoogle Scholar
Parker, A.J. 1995. Comparative gradient structure and forest cover types in Lassen Volcanic and Yosemite National parks, California. Bulletin of the Torrey Botanical Club 122: 5868.CrossRefGoogle Scholar
Parker, A.J., Parker, K.C. & Wiggins-Brown, H. 2000. Disturbance and scale-effects on southern old-growth forests (USA): the sand pine example. Natural Areas Journal 20: 273279.Google Scholar
Parker, A.J., Parker, K.C. & McCay, D.H. 2002. Geographic and ecological variation in seedling growth rates of sand pine (Pinus clausa). Southeastern Geographer 42: 2028.CrossRefGoogle Scholar
Parris, B.S., Beaman, R.S. & Beaman, J.H. 1992. The Plants of Mount Kinabalu. London: Royal Botanic Gardens, Kew.Google Scholar
Parrish, J.M. 2006. The origins of high-browsing and the effect of phylogeny and scaling on neck length. Pp 221224 in Amniote Paleobiology. Berkley, CA: University of Caifornia Press.Google Scholar
Parrish, J.M., Peterson, F. & Turner, C.E. 2004. Jurassic ‘savannah’: plant taphonomy and climate of the Morrison Formation (Jurassic, western USA). Sedimentary Geology 167: 139164.CrossRefGoogle Scholar
Parrish, J.T. 1987. Global palaeogeography and palaeoclimate of the Late Cretaceous and Early Tertiary. Pp 5172 in Friis, E.M., Chaloner, W.G. & Crane, P.R. (eds.), The Origins of Angiosperms and their Biological Consequences. Cambridge: Cambridge University Press.Google Scholar
Parrish, J.T. & Spicer, R.A. 1988. Cretaceous (Nanushuk Group, Albian–Cenomanian) wetland environments of the North Slope, Alaska. Abstracts, Geological Society of America 30: 366.Google Scholar
Parsapajouh, D., Braker, O.U., Habibi, H. & Schar, E. 1986. Etude dendroclimatologique du bois de taxus baccata di norde de l’Iran. Schweizerische Zeitschrift fur Forstwesen 137: 853868.Google Scholar
Parsons, D.J. & Benedetti, S.H. 1979. Impact of fire suppression on a mixed-conifer forest. Forest Ecology and Management 2: 2133.CrossRefGoogle Scholar
Pasquini, S.C. & Santiago, L.S. 2012. Nutrient limited photosynthesis in seedlings of a lowland forest tree species. Oecologia 168: 311319.CrossRefGoogle ScholarPubMed
Passini, M.F. 1985. Structure et régénération des formationes ligneuses à Pinus maximartinezii Rzed., Mexique. Bulletin Société Botanique de France, Lettre Botaniques 132: 327339.CrossRefGoogle Scholar
Pate, J.S. 2001. Haustoria in action: case studies of nitrogen acquisition by woody xylem-tapping hemiparasites from their hosts. Protoplasma 215: 204217.CrossRefGoogle ScholarPubMed
Patton, R.F. 1966. Interspecific hybridisation in breeding for white pine blister rust resistance. Pp 367376 in Gerhold, H.D. (ed.), Breeding Pest-Resistant Trees. Oxford: Pergamon Press.CrossRefGoogle Scholar
Paul, G.S. 1988. Physiological, migratorial, climatological, geophysical, survival and evolutionary implications of Cretaceous Polar dinosaurs. Journal of Paleontology 62: 640652.Google Scholar
Pauluch, J.G. & Stępniewska, H.J. 2012. Effects of microsite on the survival, density, and ectomycorrhizal status of shade-tolerant Abies alba regeneration attacked by fungal pathogens. Canadian Journal of Forest Research 42: 720732.CrossRefGoogle Scholar
Pearce, A.J. & O’Loughlin, C.L. 1985. Landsliding during a M 7.7 earthquake: influence of geology and topography. Geology 13: 855858.2.0.CO;2>CrossRefGoogle Scholar
Pearcy, R.W., Ehleringer, J., Money, H.A. & Rundel, P.W. (eds.). 1989. Plant Physiological Ecology. London: Chapman & Hall.CrossRefGoogle Scholar
Pearson, S.G. & Searson, M.J. 2002. High-resolution data from Australian trees. Australian Journal of Botany 50: 431439.CrossRefGoogle Scholar
Pearson, T.R.H., Burslem, D.F.R.P., Goeriz, R.E. & Dalling, J.W. 2003. Regeneration niche-partitioning in neotropical pioneers: effects of gap-size, seasonal drought and herbivory on growth and survival. Oecologia 137: 456465.CrossRefGoogle ScholarPubMed
Peay, K.G., Bruns, T.D. & Garbelotto, M. 2010. Testing the ecological stability of ectomycorrhizal symbiosis: effects of heat, ash and mycorrhizal colonisation on Pinus muricata seedling performance. Plant and Soil 330: 291302.CrossRefGoogle Scholar
Peirce, A.S. 1937. Systematic anatomy of the woods of the Cupressaceae. Tropical Woods 49: 521.Google Scholar
Pérez, C.A., Hedin, L.O. & Armesto, J.J. 1998. Nitrogen mineralisation in two unpolluted old-growth forests of contrasting biodiversity and dynamics. Ecosystems 1: 361373.Google Scholar
Pérez, C.A., Armesto, J.J., Torralba, C. & Carmona, M.R. 2003. Litterfall dynamics and nitrogen use efficiency in two evergreen temperate rainforests of southern Chile. Austral Ecology 28: 591600.CrossRefGoogle Scholar
Pérez de la Rosa, J.A. 1985. Una nueva species de Juniperus de Mexico. Phytologia 57: 8186.Google Scholar
Pérez de le Rosa, J.A. 1999. Regional action plan: conifers threatened in Nuva Galicia, Mexico. Pp 5558 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Pérez Rodríguez, R., Marques, M.J. & Bienes, R. 2007. Use of dendrochronological method in Pinus halepensis to estimate the soil erosion in the south east of Madrid (Spain). Science of the Total Environment 378: 156160.CrossRefGoogle Scholar
Perkins, D.L. & Swetnam, T.W. 1996. A dendrochronological assessment of Whitebark pine in Sawtooth-Salmon region, Idaho. Canadian Journal of Forest Research 26: 21232133.CrossRefGoogle Scholar
Perrin, P.M., Kelly, D.M. & Mitchesll, F.J.G. 2006. Long-term deer exclusion in yew-wood and oakwood habitats in southwest Ireland: natural regeneration and stand dynamics. Forest Ecology and Management 236: 356367.CrossRefGoogle Scholar
Perry, G.L.W. & Enright, N.J. 2002. Humans, fire and landscape pattern: understanding a maquis-forest complex, Mont Do, New Caledonia, using a ‘state-and-transition’ model. Journal of Biogeography 29: 11431158.CrossRefGoogle Scholar
Perry, L. & Brownsey, P. 2007. Molecular evidence for long-distance dispersal in the New Zealand pteridophyte flora. Journal of Biogeography 34: 20282038.CrossRefGoogle Scholar
Peters, M.D. & Christophel, D.C. 1978. Austrosequoia wintonenesis, a new taxodiaceaous cone from Queensland, Australia. Canadian Journal of Botany 56: 31193128.CrossRefGoogle Scholar
Petersen, P.J. 1982. Mineral nutrition of Agathis australis Salisb., the kauri. Part I. Effects of deficiencies of essential elements on the growth and foliar mineral composition of seedlings. New Zealand Journal of Science 5: 141164.Google Scholar
Peterson, C.J., Carson, W.P., McCarthy, B.C. & Pickett, S.T.A. 1990. Microsite variation and soil dynamics within newly created treefall pits and mounds. Oikos 58: 3946.CrossRefGoogle Scholar
Peterson, J.A. 1968. Cirque morphology and Pleistocene ice formation conditions in southeastern Australia. Australian Geographical Studies 6: 6776.CrossRefGoogle Scholar
Pettitt, J.M. 1970. Heterospory and the origin of the seed habit. Biological Reviews, Cambridge University 45: 401415.CrossRefGoogle Scholar
Pettitt, J.M. & Chaloner, W.G. 1964. The ultrastructure of the Mesozoic pollen Classopollis. Pollen et Spores 6: 611620.Google Scholar
Philippe, M., Barbacka, M., Gradinaru, E., et al. 2006. Fossil wood and mid-Eastern Europe terrestrial palaeobiogeography during the Jurassic–early Cretaceous interval. Review of Palaeobotany and Palynology 142: 1532.CrossRefGoogle Scholar
Philips, J.A. 2002. Erosion, isostatic response, and the missing peneplains. Geomorphology 45: 225241.CrossRefGoogle Scholar
Philipson, W.R. & Molloy, B.P.J. 1990. Seedling, shoot, and adult morphology of New Zealand conifers: the genera Dacrycarpus, Podocarpus, Dacrydium and Prumnopitys. New Zealand Journal of Botany 28: 7384.CrossRefGoogle Scholar
Phillipps, J. 1932. Root nodules of Podocarpus. Ecology 13: 189195.CrossRefGoogle Scholar
Phillips, F.M., Zreda, M.G., Benson, L.V., et al. 1996. Chronology for fluctuations in Late Pleistocene Sierra Nevada glaciers and lakes. Science 274: 749751.CrossRefGoogle Scholar
Phipps, C.J. & Taylor, T.N. 1997. Mixed arbuscular mycorrhizas from the Triassic of Antarctica. Mycologia 88: 707714.CrossRefGoogle Scholar
Pichi-Sermolli, R.E.G. & Bizzari, M.P. 1978. The botanical collections (Pteridophyta and Spermatophyta) of the AMF Mares-G.R.S.T.S. expedition to Patagonia, Tierra del Fuego and Antarctica. Webbia 32: 455534.CrossRefGoogle Scholar
Pichot, C. & El Maataoui, M. 2001. Surrogate mother for endangered Cupressus. Nature 412: 39.CrossRefGoogle Scholar
Pichot, C., Borrut, A. & El Maataoui, M. 1998. Unexpected DNA content in the endosperm of Cupressus dupreziana A. Camus seeds and its implications in the reproductive process. Sexual Plant Reproduction 11: 14152.CrossRefGoogle Scholar
Pichot, C., Faby, B. & Hochu, I. 2000. Lack of mother tree alleles in zymograms of Cupressus dupreziana A. Camus embryos. Annals of Forest Science 57: 1722.CrossRefGoogle Scholar
Pickett, S.T.A. & White, P.S. 1985. The Ecology of Natural Disturbance and Patch Dynamics. New York: Academic Press.Google Scholar
Pielou, E.C. 1991. After the Ice Age: The Return of Life to Glaciated North America. Chicago, IL: University of Chicago Press.CrossRefGoogle Scholar
Pierce, A.S 1936. Anatomical interrelationships of the Taxodiaceae. Tropical Woods 46: 115.Google Scholar
Pierce, R.L. 1957. Minnesota Cretaceous pine pollen. Science 125: 26.CrossRefGoogle ScholarPubMed
Pisano, E. 1983. The Magellanic tundra complex. Pp 295329 in Gore, A.J.P. (ed.), Ecosystems of the World, Vol. 4B. Amsterdam: Elsevier.Google Scholar
Pisek, A., Larcher, W., & Unterholzner, R. 1968. Kardinale Temperaturberiche der Photosynthese und Grenztemperaturen des Lebens der Blatter verschiedner Spermatophyten. I. Temperaturminimum der Nettoassimilation, Gefrier- und Frostschdensbereiche der Blatter. Flora 157: 239264.Google Scholar
Pisek, A., Larcher, W., Pack, I. & Unterholzner, R. 1968. Kardinale Temperaturberiche der Photosynthese und Grenztemperaturen des Lebens der Blatter verschiedner Spermatophyten. II. Temperaturmaximum der Netto-Photosynthese und Hitzeresistenz der Blatter. Flora 158: 110128.Google Scholar
Platt, W.J., Evans, C.W. & Rathbun, S.J. 1988. The population dynamics of a long-lived conifer (Pinus palustris). American Naturalist 131: 491525.CrossRefGoogle Scholar
Plavsic, J. 1938. Die Standorte von Picea omorika im sudlichen Drina-Gebiet. Osterreiches Botaniches Zeitschrift 87: 140145.CrossRefGoogle Scholar
Playford, G. & Dettmann, M.E. 1978. Pollen of Dacrydium franklinii Hook. f. and comparable early Tertiary micro-fossils. Pollen et Spores 20: 513534.Google Scholar
Poinar, G. Jr, Lamber, J.B. & Wu, Y. 2004. NMR analysis of amber in the Zubaair formation, Khafji oilfield (Saudi Arabia – Kuwait): coal as an oil source rock? Journal of Petroleum Geology 27: 207209.CrossRefGoogle Scholar
Pokorný, J. 1972. Some notes on the biology of dawn redwood (Metasequoia glyptostroboides Hu et Cheng). Silvicultura Tropica et Cubtropica 2: 149154.Google Scholar
Pole, M. 1995. Late Cretaceous macrofloras of eastern Otago, New Zealand: gymnosperms. Australian Systematic Botany 8: 10671106.CrossRefGoogle Scholar
Pole, M.S. 1997. Miocene conifers from the Manuherika Group, New Zealand. Journal of the Royal Society of New Zealand 27: 355370.CrossRefGoogle Scholar
Pole, M.S. 2009. Vegetation and climate of the New Zealand Jurassic. GFF 131: 105111.CrossRefGoogle Scholar
Pole, M. & Moore, P.R. 2011. A late Miocene leaf assemblage from Coromandel peninsula, New Zealand, and its climatic implications. Alcheringa 35: 103121.CrossRefGoogle Scholar
Pole, M.S., Hill, R.S., Green, N. & Macphail, M.K. 1993. The Late Oligocene Berwick Quarry flora: rainforest in a drying environment. Australian Systematic Botany 6: 399428.CrossRefGoogle Scholar
Polhemus, D.A. 1996. Island arcs, and their influence on Indo-Pacific biogeography. Pp 5166 in Keast, A. & Miller, S.E. (eds.), The Origin and Evolution of Pacific Island Biotas, New Guinea to Western Polynesia: Patterns and Processes. Amsterdam: SPB Academic Publishing.Google Scholar
Ponto, L.L. & Schultz, S.K. 2003. Ginkgo biloba extract: review of CNS effects. Annals of Clinical Psychiatry 2: 683686.Google Scholar
Poole, A.L. & Adams, N.M. 1986. Trees and Shrubs of New Zealand. Wellington: Government Printing Office.Google Scholar
Poole, I. 2000. Fossil angiosperm wood: its role in the reconstruction of biodiversity and palaeoenvironment. Botanical Journal of the Linnean Society 134: 361381.CrossRefGoogle Scholar
Poole, I. & Cantrill, D. 2001. Fossil woods from Williams Point Beds, Livingston Island, Antarctica: a Late Cretaceous southern high latitude flora. Palaeontology 44: 10811112.CrossRefGoogle Scholar
Poole, I. & Page, C.N. 2000. A fossil fern indicator of epiphitism in a Tertiary flora. New Phytologist 148: 117125.CrossRefGoogle Scholar
Poole, I., Mennega, A.M.W. & Cantrill, D.J. 2003. Valdivian ecosystems in the Late Cretaceous and early Tertiary of Antarctica: further evidence from myrtaceous and eucryphiaceous fossil wood. Review of Palaeobotany and Palynology 124: 927.CrossRefGoogle Scholar
Porde, O. & Ramachandran, S. 2012. The phosphorous concentration of common rocks – a potential driver of ecosystem P status. Plant and Soil 367: 115.Google Scholar
Potte, M.D., Ashton, P.S., Kaufmann, L.S. & Plotkin, J.B. 2002. Habitat patterns in tropical rain forests: a comparison of 105 plots in northwest Borneo. Ecology 83: 27822797.CrossRefGoogle Scholar
Potts, M.D. 2003. Drought in a Bornean everwet rain forest. Journal of Ecology 91: 467474.CrossRefGoogle Scholar
Praeger, E.M., Fowler, D. & Wilson, A.C. 1976. Rates of evolution in conifers (Pinaceae). Evolution 30: 637649.CrossRefGoogle Scholar
Prasad, V., Strömberg, C.A.E. Alimohammadian, H. & Sanhi, A. 2005. Dinosaur copralites and the early evolution of grasses and grazers. Science 310: 11771180.CrossRefGoogle ScholarPubMed
Price, R.A., Olsen-Stojkovich, J. & Lowenstein, J.M. 1987. Relationships among the genera of Pinaceae: an immunological comparison. Systematic Botany 12: 9197.CrossRefGoogle Scholar
Pridnya, M.V. 1984. Phytocoenotic status and structure of the Khosta common-yew population in the Caucasus Biosphere Reserve. Soviet Journal of Ecology 15: 16.Google Scholar
Proctor, J. 1971. The plant ecology of serpentine. II. Plant response to serpentine soils. Journal of Ecology 59: 397411.CrossRefGoogle Scholar
Proctor, J. 1971. 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: 827842.CrossRefGoogle Scholar
Proctor, J. & Woodell, S.R.J. 1975. The ecology of serpentine soils. Advances in Ecological Research 9: 255366.CrossRefGoogle Scholar
Pryer, K.M., Smith, A.R. & Skog, J. 1995. Phylogenetic relationships of extant ferns based on evidence from morphology and rbcL sequences. American Fern Journal 85: 205282.CrossRefGoogle Scholar
Pryer, K.M., Schneider, H., Smith, A.R., et al. 2001. Horsetails and ferns are a monophyletic group and the closest living relatives to seed plants. Nature (London) 409: 618622.CrossRefGoogle ScholarPubMed
Punyasena, S.W., Eshel, G. & McElwain, J.C. 2008. The influence of climate on the spatial patterning of Neotropical plant families. Journal of Biogeography 35: 117130.CrossRefGoogle Scholar
Quan, C., Sub, G. & Zhou, Z. 2010. A new Tertiary Ginkgo (Ginkgoaceae) from the Wuyun Formation of Jiaylin, Heilongiang, northeastern China and its paleoenvironmental implications. American Journal of Botany 97: 446457.CrossRefGoogle Scholar
Quinn, C.J. 1964. Gametophyte development and embryology in the Podocarpaceae. I. Podocarpus section Dacrycarpus. Phytomorphology 14: 342351.Google Scholar
Quinn, C.J. 1965. Gametophyte development and embryology in the Podocarpaceae. II. Dacrydium bidwillii. Phytomorphology 16: 8191.Google Scholar
Quinn, C.J. 1966. Gametophyte development and embryology in the Podocarpaceae. III. Dacrydium laxifolium. Phytomorphology 15: 3745.Google Scholar
Quinn, C.J. 1966. Gametophyte development and embryology in the Podocarpaceae. IV. Dacrydium colensoi. General conclusions. Phytomorphology 16: 199210.Google Scholar
Quinn, C.J. 1969. Generic boundaries in the Podocarpaceae. Proceedings of the Linnean Society of New South Wales 94: 166172.Google Scholar
Quinn, C.J. 1986. Embryology in Phyllocladus. New Zealand Journal of Botany 24: 575580.CrossRefGoogle Scholar
Quinn, C.J. 1989. Leaf and wood anatomy as data sources in the Cupressaceae s.l. American Journal of Botany 76 (suppl.): 222.Google Scholar
Quinn, C.J. & Gadek, P. 1981. The sequence of xylem differentiation in the leaves of Cupressaceae s.l. American Journal of Botany 75: 13441351.CrossRefGoogle Scholar
Quinn, C.J. & Gadek, P. 1981. Biflavones in Dacrydium sensu lato. Phytochemistry 20: 677–81.CrossRefGoogle Scholar
Quinn, C.J., Strenge, D.D. & Ranker, T.A. 1994. Phylogenetic relationships amongst the genera of Taxodiaceae and Cupressaceae: evidence from rbcL sequences. Systematic Botany 19: 253262.Google Scholar
Radeloff, V.C., Mladenoff, D.J., Guries, & Boyce, M.S. 2004. Spatial patterns of cone serotiny in Pinus banksiana in relation to forest disturbance. Forest Ecology and Management 189: 133141.CrossRefGoogle Scholar
Radford, A.E., Ahies, H.E. & Bell, C.R. 1968. Manual of the Vascular Flora of the Carolinas. Chapel Hill, NC: University of North Carolina Press.Google Scholar
Raharvelomanana, P., Faure, R., Cambon, A. & Azzaro, M. 1993. B-acoradienol, a sesquiterpene alcohol from Neocallitropsis pancheri. Phytochemistry 33: 235236.CrossRefGoogle Scholar
Rahut, O., Martin, T., Ortiz-Jaureguizar, E. & Puerta, P. 2002. A Jurassic mammal from South America. Nature 416: 165168.CrossRefGoogle Scholar
Rahut, O.W.M., Remes, K., Fechner, R., Cladera, G. & Puerta, P. 2005. Discovery of a short-necked sauropod dinosaur from the Late Jurassic period of Patagonia. Nature 435: 670672.CrossRefGoogle Scholar
Raimondi, N., Lesher, R.D. & Henderson, J.A. 2010. Ecology and distribution of Western redcedar and Alaska yellowcedar in northwestern Washington. US Forest Service Pacific Northwest Research Station General Technical Report PNW-GTR 828.Google Scholar
Rao, A.R. 1943. Nipaniostrobus, a new genus of Dacrydium-like seed-bearing cones, and other silicified plants from the Rajmahal Series. Proceedings of the National Academy of Sciences, India 13: 113133.Google Scholar
Rao, A.R. & Bose, M.N. 1970. Podostrobus gen. nov., a petrified podocarpaceous male cone from the Rajmahal Hills, India. Paleobotanist 19: 8385.Google Scholar
Rao, C., Sharma, G. & Shukla, A. 1997. Distribution of ectomycorrhizal fungi in pure stands of different age groups of Pinus keysia. Canadian Journal of Microbiology 43: 8591.CrossRefGoogle Scholar
Ratter, J.A. & Milne, C. 1973. Chromosome numbers of some primitive angiosperms. Notes from the Royal Botanic Garden Edinburgh 32: 423428.Google Scholar
Raubeson, L.A. & Jansen, R.K. 1992. A rare chloroplast DNA structural mutation is shared by all conifers. Biochemical Systematics and Evolution 20: 1724.CrossRefGoogle Scholar
Raubeson, L.A. & Jansen, R.K. 1993. Phylogenetically informative chloroplast DNA rearrangements in the Podocarpaceae. American Journal of Botany 80(suppl.): 173.Google Scholar
Rauhut, O.W.M. 2006. A brachiosaurid sauropod from the Late Jurassic Cañadón Calcáreo Formation of Chubut, Argentina. Fossil Record 9: 226237.CrossRefGoogle Scholar
Raunkiaer, C. 1934. The Life Forms of Plants and Statistical Plant Geography. Oxford: Clarendon Press.Google Scholar
Raven, P.H. 1977. A suggestion concerning the Cretaceous rise to dominance of the angiosperms. Evolution 31: 451452.CrossRefGoogle ScholarPubMed
Raven, P.H. & Axelrod, D.I. 1972. Plate tectonics and Australasian paleobiogeography. Science 176: 13791386.CrossRefGoogle ScholarPubMed
Read, D.J. & Pérez-Moreno, J. 2003. Mycorrhizas and nutrient cycling in ecosystems: a journey towards relevance? New Phytologist 157: 475492.CrossRefGoogle ScholarPubMed
Read, J., Hope, G.S. & Hill, R. 1990. The dynamics of some Nothofagus dominated rain forests in Papua New Guinea. Journal of Biogeography 17: 185204.CrossRefGoogle Scholar
Redecker, D., Morton, J.B. & Bruns, T.D. 2000. Ancestral lineages of arbuscular mycorrhizal fungi. Molecular Phylogenetics and Evolution 14: 276284.CrossRefGoogle ScholarPubMed
Reed, D. & Nagel, L. 2003. Carbon pools and storage along a temperate to boreal transect in northern Scots pine (Pinus sylvestris) forests. Polish Journal of Ecology 51: 545552.Google Scholar
Regal, P.J. 1977. Ecology and evolution of flowering plant dominance. Science 196: 622629.CrossRefGoogle ScholarPubMed
Rehder, A. 1940. Manual of Cultivated Trees and Shrubs, 2nd edn. New York: MacMillan.Google Scholar
Rehder, A. & Wilson, E.H. 1914. Pinaceae. In Sargent, C.S. (ed.), Plantae Wilsonianae. Cambridge: Cambridge University Press.Google Scholar
Reich, P.B. & Schoettle, A.W. 1988. Role of phosphorous and nitrogen in photosynthesis and whole plant carbon-gain and nutrient use efficiency in eastern white pine. Oecologia 77: 2533.CrossRefGoogle ScholarPubMed
Reich, P.B., Walters, M.B. & Ellsworth, D.S. 1992. Leaf life-span in relation to leaf, plant, and stand characteristics among diverse ecosystems. Ecological Monographs 62: 365392.CrossRefGoogle Scholar
Reid, E.M. & Chandler, M.E.J. 1933. The London Clay Flora. London: British Museum (Natural History).CrossRefGoogle Scholar
Remy, W., Taylor, T.N., Hess, H. & Kerp, H. 1994. Four hundred-million-year-old vesicular arbuscular mycorrhizae. Proceedings of the National Academy of Sciences 91: 1184111843.CrossRefGoogle ScholarPubMed
Ren, Y., Wei, X., Wei, X.-H., et al. 2011. Potential for forest vegetation carbon storage in Fukian Province, China, determined from forest inventories. Plant and Soil 345: 125140.CrossRefGoogle Scholar
Rendle, A.B. 1894. The plants of Milanji, Nyasaland, collected by Mr. Alexander Whyte: Gymnospermae. Transactions of the Linnean Society, London, Botany 4: 6062.Google Scholar
Rendle, A.B. 1911. A contribution to our knowledge of the flora of Gazaland: gymnosperms. Journal of the Linnean Society, Botany 40: 235237.CrossRefGoogle Scholar
Resch, T. 1974. Essai de description morphologique des races majeours de Pinus pinaster. Annales Rech. Forestière Marocana, Tome 4: 116.Google Scholar
Retallack, G.J. 1977. Reconstructing Triassic vegetation of eastern Australasia: a new approach for the biostratigraphy of Gondwanaland. Alcheringa 1: 247278.CrossRefGoogle Scholar
Retallack, G. & Dilcher, D.L. 1981. A coastal hypothesis for the dispersal and rise to dominance of flowering plants. Pp 2777 in Niklas, K.J. (ed.), Paleobotany, Paleoecology and Evolution. New York: Praeger.Google Scholar
Reymanowna, M. & Watson, J. 1976. The genus Frenopsis Schenk and the type species Frenlopsis hoheneggeri (Ettinshausesn) Schenk. Acta Palaeobotanica 17: 1726.Google Scholar
Reyre, Y. 1968. La sculpture de l’exine des pollens des gymnosperms et des chlamydosperms et son utilisation dans l’identification des pollen fossils. Pollen & Spores 10: 197220.Google Scholar
Ribbens, E., Silander, J.A. & Pacala, S.W. 1994. Seedling recruitment in forests: calibrating models to predict patterns of tree seedling dispersion. Ecology 75: 17941806.CrossRefGoogle Scholar
Richards, P.W. 1964. The Tropical Rainforest: an Ecological Study, 2nd edn. Cambridge: Cambridge University Press.Google Scholar
Richardson, B.A., Brunsfeld, J., & Klopfenstein, N.B. 2002. DNA from bird-dispersed seed and wind-disseminated pollen provides insights into postglacial colonization and population genetic structure of white bark pine (Pinus albicaulis). Molecular Ecology 11: 215227.CrossRefGoogle Scholar
Richardson, S.J., Peltzer, D.A., Allen, R.B., McGlone, M.S. & Parfitt, R.L. 2004. Rapid development of phosphorous limitation in temperate rainforest along the Franz Josef soil chronosequence. Oecologia 139: 267276.CrossRefGoogle Scholar
Riebe, C.S., Kirchner, J.W. & Finkel, R.C. 2004. Sharp decrease in long-term chemical weathering rates along an altitudinal transect. Earth and Planetary Science Letters 218: 421434.CrossRefGoogle Scholar
Rigg, C.M. 2001. Orchestrating ecosystem management: challenges and lessons from Sequoia National Forest. Conservation Biology 15: 7890.CrossRefGoogle Scholar
Ritchie, J.C. 1986. Climate change and vegetation response. Vegetatio 67: 6574.CrossRefGoogle Scholar
Ritchie, J.C. & MacDonald, G.M. 1986. The patterns of post-glacial spread of white spruce. Journal of Biogeography 13: 527540.CrossRefGoogle Scholar
Roberts, D.L., Bamford, M. & Millsteed, B. 1997. Permo-Triassic macro-plant fossils in the Fort Grey silcrete, east London. South African Journal of Geology 100: 157168.Google Scholar
Robertson, A. Some points in the morphology of Phyllocladus alpinus Hook. Annals of Botany (London): 20: 256265.Google Scholar
Robertson, F. 1907. The Taxoideae: a phylogenetic study. New Phytologist 6: 92102.CrossRefGoogle Scholar
Rodewald, A.D. & Vitz, A.C. 2005. Edge- and area-sensitivity of shrubland birds. Journal of Wildlife Management 69: 681688.CrossRefGoogle Scholar
Rodríguez, R., Matthei, O. & Quezada, M. 1983. Flora Arbórea de Chile. Concepción: Editorial de la Universidad de Concepción.Google Scholar
Rogers, D.L. 2004. In situ genetic conservation of a naturally restricted and commercially widespread species, Pinus radiata. Forest Ecology and Management 197: 311322.CrossRefGoogle Scholar
Rogers, H.M. 1999. Stand dynamics of Dacrydium cupressinum dominated forest on glacial terraces, south Westland, New Zealand. Forest Ecology and Management 117: 111128.CrossRefGoogle Scholar
Romero, E.J. 1986. Paleogene phytogeography and climatology of South America. Annals of the Missouri Botanical Garden 73: 449461.CrossRefGoogle Scholar
Romme, W.H. & Knight, D.H. 1981. Fire frequency and subalpine forest succession along a topographic gradient in Wyoming. Ecology 62: 319326.CrossRefGoogle Scholar
Rook, D.A. & Page, C.N. 1992. The importance of arboreta in the eventuality of climate change. Pp 85100 in Mustila Arboretum, 90th Anniversary Jubilee Symposium of Mustila Arboretum. Helsinki: Mustila Arboretum.Google Scholar
Roselt, G. 1958. Neue Koniferen aus dem Unteren Keuper und ihre Bezichungen zu verwandten fossilen und rezenten. Freidrich-Schiller-Universitat Wiss. Zeitschr. Palynol. Jahrg. 5: 75115.Google Scholar
Ross, J.R. & Duncan, R.E. 1949. Cytological evidence of hybridisation between Juniperus virginiana and J. horizontalis. Bulletin of the Torrey Botanical Club 76: 414429.CrossRefGoogle Scholar
Rothwell, G.W. 1976. Primary vasculature and gymnosperm systematics. Review of Paleobotany and Palynology 22: 193206.CrossRefGoogle Scholar
Rothwell, G.W. 1981. The Callistophytales (Pteridospermopsida): reproductive sophisticated Paleozoic gymnosperms. Review of Palaeobotany and Palynology 32: 103121.CrossRefGoogle Scholar
Rothwell, G.W. 1982. New interpretation of the earliest conifers. Review of Palaeobotany and Palynology 37: 728.CrossRefGoogle Scholar
Rothwell, G.W. 1982. Cordaianthus dusquesnensis sp. nov., anatomically preserved ovulate cones from the Upper Pennsylvanian of Ohio. American Journal of Botany 69: 239243.CrossRefGoogle Scholar
Rothwell, G.W. 1985. The role of comparative morphology and anatomy in interpreting the systematics of fossil gymnosperms. Botanical Review 51: 319327.CrossRefGoogle Scholar
Rothwell, G.W. 1987. Complex Paleozoic Filicales in the evolutionary radiation of ferns. American Journal of Botany 74: 458461.CrossRefGoogle Scholar
Rothwell, G.W. 1987. The role of development in plant phylogeny: a palaeobotanical perspective. Review of Paleobotany and Palynology 50: 97114.CrossRefGoogle Scholar
Rothwell, G.W. 1991. Botryopteris forensis (Botryopteridaceae), a trunk epiphyte of the tree-fern Psaroniums. American Journal of Botany 78: 782788.CrossRefGoogle Scholar
Rothwell, G.W. 1996. Pteridophytic evolution, an often underappreciated phytological success story. Review of Palaeobotany and Palynology 90: 209222.CrossRefGoogle Scholar
Rothwell, G.W. 1996. Phylogenetic relationships of ferns, a palaeobotanical perspective. Pp 395404 in Camus, J.M., Gibby, M. & Johns, R.J. (eds.), Pteridology in Perspective. London: Royal Botanic Gardens, Kew.Google Scholar
Rothwell, G.W. 1999. Fossils and ferns in the resolution of land plant phylogeny. Botanical Review 65: 188218.CrossRefGoogle Scholar
Rothwell, G.W. & Scott, A.C. 1985. Ecology of the Lower Carboniferous plant remains from Oxroad Bay, East Lothian, Scotland. American Journal of Botany 2: 899.Google Scholar
Rothwell, G.W. & Serbet, R. 1944. Lignophyte phylogeny and the evolution of spermatophytes: a numerical cladistic analysis. Systematic Botany 19: 443482.CrossRefGoogle Scholar
Rothwell, G.W. & Stockey, R.A. 1991. Onoclea sensibilis in the Paleocene of North America, a dramatic example of structural and ecological stasis. Review of Palaeobotany and Palynology 70: 113124.CrossRefGoogle Scholar
Rouane, M.L., Gondran, M. & Woltx, P. 1988. Hétérocotylie et évolution cotyledonaire chex Podocarpineae. Gaussenia 4: 715.Google Scholar
Rougier, G.W., Martinelli, A.G., Forasiepi, A.M. & Novacek, M.J. 2007. New Jurassic mammals from Patagonia, Argentina: a reappraisal of australosphenidan morphology and interrelationships. American Museum Novitates 3566: 154.CrossRefGoogle Scholar
Rourke, J.P. 1991. Tetraclinis articulata, a hitherto unrecorded naturalised alien conifer in South Africa. Bothalia 21: 6264.CrossRefGoogle Scholar
Rouse, G.E. 1959. Plant microfossils from Kootenay coal-measures strata of British Columbia. Micropaleontology 5: 303324.CrossRefGoogle Scholar
Royer, D.L. 2001. Stomatal density and stomatal index as indicators of paleo-atmospheric CO2 concentration. Review of Palaeobotany and Palynology 114: 128.CrossRefGoogle Scholar
Royer, D.L., Osborne, C.P. & Beerling, D.J. 2002. High CO2 increases the freezing sensitivity of plants: implications for paleoclimatic reconstructions from fossil floras. Geology 30: 963966.2.0.CO;2>CrossRefGoogle Scholar
Royer, D.L., Wilf, P., Janesko, D.A., Kowalski, E.A. & Dilccher, D.J. 2005. Correlations of climate and plant ecology to leaf size and shape: potential proxies for the fossil record. American Journal of Botany 92: 11411151.CrossRefGoogle ScholarPubMed
Rozas, V., Olano, J.M., DeSoto, L. & Bartolome, D. 2008. Large-scale structural variation and long-term growth dynamics of Juniperus thurifera trees in a managed woodland in Soria, central Spain. Annals of Forest Science 65: 809820.CrossRefGoogle Scholar
Rubner, K. 1943. Das Area der Sudetenlärche. Tharandler Forestlia Jahrbucher 94: 199.Google Scholar
Rudloff, E. von. 1964. Gas-liquid chromatography of terpenes x: the volatile oils of the leaves of sitka and engelmann spruce. Canadian Journal of Chemistry 42: 10571062.CrossRefGoogle Scholar
Rudolph, D.C., Ely, C.A., Schaefer, R.R., Williamson, J.H. & Thill, R.E. 2006. Monarch (Danaus plexippus L, Nymphalidae) migration, nectar resources and forest regimes in the Ouachita Mountains of Arkansas. Journal of the Lepidopterists Society 60: 165170.Google Scholar
Rudolph, T.D. & Yeayman, C.W. 1982. Genetics of Jack pine. US Department of Agriculture, Forest Service, Research Paper WO-38.Google Scholar
Rundel, P.W. 1972. Habitat restriction in giant Sequoia: the environmental control of grove boundaries. American Midland Naturalist 87: 8199.CrossRefGoogle Scholar
Rundel, P.W., Sharfi, M.A., Kohl-Rundel, J. & Middleton, D.J. 2016. Dacrydium elatum (Podocarpaceae) in the mountain cloud forests of Bokor Mountain, Cambodia. Cambodian Journal of Natural History 2: 9097.Google Scholar
Rundell, P., Parsons, D. & Gordon, D. 1977. Montane and subalpine vegetation of the Sierra Nevada and Cascade ranges. Pp 559599 in Barbour, M. & Major, J. (eds.), Terrestrial Vegetation of California. New York: Wiley.Google Scholar
Rune, O. 1953. Plant life on serpentine and related rocks in the north of Sweden. Acta Phytogeographica Suecica 36: 1139.Google Scholar
Runions, C.J. & Owens, J.N. 1996. Pollen scavenging and rain involvement in the pollination mechanism of interior spruce. Canadian Journal of Botany 74: 115124.CrossRefGoogle Scholar
Runkle, J.R. 1981. Gap-phase regeneration in some old-growth forests of the eastern United States. Ecology 62: 10411051.CrossRefGoogle Scholar
Rushforth, K. 1976. Tree genera – 7. The Spruces – Picea. Arboricultural Journal 3: 246255.CrossRefGoogle Scholar
Rushforth, K.D. 1986. Notes on Chinese firs 3. Notes from the Royal Botanic Garden Edinburgh 43: 269275.Google Scholar
Rushforth, K.D. 1989. Two new species of Abies (Pinaceae) from western Mexico. Notes from the Royal Botanic Garden Edinburgh 46: 101109.Google Scholar
Russo, S.E., Davies, S.J., King, D.A. & Tan, S. 2005. Soil-related performance variations and distributions of tree species in a Bornean rain forest. Journal of Ecology 93: 879889.CrossRefGoogle Scholar
Russo, S.E., Brown, P., Tan, S. & Davies, S. 2008. Interspecific demographic trade-offs and soil-related habitat associations of tree species along resource gradients. Journal of Ecology 96: 192203.CrossRefGoogle Scholar
Rydgren, K. & Hestmark, G. 1996. The soil propagule bank in a boreal old-growth spruce forest: changes with depth and relationship to aboveground vegetation. Canadian Journal of Botany 75: 121128.CrossRefGoogle Scholar
Rzedowski, J. 1978. The Vegetation of México. Limusa: Editorial Limusa.Google Scholar
Sabzehzari, M., Zeinali, M. & Nashawi, M.R. 2020. Alternative advances and metabolic engineering of taxol: advances and future perspectives. Biotechnology Advances 43.CrossRefGoogle ScholarPubMed
Safarov, I.S. 1986. Rare and disappearing species of eastern Transcaucasian USSR dendroflora and their preservation. Botanicheskii Zhurnal 71: 102108.Google Scholar
Sage, R.F. 2004. The evolution of C-4 photosynthesis. New Phytologist 161: 341370.CrossRefGoogle Scholar
Sahni, B. & Mitra, A.K. 1927. Notes on the anatomy of some New Zealand species of Dacrydium. Annals of Botany 41: 7688.Google Scholar
Saika, D., Khanuja, S.P.S., Shasany, A.K., et al. 2000. Assessment of diversity amongst Taxus walichianus accessions from northeast India using RAPD analysis. Plant Genetic Resources Newsletter 121: 2731.Google Scholar
Saitou, N & Nei, M. 1987. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Molecular Biology and Evolution 4: 406425.Google Scholar
Saki, K. & Suzuki, Y. 1955. Studies on competition in plants. V. Competition between allopolyploids and their diploid parents. Journal of Genetics 53: 585590.CrossRefGoogle Scholar
Sakisaka, M. 1929. On the morphological significance of seed-bearing leaves of Ginkgo biloba. Botanical Magazine Tokyo 41: 273278.CrossRefGoogle Scholar
Salter, J. 2007. Matai and miro (Prumnopitys): the inside story on the rumour of their separation. New Zealand Journal of Botany 45: 294295.Google Scholar
Salzer, M.W. & Hughes, M.K. 2007. Bristlecone pine tree rings and volcanic eruptions over the last 1500 yr. Quaternary Research 67: 5768.CrossRefGoogle Scholar
Sánchez-Zabala, J., Majadan, J., Martin-Rodrigues, N., et al. 2013. Physiological aspects underlying the improved outplanting performance of Pinus pinaster Ait. seedlings associated with ectomycorrhizal inoculation. Mycorrhiza 23: 627640.CrossRefGoogle ScholarPubMed
Sander, P.M, Christian, A., Clauss, M., et al. Biology of the sauropod dinosaurs: the evolution of gigantism. Biological Reviews 86: 117155.CrossRefGoogle Scholar
Sanhi, B. 1920. The structure and affinities of Acmopyle pancheri Pilger. Philosophical Transactions of the Royal Society of London B 210: 253310.Google Scholar
Sanhi, B. 1948. The Pentoxyleae: a new group of Jurassic Gymnosperms from the Rajmahal Hills of India. Botanical Gazette 110: 4780.Google Scholar
Sanmartin, I. & Ronquist, F. 2004. Southern hemisphere biogeography inferred by event-based models: plant versus animal patterns. Systematic Biology 53: 216243.CrossRefGoogle ScholarPubMed
Santiago, L.S. & Dawson, T.E. 2014. Light use efficiency of California redwood forest understory plants along a moisture gradient. Oecologia 174: 351363.CrossRefGoogle Scholar
Santisuk, T. 1988. An account of the vegetation of northern Thailand. Geoecological Research 5: 1101.Google Scholar
Sarmaja-Korjonen, K., Vasari, Y. & Haeggstrom, C.-A. Taxus baccata and influence of Iron Age man on the vegetation in Adland, SW Finland. Annales Botanici Fennici 28: 143159.Google Scholar
Satake, Y. 1934. On the systematic importance of the vascular course in the cone scale of the Japanese Taxodiaceae (preliminary report). Botanical Magazine Tokyo 48: 186205 (in Japanese, with English summary).CrossRefGoogle Scholar
Sauquet, H., Weston, P.H., Anderson, C.L., et al. 2009. Contrasted patterns of hyperdiversification in Mediterranean hotspots. Proceedings of the National Academy of Sciences, USA 106: 221225.CrossRefGoogle ScholarPubMed
Savage, T. 1985. A Georgia station for Torreya taxifolia survives. Florida Scientist 46: 6264.Google Scholar
Savolainen, O. & Kuittinnen, H. 2000. Small population processes. Pp 91100 in Young, A. (ed). Forest Conservation: Genetics. Collingwood: CSIRO Publishing.Google Scholar
Savva, J.W. & Vaganova, E.A. 2006. Genetic and environmental effects assessment in Scots pine provenances planted in Central Siberia. Mitigation and Adaptation Strategies for Global Change 11: 269290.CrossRefGoogle Scholar
Sawyer, J.O., Sillett, S.C., Libby, W.J., et al. 2000. Redwood trees, communities, and ecosystems: a close look. Pp 81118 in Noss, R.F. (ed.), The Redwood Forest: History, Ecology and Conservation of the Coast Redwoods. Washington, DC: Island Press.Google Scholar
Sawyer, J.O., Sillett, S.C., Popenoe, J.H., et al. 2000. Characteristics of redwood forests. Pp 3979 in Noss, R.F. (ed.), The Redwood Forest: History, Ecology and Conservation of the Coast Redwoods. Washington, DC: Island Press.Google Scholar
Saxton, W.T. 1901. Contributions to the life history of Callitris. Annals of Botany 24: 557569.Google Scholar
Saxton, W.T. 1909. Preliminary account of the ovule, gametophytes and embryo of Widdringtonia cupressoides. Botanical Gazette 48: 161178.CrossRefGoogle Scholar
Saxton, W.T. 1910. Contributions to the life-history of Widdringtonia cupressoides. Botanical Gazette 50: 3148.CrossRefGoogle Scholar
Saxton, W.T. 1913. Contributions to the life-history of Tetraclinis articulata Masters, with some notes on the phylogeny of the Cupressoideae and Callitroideae. Annals of Botany 27: 577605.CrossRefGoogle Scholar
Saxton, W.T. 1913. The classification of conifers. New Phytologist 12: 242262.CrossRefGoogle Scholar
Saxton, W.T. 1913. Contributions to the life-history of Actinostrobus pyramidalis. Annals of Botany 27: 321345.CrossRefGoogle Scholar
Saxton, W.T. 1929. Notes on conifers. 1. The older fertile ovule of Saxegothaea. Annals of Botany 43: 375377.CrossRefGoogle Scholar
Saxton, W.T. 1930. The root nodules of the Podocarpaceae. South African Journal of Science 27: 323325.Google Scholar
Saxton, W.T. 1934. Notes on conifers. IX. The ovule and embryogeny of Widdringtonia. Annals of Botany 48: 429431.CrossRefGoogle Scholar
Scattolin, L., Montecchio, L., Mosca, E. & Agerer, R. 2015. Vertical distribution of the ectomycorrhizal community in the top soil of Norway spruce stands. European Journal of Forest Research 101: 111.Google Scholar
Schauer, A.J., Schoettle, A.W. & Boyce, R.L. 2001. Partial cambial mortality in high-elevation Pinus aristata (Pinaceae). American Journal of Botany 88: 646652.CrossRefGoogle ScholarPubMed
Scheiner, S. & Istock, C. 1994. Species enrichment in a traditional landscape, northern lower Michigan. Canadian Journal of Botany 72: 217226.CrossRefGoogle Scholar
Schellevis, N. & Schouten, J. 1999. Species accounts: Clanwilliam cedar (Widdringtonia cedarbergensis J.A.Marsh). Pp 9091 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Schellevis, N. & Schouten, J. 1999. Species accounts: giant sequoia (Sequoiadendron giganteum (Lindl.) Buchholz). Pp 9294 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Schellevis, N. & Schouten, J. 1999. Species accounts: alerce (Fitzroya cupressoides (Molina) I.M. Johnston). Pp 9597 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Schellevis, N. & Schouten, J. 1999. Species accounts: Sicilian fir (Abies nebrodensis (Lojac) Mattei). Pp 9798 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Schemske, D., Husband, B.B., Ruckelhaus, M. et al. 1994. Evaluating approaches to the conservation of rare and endangered plants. Ecology 75: 854–606.CrossRefGoogle Scholar
Scheublin, T.R., Ridgway, K.P., Young, J.P.W. & van der Heijden, M.G.A. 2004. Nonlegumes, legumes, and root nodules harbor different arbuscular mycorrhizal fungal communities. Applied Environmental Microbiology 70: 62406246.CrossRefGoogle ScholarPubMed
Schirone, B., Piovesan, G. & Bellarosa, R. 1991. A taxonomic analysis of seed proteins in Pinus spp. (Pinaceae). Plant Systematiccs and Evolution 178: 4353.CrossRefGoogle Scholar
Schlarbaum, S.E. & Tschiya, T. 1975. Chromosomes of incense cedar. Journal of Heredity 66: 4142.CrossRefGoogle Scholar
Schlarbaum, S.E. & Tsuchiya, T. 1985. Karyological derivation of Sciadopitys vertilillata Sieb et Zucc. from a pro-taxodiaceous ancestor. Botanical Gazette 146: 264267.CrossRefGoogle Scholar
Schluter, D. 2000. The Ecology of Adaptive Radiation. Oxford: Oxford University Press.CrossRefGoogle Scholar
Schluter, D. 2001. Ecology and the origin of species. Trends in Ecology and Evolution 16: 372380.CrossRefGoogle ScholarPubMed
Schmid, M. 1974. Végétation du Viet-Nam. Paris: ORSTOM.Google Scholar
Schmid, R. 1999. Regional action plan: conifers of the Californian floristic province. Pp 7588 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Schmid, R. & Schmid, M. 2012. Naturalization of Sequoiadendron giganteum (Cupressaceae) in montane southern California. Aliso 30: 1932.CrossRefGoogle Scholar
Schmidtling, R.C. 1983. Genetic variation in fruitfulness in a loblolly pine (Pinus taeda L.) seed orchard. Silvae Genetica 32: 7680.Google Scholar
Schmidtling, R.C. & Hipkins, V. 1998. Genetic diversity in longleaf pine (Pinus palustris): influence of historical and prehistorical events. Canadian Journal of Forest Research 28: 11351145.CrossRefGoogle Scholar
Schmidtling, R.C. & Hipkins, V. 2000. Evolutionary relationships of slash pine (Pinus elliotti) with its temperate and tropical relatives. South African Forest Journal 190: 7378.Google Scholar
Schmithusen, J. 1960. Conifers in the forest association of the northern Andes. Vegetation 9: 313327.Google Scholar
Schmithusen, J. 1960. Die nadelhölzer in den waldgesellschaften der südlichen anden. Vegetation 9: 313327.CrossRefGoogle Scholar
Schneider, H., Schuettpelz, E., Pryer, K.M. 2004. Ferns diversified in the shadow of angiosperms. Nature 428: 553557.CrossRefGoogle Scholar
Schoenhut, K. 2005. Environmental implications of the preservation of chloroplast ultrastructure in Eocene Metasequoia leaves. Paleobiology 31: 424433.CrossRefGoogle Scholar
Schoennagel, T., Turner, M.G. & Romme, W.H. 2003. The influence of fire interval and serotiny on post-fire lodgepole pine density in Yellowstone National Park. Ecology 84: 29672978.CrossRefGoogle Scholar
Schonfeld, E. 1955. Metasequoia in der Westdeutschen Braunkohle. Senkenbergiana Lethaea 36: 389399.Google Scholar
Schoonraad, E. & van der Schiff, H.P. 1974. Anatomy of the leaves of the genus Podocarpus in South Africa. Phytomorphology 24: 7585.Google Scholar
Schoonraad, E. & van der Schiff, H.P. 1975. Distribution of some interesting morphological aspects of South African Podocarpaceae. Biossiera 24: 135144.Google Scholar
Schrepfer, P., Buettner, A., Goerner, C., et al. 2016. Identification of amino acid networks governing catalysis in the closed complex of class I terpene synthases. Proceedings of the National Academy of Sciences, USA 113(8): E958E967. doi: 10.1073/pnas.1519680113.CrossRefGoogle ScholarPubMed
Schrumpf, M., Axmacher, J.C., Zech, W., Lehmann, J., & Lyaruu, H.V.C. 2007. Long-term effects of rainforest disturbance on the nutrient composition of throughfall, organic layer percolate and soil solution at Mt. Kilimanjaro. Science of the Total Environment 376: 241254.CrossRefGoogle Scholar
Schuettpelz, E., Korall, P. & Pryer, M. 206. Plastid atpA data provide improved support for deep relationships among ferns. Taxon 55: 897906.CrossRefGoogle Scholar
Schulte, P., Alegret, L., Arenillas, I. 2010. The Chixulub asteroid impact and mass extinction at the Cretaceous–Paleogene boundary. Science 327: 12141218.CrossRefGoogle Scholar
Schulter, D. 2000. The Ecology of Adaptive Radiation. Oxford: Oxford University Press.CrossRefGoogle Scholar
Schultes, R.E. & Raffauf, R.F. 1990. The Healing Forest: Medicinal and Toxic Plants of the Northwest Amazonia. Portland, OR: Dioscoroides Press.Google Scholar
Schweitzer, H.-J. 1963. Der weibliche Zapfen von Pseudovoltzia liebeanna und seine Bedeuttung für die Phylogenie der Koniferen. Palaeontographica B 113: 129.Google Scholar
Schweitzer, H.-J. 1966. Volzia hexagona (Bischoff) Geinitz aus dem Mittelerenperm Westdeutschlands. Palaeontographica B 239: 122.Google Scholar
Schwendemann, A.B., Decombeix, A.-L., Taylor, T.N., Taylor, E.L. & Krings, M. 2011. Morphological and functional stasis in mycorrhizal root nodules as exhibited by a Triassic conifer. Proceedings of the National Academy of Sciences, USA 108: 1363013634.CrossRefGoogle ScholarPubMed
Schwilk, D.W. & Keeley, J.E. 2006. The role of fire refugia in the distribution of Pinus sabiniana (Pinaceae) in the southern Sierra Nevada. Madrõno 53: 364372.CrossRefGoogle Scholar
Scotland, R.W. & Sanderson, M.J. 2004. The significance of few versus many in the tree of life. Science 303: 643.CrossRefGoogle Scholar
Scott, A.C. 1974. The earliest conifer. Nature 251: 707708.CrossRefGoogle Scholar
Scott, A.C. & Chaloner, W.G. 1983. The earliest fossil conifer from the Westphalian B of Yorkshire. Proceedings of the Royal Society of London B 220: 163182.Google Scholar
Seddon, G. 1974. Xerophytes, xeromorphs and sclerophylls: the history of some concepts in ecology. Biological Journal of the Linnean Society 6: 6587.CrossRefGoogle ScholarPubMed
Seddon, G. 1984. Characteristics and classification of rainforest. Landscape Australia 4(84): 276285.Google Scholar
Selosse, M.-A. & Le Tacon, F. 1998. The land flora: a phototroph–fungus partnership? Trends in Ecology and Evolution 13: 1520.CrossRefGoogle Scholar
Selwood, B.W. & Valdes, P.J. 2006. Mesozoic climates: general circulation models and the rock record. Sedimentary Geology 190: 269287.CrossRefGoogle Scholar
Serbet, R. & Stockey, R.A. 1991. Taxodiaceous pollen cones from the Upper Cretaeous (Horseshoe Canyon Formation) of Drumheller, Alberta, Canada. Review of Palaeobotany and Palynology 70: 6776.CrossRefGoogle Scholar
Serlin, B.S., Delevoryas, T. & Weber, R. 1981. A new conifer pollen cone from the Upper Cretaceous of Coahuila, Mexico. Review of Palaeobotany and Palynology 31: 241248.CrossRefGoogle Scholar
Seward, A.C. 1895. The Wealden Flora, part 2. Gymnospermae. Catalogue of Mesozoic Plants in the Department of Geology. London: British Museum (Natural History).Google Scholar
Seward, A.C. 1917. Fossil Plants: A Textbook for Students of Botany and Geology. Cambridge: Cambridge University Press.Google Scholar
Seward, A.C. 1926. The Cretaceous plant-bearing rocks of western Greenland. Philosophical Transactions of the Royal Society of London B 215: 57175.Google Scholar
Seward, A.C. & Ford, S.O. 1906. The Araucariaceae, recent and extinct. Philosophical Transactions of the Royal Society of London B 198: 305411.Google Scholar
Seward, A.C. & Gowan, J. 1900. The maidenhair tree (Ginkgo biloba L.). Annals of Botany 14: 109154.CrossRefGoogle Scholar
Seward, A.C. & Sanhi, B. 1920. Indian Gondwana plants: a revision. Memoirs of the Geological Survey of India, Palaeontolgica Indica, n.s. 7: 141.Google Scholar
Shah, A., Li, D.-Z, Gao, L.M., Li, H.-T. & Möller, M. 2008. Genetic diversity within and among populations of the endangered species Taxus fauna (Taxaceae) from Pakistan and implications for its conservation. Biochemical Systematics and Ecology 36: 183193.CrossRefGoogle Scholar
Shao, X. & Wu, X. 1994. Tree-ring chronologies for Pinus armandii Franch from Huashan, China. Acta Geographica Sinica 49: 174181.Google Scholar
Shaw, D.C., Huso, M. & Bruner, H. 2000. Basal area growth impacts of dwarf mistletoe on western hemlock in an old-growth forest. Canadian Journal of Forest Research 36: 576583.Google Scholar
Shaw, W.R. 1896. Contribution to the life history of Sequoia sempervirens. Botanical Gazette 21: 332339.CrossRefGoogle Scholar
Sheikh, M.T. & Kolhe, P.D. 1982. A new petrified Podocarpoovuites chitaleyi from a new locality of Nagpur, Decan Maharashtra (India). Votanique 10: 99108.Google Scholar
Shen, Y.-B. 1999. Subdivision and correlation of the Eocene Fossil Hill Formation from King George Island, West Antarctica. Korean Journal of Polar Research 10: 9195.Google Scholar
Sherriff, D.W., Nambiar, E.K.S. & Fife, D.N. 1986. Relationships between nutrient status, carbon assimilation and water use efficiency in Pinus radiata (D. Don) needles. Treee Physiology 2: 7388.CrossRefGoogle Scholar
Shi, G.L., Zhou, Z.Y. & Xie, Z.M. 2010. A new Cephalotaxus and associated epiphyllous fungi from the Oligocene of Guangxi, South China. Review of Palaeobotany and Palynology 161: 179195.CrossRefGoogle Scholar
Shinneman, D.J. & Baker, W.L. 1997. Nonequilibrium dynamics between catastrophic disturbances and old-growth forests in ponderosa pine landscapes of the Black Hills. Conservation Biology 11: 12761288.CrossRefGoogle Scholar
Shirasawa, H. & Koyama, M. 1917. Some new species of Picea and Abies in Japan. Botanical Magazine Tokyo 27.Google Scholar
Shropshire, C., Wagner, R.G., Bell, F.W. & Swanton, C.J. 2001. Light attenuation by early successional plants of the boreal forest. Canadian Journal of Forest Research 31: 812823.CrossRefGoogle Scholar
Siepielski, A.M. & Benkman, C.W. 2004. Interactions among moths, crossbills, squirrels, and lodgepole pine in a geographic selection mosaic. Evolution 58: 95101.Google Scholar
Siepielski, A.M. & Benkman, C.W. 2007. Convergent patterns in the selection mosaic for two North American bird-dispersed pines. Ecological Monographs 77: 203220.CrossRefGoogle Scholar
Siepielski, A.M. & Benkman, C.W. 2007. Selection by predispersal seed predator constrains the evolution of avian seed dispersal in pines. Functional Ecology 21: 611618.CrossRefGoogle Scholar
Siepielski, A.M. & Benkman, C.W. 2008. A seed predator drives the evolution of a seed-dispersal mutualism. Proceedings of the Royal Society of London B 275: 19171925.Google Scholar
Silba, J. 1981. An International Census of the Coniferae. I. Phytologia Memoirs VII. Plainfield, NJ: Moldenke.Google Scholar
Silba, J. 1981. Revised generic concepts of Cupressus L. (Cupressceae). Phytologia 49: 390399.Google Scholar
Silba, J. 1983. Addendum to revised generic concepts of Cupressus L. (Cupressaceae). Phytologia 52: 349361.Google Scholar
Silba, J. 1988. A new species of Cupressus from Tibet (Cupressaceae). Phytologia 65: 333336.Google Scholar
Silba, J. 2006. The chronological history and taxonomic variation of the genus Cupressus (Cupressaceae) in India. Acta Botanica Yunnanica 28: 469470.Google Scholar
Silver, W.L., Scatena, F.N., Johnson, A.H., Siccama, T.G. & Sánchez, M.J. 1994. Nutrient availability in montane wet tropical forest: spatial patterns and methodological considerations. Plant and Soil 164: 129145.CrossRefGoogle Scholar
Silvertown, J., Franco, M., & Menges, M. 1996. Interpretation of elasticity matrices as an aid to the management of plant populations for conservation. Conservation Biology 10: 591597.CrossRefGoogle Scholar
Silvertown, J., Dodd, M.E., Gowing, D.J.D. & Mountford, J.O. 1999. Hydrologically defined niches reveal a basis for species richness in plant communities. Nature 400: 6163.CrossRefGoogle Scholar
Simak, M. 1962. Karyotype analysis of Larix deciduas Mill from different provenances. Medd. Statens Skogsfersk Institut 51: 122.Google Scholar
Simak, M. 1964. Karyotype analysis of Siberian Larch (Larix sibirica Lebed. and Larix sukaczewii Dyl.). Studia Forestalia Suecica 17: 115.Google Scholar
Simal, M. 1967. Seed weight of larch from different provenances (Larix decidua Mill). Studia Forestalia Suecica 57: 130.Google Scholar
Simkin, S.M., Michner, W.K. & Wyatt, R. 2004. Mound microclimate, nutrients and seedling survival. American Midland Naturalist 152: 1224.CrossRefGoogle Scholar
Simon, L., Bousquet, J., Lévesque, R.C. & Lalonde, M. 1993. Origin and diversification of endomycorrhizal fungi and coincidence with vascular land plants. Nature 363: 6769.CrossRefGoogle Scholar
Simonin, K.A., Santiago, L.S. & Dawson, T.E. 2009. Fog interception by Sequoia sempervirens (D. Don) crowns decouples physiology from soil water deficit. Plant, Cell and Environment 32: 882892.CrossRefGoogle ScholarPubMed
Sinard, S.W. & Durall, D.M. 2004. Mycorrhizal networks: a review of their extent, function and importance. Canadian Journal of Botany 82: 11401165.CrossRefGoogle Scholar
Sinclair, W.T., Morman, J.D. & Ennos, R.A. 1999. The postglacial history of Scots pine (Pinus sylvestris L.) in western Europe: evidence from mitochondrial DNA variations. Molecular Ecology 8: 8388.CrossRefGoogle Scholar
Singh, H. 1961. The life history and systematic position of Cephalotaxus drupacea Sieb & Zucc. Phytomorphology 11: 153197.Google Scholar
Singh, H. 1978. Embryology of Gymnosperms. Berlin: Bruder Borntrager.Google Scholar
Sinha, B. 2002. Pines in the Himalayas: past, present and future scenario. Energy and Environment 13: 873882.CrossRefGoogle Scholar
Sinninghe Damsté, J.S., van Bentum, E.C., Reichart, G.J., Pros, J. & Schouten, S. 2010. A CO2 decrease-driven cooling and increased latitudinal temperature gradient during the mid-Cretaceous Oceanic Anoxic Event 2. Earth and Planetary Science Letters 293: 97103.CrossRefGoogle Scholar
Sinnott, E.W. 1913. The morphology of the reproductive structure in the Podocarpinae. Annals of Botany 27: 3982.CrossRefGoogle Scholar
Siwecki, R. 1978. Diseases and parasitic insects of the yew. Pp 103109 in Bartkowiak, S., Bugala, W., Czartoryski, A., et al. (eds.) The Yew: Taxus baccata. Warsaw: Department of the National Center for Scientific and Technical, and Economic Information.Google Scholar
Sizonenko, T. & Zagirova, S. 2012. Seasonal changes in the structure of Picea obovata mycorrhizas in the middle taiga subzone. Russian Journal of Ecology 43: 107110.CrossRefGoogle Scholar
Skottsberg, C. 1960. Remarks of the plant geography of the southern cold temperate zone. Proceedings of the Royal Society B 152: 447457.Google Scholar
Smellie, J.L. 1981. A complete arc-trench system recognised in Gondwana sequences of the Antarctic peninsula region. Geological Magazine 118: 139159.CrossRefGoogle Scholar
Smiet, A.C. 1992. Forest ecology on Java: human impact and vegetation of montane forest. Journal of Tropical Ecology 8: 129152.CrossRefGoogle Scholar
Smith, A.C. 1972. An appraisal of the orders and families of primitive extant angiosperms. Journal of the Indian Botanical Society 50A: 215226.Google Scholar
Smith, C.A. 1955. Early 19th century records of the Clanwilliam cedar (Widdringtonia juniperoides Endl.). Journal of the South African Forestry Association 25: 18.CrossRefGoogle Scholar
Smith, C.C. 1970. The coevolution of Pine squirrels (Tamiasciurus) and conifers. Ecological Monographs 40: 349371.CrossRefGoogle Scholar
Smith, D.L. 1964. The evolution of the ovule. Biological Reviews 39: 137159.CrossRefGoogle Scholar
Smith, M.J.B. 1982. Origins of the tropical alpine flora. Pp 283304 in Gressitt, J.L. (ed.), Biogeography and Ecology of New Guinea. The Hague: W. Junk.Google Scholar
Smith, S.E. & Read, D.J. 1997. Mycorrhizal Symbiosis, 2nd edn. London: Academic Press.Google Scholar
Smith, S.Y. & Stockey, R.A. 2002. Permineralized pine cones from the Cretaceous of Vancouver Island, British Columbia. International Journal of Plant Sciences 163: 185196.CrossRefGoogle Scholar
Smith, T.M. & Huston, M.A. 1989. A theory of the spatial and temporal dynamics of plant communities. Vegetatio 83: 4969.CrossRefGoogle Scholar
Smith-White, S. 1959. Cytological evolution in the Australian flora. Cold Spring Harbor Symposium in Quantative Biology 24: 273289.CrossRefGoogle ScholarPubMed
Smouse, P.E. & Saylor, L.C. 1973. Studies in the Pinus rigidaserotina complex. II. Natural hybridization among the Pinus rigidaserotina complex, P. taeda, and P. echinata. Annals of the Missouri Botanical Garden 60: 192203.CrossRefGoogle Scholar
Snook, L.C. 1993. Conservation of the Monarch butterfly reserves in Mexico: focus on the forest. Pp 363375 in Malcolm, S.B. & Zalucki, M.P. (eds.), Biology and Conservation of the Monarch Butterfly. Los Angeles, CA: Los Angeles Natural History Museum.Google Scholar
Soejarto, D.D. & Farnsworth, N.R. 1989. Tropical rain forests: potential sources of new drugs? Perspectives in Biology and Medicine 32: 244256.CrossRefGoogle ScholarPubMed
Solins, P., Grier, C.C., McCorison, F.M., Cromack, K. & Fogel, R. 1980. The internal element cycles of an old-growth Douglas-fir ecosystem in western Oregon. Ecological Monographs 50: 261285.CrossRefGoogle Scholar
Sonett, C.P. & Suess, H.E. 1984. Correlation of bristlecone pine ring widths with atmospheric C14 variations: a climate–sun relation (Campito Mountain, California). Nature 307: 141143.CrossRefGoogle Scholar
Sorensen-Corhern, K.A., Ford, E.D. & Sprugel, D.G. 1993. A model of a competition incorporating plasticity through modular foliage and crown development. Ecological Monographs 63: 277304.CrossRefGoogle Scholar
Soto-Nuñez, J.C. & Vázquez-García, L. 1993. Vegetation types of Monarch butterfly overwintering habitat in Mexico. Pp 287293 in Malcolm, S.B. & Zalucki, M.P. (eds.), Biology and Conservation of the Monarch Butterfly. Los Angeles, CA: Los Angeles Natural History Museum.Google Scholar
Soule, M.E. (ed.). 1986. Conservation Biology: The Science of Scarcity and Diversity. Sunderland, MA: Sinaver Associates.Google Scholar
Soule, M.E. (ed.). 1987. Viable Populations for Conservation. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Sousa, N., Ramos, M., Franco, A., Oliveira, R. & Castro, P. 2012. Mycorrhizal symbiosis affected by different genotypes of Pinus pinaster. Plant and Soil 359: 245253.CrossRefGoogle Scholar
Specht, R.L. 1958. The climate, geology, soils and plant ecology of the northern portion of Arnhem Land. Pp 333414 in Specht, R.L. & Mountford, C.P. (eds.), Records of the American–Australian Scientific Expeditions to Arnhem Land, Vol. 3. Botany and Plant Ecology. Melbourne: Melbourne University Press.Google Scholar
Specht, R.L. 1958. The geographical relationships of the flora of Arnhem Land. Pp 415478 in Specht, R.L. & Mountford, C.P. (eds.), Records of the American-Australian Scientific Expeditions to Arnhem Land. Vol. 3. Botany and Plant Ecology. Melbourne: Melbourne University Press.Google Scholar
Specht, R.L. 1969. A comparison of the sclerophyllous vegetation characteristic of Mediterranean-type climates in France, California and southern Australia. II. Dry matter, energy and nutrient accumulation. Australian Journal of Botany 17: 298308.Google Scholar
Specht, R.L. 1970. Vegetation. Pp 4767 in Leeper, G.W. (ed.). The Australian Environment. Melbourne: CSIRO and Melbourne University Press.Google Scholar
Specht, R.L. (ed.). 1978. Ecosystems of the World: Heathlands and Related Shrublands. Amsterdam: Elsevier.Google Scholar
Spegazzini, C. 1924. Coniferales fossiles Patagonicas. Anales de la Sociedad Cientitica Argentina 98: 125139.Google Scholar
Sperry, J.S., Hacke, U.G. & Pittermann, J. 2006. Size and function of conifer tracheids and angiosperm vessels. American Journal of Botany 93: 14901500.CrossRefGoogle ScholarPubMed
Spicer, R.A. 1989. Plants at the Cretaceous–Tertiary boundary. Philosophical Transactions of the Royal Society of London B 325: 291305.Google Scholar
Spicer, R.A. & Herman, A.B. 2001. The Albian–Cenomanian flora of the Kukpowruk River, western North Slope, Alaska: stratigraphy, palaeofloristics, and plant communities. Cretaceous Research 22: 140.CrossRefGoogle Scholar
Spies, T.A. & Franklin, J.F. 1989. Gap characteristics and vegetation response in coniferous forests of the Pacific Northwest. Ecology 70: 543545.CrossRefGoogle Scholar
Sporne, K.R. 1965. The Morphology of Gymnosperms. London: Hutchinson.Google Scholar
Sporne, K.R. 1967. The Morphology of Gymnosperms, 3rd edn. London: Hutchinson University Library.Google Scholar
Sprugel, D.G. 1991. Disturbance, equilibrium, and environmental variability: what is ‘natural’ vegetation in a changing environment? Biological Conservation 58: 118.CrossRefGoogle Scholar
Squilace, A.E. & Wells, O.O. 1981. Geographic variation of monoterpenes in cortical oleoresin of Loblolly pine. Silvae Genetica 30: 127135.Google Scholar
Srikiatis, L. 2014. The De Geer, Trulian and Beringia routes: key concepts for understanding early Cenozoic biogeography. Journal of Biogeography 41: 136154.Google Scholar
Srodon, A. 1978. History of yew in Poland. Pp 514 in Bartkowiak, S., Bialobok, C., Bugala, W., et al. (eds.), The Yew: Taxus baccata L. Springfield, VA: US Department of Commerce, National Technical Information Service.Google Scholar
Stace, C.A. 2005. Plant taxonomy and biosystematics: does DNA provide all the answers? Taxon 54: 9991007.CrossRefGoogle Scholar
Stace, C.A. 2010. Classification by molecules: what’s in it for field botanists? Watsonia 28: 103122.Google Scholar
Stanislavsky, F.A. 1973. A new genus Toretzia from the Upper Triassic of the Donetz Basin, and its relation to the genera of the order Ginkgoales. Palaeontologiczesky Zhurnal 1: 8896.Google Scholar
Stanley, T.D. & Ross, E.M. 1989. Flora of South-Eastern Queensland, Vol 4. Brisbane: Queensland Department of Primary Industries.Google Scholar
Stapf, O. 1922. Picea brachytyla. Botanical Magazine 148.Google Scholar
Staplin, F.L., Pocock, S.J. & Jansonius, J. 1967. Relationships amongst gymnospermous pollen. Review of Palaeobotany and Palynology 3: 297310.CrossRefGoogle Scholar
Stark, N. 1968. The environmental tolerance of the seedling stage of Sequoiadendron giganteum. American Midland Naturalist 80: 8495.CrossRefGoogle Scholar
Starr, A.M. 1910. The microsporophylls of Ginkgo. Botanical Gazette 49: 5155.Google Scholar
Stearn, W.T. 1948. Diiscovery of a living Metasequoia in China. Nature (London) 161: 594.Google Scholar
Stearn, W.T. 1959. The background of Linnaeus’s contributions to the methods and nomenclature of systematic biology. Systematic Zoology 7: 422.CrossRefGoogle Scholar
Stearn, W.T. 1973. Botanical Latin. Newton Abbot: David & Charles.Google Scholar
Stebbins, G.L. 1958. Longevity, habitat and release of genetic variability in higher plants. Cold Spring Harbour Symposium, Quantitative Biology 23: 365378.CrossRefGoogle Scholar
Stebbins, G.L. 1971. Chromosomal Evolution in Higher Plants. London: Edward Arnold.Google Scholar
Stebbins, G.L. 1974. Flowering Plants: Evolution Above the Species Level. Cambridge, MA: Harvard University Press.CrossRefGoogle Scholar
Stefenon, V.M., Galling, O. & Fineldey, R. 2006. Phylogenetic relationships within genus Araucaria (Araucariaceae) assessed by means of Azp fingerprints. Silvae Genetica 55: 4552.CrossRefGoogle Scholar
Stein, N. 1978. Coniferen im westlichen malayischen Archipel. Biogeographica 11: 1168.Google Scholar
Stein, W.E. 1987. Phylogenetic analysis of fossil plants. Review of Palaeobotany and Palynology 50: 3161.CrossRefGoogle Scholar
Stein, W.E., Wight, D. & Beck, C. 1984. Possible alternatives for the origin of the Sphenopsids. Systematic Botany 9: 102118.Google Scholar
Steinhoff, R.J., Joyce, D.G. & Fins, L. 1983. Isozyme variation in Pinus monticola. Canadian Journal of Forest Research 13: 11221132.CrossRefGoogle Scholar
Stephens, S.L. & Libby, W.J. 2006. Anthropogenic fire and bark thickness in coastal and island pine populations from Alta and Baja California. Journal of Biogeography 3: 648652.CrossRefGoogle Scholar
Stephens, S.L., Skinner, C.N. & Gill, S.J. 2003. Dendrochronology-based fire history of Jeffreyi pine: mixed conifer forests in the Sierra San Pedro Martin, Mexico. Canadian Journal of Forest Research 33: 10901101.CrossRefGoogle Scholar
Stephens, S.L., Fry, D.L., Franco-Vizcaino, E., Collins, B.M. & Moghaddas, J.M. 2007. Coarse woody debris and canopy cover in an old-growth Jeffrey pine – mixed conifer forest from the Sierra San Pedro Martir, Mexico. Forest Ecology and Management 240: 8795.CrossRefGoogle Scholar
Stephenson, N.L. 1999. Reference conditions for giant sequoia forest restoration, structure, process, and precision. Ecological Applications 9: 12501263.CrossRefGoogle Scholar
Stephenson, N.L. & Demetry, A. 1995. Estimating ages of giant sequoias. Canadian Journal of Forest Research 25: 223233.CrossRefGoogle Scholar
Sterling, C. 1949. Some features in the morphology of Metasequoia. American Journal of Botany 36: 461471.CrossRefGoogle Scholar
Stern, M.J. 1995. Vegetation recovery on earthquake-triggered landslide sites in the Equadorian Andes. Pp 207220 in Churchill, S.P., Balslev, H., Forero, E. & Luteyn, J.L. (eds.), Biodiversity and Conservation of Neotropical Montane Forests. New York: New York Botanic Garden.Google Scholar
Steven, H.M. & Carlisle, A. 1959. The Native Pinewoods of Scotland. Edinburgh: Oliver & Boyd.Google Scholar
Stevens, K.A. & Parrish, J.M. 2005. Digital reconstructions of sauropod dinosaurs and implications for feeding. Pp 178200 in Curry, K.A. & Wilson, J.A. (eds.), The Sauropods: Evolution and Paleobiology. Berkeley, CA: University of California Press.Google Scholar
Stevens, U.A. & Parrish, J.M. 1999. Neck posture and feeding habits of two Jurassic sauropod dinosaurs inferred from extant animals. Science 284: 798800.CrossRefGoogle Scholar
Stiles, W. 1908. The anatomy of Saxegothaea conspicua. New Phytologist 6: 209223.CrossRefGoogle Scholar
Stiles, W. 1912. The Podocarpaceae. Annals of Botany 21: 443514.CrossRefGoogle Scholar
Stockey, R.A. 1975. Seeds and embryos of Araucaria mirabilis. American Journal of Botany 62: 856868.CrossRefGoogle Scholar
Stockey, R.A. 1977. Reproductive biology of the Cerro Cuadrado (Jurassic) fossil conifers: Pararaucaria patagonica Weiland. American Journal of Botany 64: 733744.CrossRefGoogle Scholar
Stockey, R.A. 1978. Reproductive biology of the Cerro Cuadrado fossil conifers: ontogeny and reproductive strategies in Araucaria mirabilis (Speazzini) Windhausen. Palaeontographica 166: 115.Google Scholar
Stockey, R.A. 1980. Anatomy and morphology of Araucaria sphaerocarpa Carruthers from the Jurassic Inferior Oolite of Bruton, Somerset. Botanical Gazette 141: 116124.CrossRefGoogle Scholar
Stockey, R.A. 1983. Pinus driftwoodensis sp. n. from the early Tertiary of British Columbia. Botanical Gazette 144: 148156.CrossRefGoogle Scholar
Stockey, R.A. 1989. Antarctic and Gondwanan conifers. Pp 179191 in Taylor, T.N. & Taylor, E.L. (eds.), Antarctic Paleobiology, Its Role in the Reconstruction of Gondwana. New York: Springer.Google Scholar
Stockey, R.A. & Frevel, B.J. 1997. Cuticle micromorphology of Prumnopitys philippi (Podocarpaceae). International Journal of Plant Sciences 158: 198221.CrossRefGoogle Scholar
Stockey, R.A. & Ko, H. 1988. Cuticle micromorphology of some New Caledonian podocarps. Botanical Gazette 149: 240252.CrossRefGoogle Scholar
Stockey, R.A. & Nishida, M. 1986. Pinus harborensis sp. nov. and the affinities of permineralised leaves from the Upper Cretaceous of Japan. Canadian Journal of Botany 64: 18561866.CrossRefGoogle Scholar
Stockey, R.A. & Taylor, T.N. 1978. Scanning electron microscopy of epidermal patterns and cuticular structure in genus Araucaria. Scanning Electron Microscopy 2: 223228.Google Scholar
Stockey, R.A. & Taylor, T.N. 1978. Cuticular features and epidermal patterns in the genus Araucaria de Jussieu. Botanical Gazette 139: 490498.CrossRefGoogle Scholar
Stockey, R.A., Ko, H. & Woltz, P. 1992. Cuticle micromorphology of Falcatifolium de Laubenfels (Podoacarpaceae). International Journal of Plant Sciences 153: 589601.CrossRefGoogle Scholar
Stockey, R.A., Frevel, B.J. & Woltz, P. 1998. Cuticle micromorphology of Podocarpus subgenus Podocarpus section Scytopodium (Podocarpaceae) of Madagascar and South Africa. International Journal of Plant Science 159: 923940.CrossRefGoogle Scholar
Stockey, R.A., Graham, S.W. & Crane, P.R. 2009. Introduction to the Darwin special issue: the abominable mystery. American Journal of Botany 96: 34.CrossRefGoogle Scholar
Stockler, K., Daniel, I.L. & Lockhart, P.J. 2002. New Zealand Kauri (Agathis australis ((D.Don) Lindl., Araucariaceae) survives Oligocene drowning. Systematic Biology 51: 827832.CrossRefGoogle ScholarPubMed
Stoffberg, E. 1991. Morphological and ontogenetic studies on southern African podocarps: initiation of the seed scale complex and early development of the integument, nucellus and epimatium. Botanical Journal of the Linnean Society 105: 2135.CrossRefGoogle Scholar
Stohlgren, T.J. 1988. Litter dynamics in two Sierran mixed conifer forests. 1. Litter fall and decomposition rates. Canadian Journal of Forest Research 18: 11271135.CrossRefGoogle Scholar
Stokes, W.L. 1964. Fossilised stomach contents of a sauropod dinosaur. Science 143: 576577.CrossRefGoogle Scholar
Stone, J.O., Ballantyne, C.K. & Fifield, L.K. 1998. Exposure dating and validation of periglacial weathering limits, northwest Scotland. Geology 26: 587590.2.3.CO;2>CrossRefGoogle Scholar
Stoutjesdijk, P. & Barkman, J.J. 1992. Microclimate, Vegetation and Fauna. Uppsala: Opulus Press.Google Scholar
Strauss, S.H. & Ledig, F.T. 1985. Seedling architecture and life history evolution in pines. American Naturalist 125: 702715.CrossRefGoogle Scholar
Strobel, G.A. & Hess, W.M. 1996. A scanning electron microscopy study of Taxus leaves as related to taxonomy. Scanning Microscopy 10: 11111126.Google ScholarPubMed
Strullu-Derrien, C., Kenrick, C., Pressel, P., et al. 2014. Fungal associations in Horneophyton ligneri from the Rhynie Cjert (c. 407 million years old) closely resemble those in extant lower land plants: novel insights into ancestral plant–fungus symbioses. New Phytologist 203: 964979.CrossRefGoogle Scholar
Stubblefield, S.P., Taylor, T.N. & Trappe, J.M. 1987. Vesicular–arbuscular mycorrhizae from the Triassic of Antarctica. American Journal of Botany 74: 19041911.CrossRefGoogle Scholar
Stultz, D.Z., Axsmith, B.J., Knight, T.K. & Bingham, P.S. 2012. The conifer Araucaria bladenensis and associated large pollen and ovulate cones from the Upper Cretaceous Ingersol shale (Eutaw Formation) of Alabama. Cretaceous Research 34: 142148.CrossRefGoogle Scholar
Stützel, T. & Rowekamp, I. 1999. Female reproductive structures in taxales. Flora 194: 145157.CrossRefGoogle Scholar
Styles, B.T. 1993. Genus Pinus: a Mexican purview. Pp 397420 in Ramamoorthy, T.P. (ed.), Biological Diversity of Mexico: Origins and Distributions. Oxford: Oxford University Press.Google Scholar
Styles, B.T. & Hughes, C.E. 1983. Studies of variation in Central American pines III. Brenesia 21: 261291.Google Scholar
Sugihara, Y. 1938. The embryogeny of Cupressus funebris Endlicher. Botanical Magazine Tokyo 69: 439441.CrossRefGoogle Scholar
Sugihara, Y. 1941. The embryogeny of Cunninghamia lanceolata Hooker. Science Reports Tohoku University (Biology) 16: 187192.Google Scholar
Sugihara, Y. 1943. Embryological observations on Keteleeria davidiana Beissner var. formosana Hayata. Science Reports Tohoku University (Biology) 17: 215222.Google Scholar
Sugihara, Y. 1943. Notes on Amentotaxus. Botanical Magazine Tokyo 57: 404405.Google Scholar
Sugihara, Y. 1985. Further development of the embryo of Cupressus sempervirens. Journal of Japanese Botany 65: 353358.Google Scholar
Sugihara, Y. 1990. Embryological studies on two Mexican species of Cupressus. Journal of Japanese Botany 65: 353358.Google Scholar
Sugihara, Y. 1991. On male gametophytes in Cupressus sempervirens L. and Cupressus funebris Endl. Journal of Japanese Botany 66: 224228.Google Scholar
Sugimota, N., Kuroyanagi, M., Kato, T., et al. 2006. Identification of the main constituents in sandarac resin, a natural gum base. Journal of the Food Hygiene Society of Japan 47: 7679.CrossRefGoogle Scholar
Summers, R.W. 2004. Use of pine snags by birds in different stand types of Scots pine, Pinus sylvestris. Bird Study 51: 212221.CrossRefGoogle Scholar
Sun, C., Dilcher, D.L., Wang, H., Sun, G. & Ge, Y. 2008. A study of Ginkgo leaves from the Middle Jurassic of Inner Mongolia, China. International Journal of Plant Science 169: 11281139.CrossRefGoogle Scholar
Sun, T.X. 2008. Cuticle micromorphology of Nageia. Journal of Wuhan Botanical Research 26: 554560.Google Scholar
Suszka, B. 1985. Conditions for after-ripening and germination of seeds and for seedling-emergence of the English yew (Taxus baccata L.). Arboretum Kornickie 30: 285338.Google Scholar
Sutherland, R., Ppasojevic, S. & Gurnis, M. 2010. Mantle upwelling after Gondwana subduction death explains anomalous topography and subsidence histories of eastern New Zealand and West Antarctica. Geology 38: 155158.CrossRefGoogle Scholar
Suzuki, M. 1979. The course of resin canals in the shoots of conifers. 2 Araucariaceae, Cupressaceae and Taxodiaceae. Botanical Magazine Tokyo 92: 253274.CrossRefGoogle Scholar
Suzuki, M. 1979. The course of resin canals in the shoots of conifers. 3. Pinaceae and summary. Botanical Magazine Tokyo 92: 333353.CrossRefGoogle Scholar
Svenning, J.-C. & Magard, E. 1999. Population ecology and conservation status of the last natural population of English yew Taxus baccata in Denmark. Biological Conservation 88: 173182.CrossRefGoogle Scholar
Svenning, J.-C., Kinner, D.A., Stallard, R.F., Englebrecht, B.M.J. & Wright, S.J. 2004. Ecological determinism in plant community structure across a tropical forest landscape. Ecology 85: 25262538.CrossRefGoogle Scholar
Swanson, C.P. 1960. Cytology and Cytogenetics. London: Macmillan.Google Scholar
Swenson, U. & Hill, R.S. 2001. Most parsimonious areagram versus fossils: the case of Nothofagus (Nothofagaceae). Australian Journal of Botany 49: 367376.CrossRefGoogle Scholar
Swenson, U., Hill, R.S. & McLouighlin, S. 2001. Biogeography of Nothofagus supports the sequence of Gondwana break-up. Taxon 50: 10251041.CrossRefGoogle Scholar
Swetnam, T.W. & Betancourt, J.L. 1990. Fire-southern oscillation relations in the southwestern United States. Science 249: 10171020.CrossRefGoogle ScholarPubMed
Syrach-Larsen, C. & Westergaard, M. 1938. A triploid hybrid between Larix decidua Mill and L. occidentalis Nutt. Journal of Genetics 36: 523530.Google Scholar
Syring, J., Wilyard, A., Cron, R. & Liston, A. 2005. Evolutionary relationships among Pinus (Pinaceae) subsections inferred from multiple low-copy nuclear loci. American Journal of Botany 92: 20862100.CrossRefGoogle ScholarPubMed
Systematics Association Committee for Descriptive Terminology 1962. Terminology of simple symmetrical plane shapes. Taxon 11: 145156, 245–247.CrossRefGoogle Scholar
Szafer, W. 1949. Studies on the genus Tsuga Carr. in the Tertiary of Europe. Bulletin Internationaale de l’Academie Polonaise des Sciences et des Lettres et de Mathematic et Sciences Naturelles B 3: 2351Google Scholar
Sze, H.-C. 1951. On Sequoia remains from Fushun, South Manchuria. Science Record 4: 185191 (in English with Chinese abstract).Google Scholar
Szeicz, J.M. 1997. Growth trends and climatic sensitivity of trees in the North Patagonian rain forest of Chile. Canadian Journal of Forest Research 27: 10031014.CrossRefGoogle Scholar
Szi, X.-J. 1951. Structure of cuticle and small stomata of modern Metasequoia leaves. Science China 2: 239243 (in Chinese).Google Scholar
Tabarelli, M., Mantovani, W. & Peres, C.A. 1999. Effects of habitat fragmentation on plant guildstructure in the montane Atlantic forest of southeastern Brazil. Biological Conservation 91: 119127.CrossRefGoogle Scholar
Taberlet, P., Fumigalli, L., Wust-Saucy, A.G. & Cosson, J.F. 1998. Comparative phylogeography and postglacial colonisation routes in Europe. Molecular Ecology 7: 453464.CrossRefGoogle ScholarPubMed
Tai, A. & Ueno, J. 1965. On the fossil pollen grains of Pseudolarix-type. Acta Phytotaxonomica Geobotanica 21: 141143.Google Scholar
Takahara, H., Tanida, K. & Miyoshi, N. 2001. The full-glacial refuge of Cryptomeria japonica in the Oki Islands, western Japan. Japanese Journal of Palynology 47: 2133.Google Scholar
Takahashi, H.A., Yonenobu, H., Kakamura, T. & Wada, H. 2001. Seasonal fluctuation of stable carbon isotopic composition in Japanese cypress rings from the last glacial period: possibility of paleoenvironmental reconstruction. Radiocarbon 43: 433438.CrossRefGoogle Scholar
Takahashii, K. 1977. Upper Cretaceous palynoflora from Quiriquina Island, Chile. Bulletin of the Faculty of Liberal Arts, Nagasaki University, Natural Science 17: 2953.Google Scholar
Takasi, S., Bottomley, H., Andreller, I., et al. 2009. Infrared radiation from hot cones on cool conifers attracts seed-feeding insects. Proceedings of the Royal Society B 276: 649655.Google Scholar
Takaso, T. & Owens, J.N. 1996. Ovulate cone, pollination drop, and pollen capture in Sequoiadendron. American Journal of Botany 83: 11751180.CrossRefGoogle Scholar
Takhtajan, A.L. 1956. Higher Plants. I. From Psilophytes to Coniferophytes. Moscow: Akademica Nauk SSSR (in Russian).Google Scholar
Takhtajan, A.L. 1967. Systema et Phylogenia Magnoliophytorum. Moscow: Nauka.Google Scholar
Takhtajan, A.L. 1986. Problems of Palaeobotany. Leningrad: Nauka (in Russian).Google Scholar
Tapias, R., Gil, L., Fuentes-Utrilla, P. & Pardos, J.A. 2001. Canopy seed banks in Mediterranean pines of southeastern Spain: a comparison between Pinus halepensis Mill., P. pinaster Ait. and P. pinea L. Journal of Ecology 89: 629638.CrossRefGoogle Scholar
Tapias, R., Climent, J., Pardos, J.A. & Gil, L. 2004. Life histories of Mediterranean pines. Plant Ecology 171: 5368.CrossRefGoogle Scholar
Taylor, A.H. 1990. Disturbance and persistence of Sitka spruce (Picea sitchensis (Bong.) Carr) in coastal forests of the Pacific Northwest, North America. Journal of Biogeography 17: 4758.CrossRefGoogle Scholar
Taylor, A.H. 1993. Fire history and structure of red fir (Abies magnifica) forests, Swain Mountain Experimental Forest, northeastern California. Canadian Journal of Forest Research 23: 16721678.CrossRefGoogle Scholar
Taylor, A.H. 2000. Fire regimes and forest changes in the mid and upper montane forests of the southern Cascades, Lassen National park, California, USA. Journal of Biogeography 27(87):104.CrossRefGoogle Scholar
Taylor, A.H. & Qin, Z. 1988. Regeneration patterns in old-growth AbiesBetula forests in the Wolong Natural Reserve, Sichuan, China. Journal of Ecology 76: 12041218.CrossRefGoogle Scholar
Taylor, A.H. & Skinner, C.N. 1998. Fire history and landscape dynamics in a late-successional reserve, Klamath Mountains, California, USA. Forest Ecology and Management 44: 117.Google Scholar
Taylor, A.H. & Zisheng, Q. 1989. Structure and composition of selectively cut and uncut AbiesTsuga forest in Wolong natural reserve and implications for panda conservation in China. Biological Conservation 47: 83108.CrossRefGoogle Scholar
Taylor, D.W. & Hickey, L.J. Phylogenetic evidence for the herbaceous origin of angiosperms. Plant Systematics and Evolution 1180: 137156.Google Scholar
Taylor, S.F. & Brodribb, T.M. 2005. A unique mode of parasitism in the conifer coral tree Parasitaxus ustus (Podocarpaceae). Plant, Cell and Environment 28: 13161325.Google Scholar
Taylor, S.R. & McClennan, S.M. The geochemical evolution of the continental crust. Review of Geophysica 111: 241265.Google Scholar
Taylor, T.N. 1981. Paleobotany: An Introduction to Fossil Plant Biology. New York: McGraw-Hill.Google Scholar
Taylor, T.N. & Krings, M. 2010. Paleomycology: the rediscovery of the obvious. Palaios 25: 283286.CrossRefGoogle Scholar
Taylor, T.N. & Millay, M.A. 1979. Pollination biology and reproduction in early seed plants. Review of Palaeobotany and Palynology 27: 329355.CrossRefGoogle Scholar
Taylor, T.N. & Taylor, E.L. 1993. The Biology and Evolution of Fossil Plants. Engelwood Cliffs, NJ: Prentice Hall.Google Scholar
Taylor, T.N., Remy, W., Hass, H. & Kerp, H. 1995. Fossil arbuscular mycorrhizae from the Early Devonian. Mycologia 87: 560573.CrossRefGoogle Scholar
Tegner, J. 1967. Anatomy and taxonomy in the Podocarpaceae. Botaniska Notiser 120: 504506.Google Scholar
Templer, P.H. & Weathers, K.C. 2011. Use of mixed ion exchange resin and the denitrifier method to determine isotopic values of nitrate in atmospheric deposition and canopy throughfall. Atmospheric Environment 45: 20172020.CrossRefGoogle Scholar
Templer, P.H., Weathers, K.C., Ewing, H.A., et al. 2015. Fog as a source of nitrogen for redwood trees: evidence from fluxes and stable isotopes. Journal of Ecology 103: 13671407.CrossRefGoogle Scholar
Tendersoo, L., Suvi, T., Jairus, T. & Kõljalg, U, 2008. Forest microsite effects on community composition of ectomycorrhizal fungi on seedlings of Picea abies and Betula pendula. Environmental Microbiology 10: 11891201.CrossRefGoogle Scholar
Teoh, S.B. & Rees, H. 1976. Nuclear DNA amounts in populations of Picea and Pinus species. Heredity 36: 123127.CrossRefGoogle Scholar
Terasmaa, T. 1971. Karyotype analysis of Norway Spruce, Picea abies (L.) Karst. Silvae Genetica 20: 179182.Google Scholar
Terborgh, J. 1985. The vertical component of plant species diversity in temperate and tropical forests. American Naturalist 126: 760766.CrossRefGoogle Scholar
Terry, R.G., Nowak, R.S. & Tausch, R.J. 2000. Genetic variation in chloroplast and nuclear ribosomal DNA in Utah juniper, Juniperus osteosperma, Cupressaceae: evidence for interspecific gene flow. American Journal of Botany 87: 250258.CrossRefGoogle ScholarPubMed
Teslenko, Y.V., Golbert, A.V. & Poliakova, I.F. 1966. The routes of dispersal of the most ancient angiosperms in Western Siberia. Botanische Zhurnal 56: 801803.Google Scholar
Thomas, H.H. 1913. On some new rare Jurassic plants from Yorkshire: Eretmophyllum, a new type of ginkgoalean leaf. Proceedings of the Cambridge Philosophical Society 17: 256262.Google Scholar
Thomas, P., Sengdala, K. Lamxay, V. & Khou, E. 2007. New records for conifers in Cambodia and Laos. Edinburgh Journal of Botany 64: 3744.CrossRefGoogle Scholar
Thomas, P.A. & Wein, R.W. 1994. Amelioration of wood ash toxicity and jack pine establishment. Canadian Journal of Forest Research 24: 748755.CrossRefGoogle Scholar
Thompson, W.P. 1912. The structure of the stomata of certain Cretaceous conifers. Botanical Gazette 54: 6367.Google Scholar
Thomson, R.B. 1945. ‘Polyembryony’ sexual and asexual embryo initiation. Transactions of the Royal Society of Canada 39: 143169.Google Scholar
Tidwell, W.D. & Medlyn, D.A. 1992. Short shoots from the Upper Jurassic Morrison Formation, Utah, Wyoming and Colorado, USA. Review of Palaeobotany and Palynology 71: 219238.CrossRefGoogle Scholar
Tiffney, B.H. 1984. Seed size, dispersal syndromes, and the rise of angiosperms: evidence and hypothesis. Annals of the Missouri Botanical Garden 71: 551576.CrossRefGoogle Scholar
Tiffney, B.H. 1985. The Eocene North Atlantic land-bridge: its importance in tertiary and modern phytogeography of the Northern Hemisphere. Journal of the Arnold Arboretum 66: 243273.CrossRefGoogle Scholar
Tiffney, B.H. 2004. Vertebrate dispersal of seed plants through time. Annual Review of Ecology, Evolution and Systematics 35: 129.CrossRefGoogle Scholar
Tiffney, B.H. 2008. Phylogeography, fossils, and northern hemisphere biogeography: the role of physiological uniformatarianism. Annals of the Missouri Botanical Garden 95: 135143.CrossRefGoogle Scholar
Tilman, D. 1988. Plant Strategies and the Structure and Dynamics of Plant Communities. Princeton, NJ: Princeton University Press.Google Scholar
Tinker, D.B., Romme, W.H., Hargrove, W.W. Gardner, R.H. & Turner, M.G. 1994. Landscape scale heterogeneity in lodgepole pine serotiny. Canadian Journal of Forest Research 24: 897903.CrossRefGoogle Scholar
Titov, E.V. 1977. Trial in crossing Pinus sibirica with other pines in the NE Altai. Lesovedenie 4: 8187 (in Russian).Google Scholar
Tomback, D.F., Hoffmann, L.A. & Sund, S.K. 1990. Coevolution of whitebark pine and nutcrackers: implications for forest regeneration. General Technical Report, US Department of Agriculture, Forest Service INT-270.Google Scholar
Tomlinson, P.B. 1992. Aspects of cone morphology and development in Podocarpaceae (Coniferales). International Journal of Plant Science 153: 572588.CrossRefGoogle Scholar
Tomlinson, P.B. & Takaso, T. 2002. Seed cone structure in conifers in relation to development and pollination: a biological approach. Canadian Journal of Botany 80: 12501273.CrossRefGoogle Scholar
Tomlinson, P.B. & Zimmermann, M.H. (eds.). 1978. Tropical Trees as Living Systems. Cambridge: Cambridge University Press.Google Scholar
Tomlinson, P.B., Takaso, T. & Rattenbury, J.A. 1989. Cone and ovule ontogeny in Phyllocladus (Podocarpaceae). Botanical Journal of the Linnean Society 99: 209221.CrossRefGoogle Scholar
Tomlinson, P.B., Braggins, J.E. & Rattenbury, J.A. 1991. Pollination drop in relation to cone morphology in Podocarpaceae: a novel reproductive mechanism. American Journal of Botany 78: 12891303.CrossRefGoogle Scholar
Tomlinson, P.B., Takaso, T. & Cameron, E.K. 1993. Cone development in Libocedrus (Cupressaceae): phenological and morphological aspects. American Journal of Botany 80: 649659.Google Scholar
Tomlinson, P.B., Braggins, J.E. & Rattenbury, J.A. 1997. Contrasted pollen capture mechanisms in Phyllocladaceae and certain Podocarpaceae (Coniferales). American Journal of Botany 84: 214223.CrossRefGoogle ScholarPubMed
Torres, O., Calvo, L. & Valbuena, L. 2006. Influence of high temperatures on seed germination of a special Pinus pinaster stand adapted to frequent fires. Plant Ecology 186: 129136.CrossRefGoogle Scholar
Tortorelli, L.A. 1956. Maderas y Bosques en la Argentina. Buenos Aires: Editorial Acme.Google Scholar
Toumi, F.B., Benyahia, M., Hamel, L., Mohamedi, H. & Boudaghen, L. 2011. Comparative study of chemical composition of essential oils in Tetraclinis articulata (Vahl) Masters originating from Algeria. Acta Botanica Gallica 158: 93100.CrossRefGoogle Scholar
Townrow, J.A. 1965. Notes on some Tasmanian pines. I. Some lower Tertiary podocarps. Papers and Proceedings of the Royal Society of Tasmania 99: 87108.CrossRefGoogle Scholar
Townrow, J.A. 1965. Notes on some Tasmanian pines II. Athrotaxis from the Lower Tertiary. Papers and Proceedings of the Royal Society of Tasmania 99: 109113.CrossRefGoogle Scholar
Townrow, J.A. 1966. Fossil plants from Alon and Carapace nunataks, and from Upper Mill and Shakleton glaciers, Antarctica. New Zealand Journal of Geology and Geophysics 10: 456473.CrossRefGoogle Scholar
Townrow, J.A. 1967. On Rissikia & Mataia, podocarpaceous conifers from the Lower Mesozoic of southern lands. Papers and Proceedings of the Royal Society of Tasmania 101: 103136.CrossRefGoogle Scholar
Townrow, J.A. 1967 The Brachyphyllum crassum complex of fossil conifers. Papers and Proceedings of the Royal Society of Tasmania 101: 149172.CrossRefGoogle Scholar
Troitsky, A.V., Melekhovets, Y.F., Rakhimova, G.M., et al. 1991. Angiosperm origin and early stages of seed plant evolution deduced from rRNA sequence comparisons. Journal of Molecular Evolution 32: 253261.CrossRefGoogle ScholarPubMed
Troncoso, A., Gnedinger, S. & Herbst, A. 2000. Hediphyllum, Risskia and Kesmophyllum (Pinophyta, Coniferales) en el Triásico del norte chico de Chile y sur de Argentina. Ameghiniana 37: 119125.Google Scholar
Truswell, E.M. 1993. Vegetation changes in the Australian tertiary in response to climatic and physiographic forcing factors. Australian Systematic Botany 6: 533558.CrossRefGoogle Scholar
Truswell, E.M. & Macphail, M.K. 2004. Carnivorous plants at high latitudes: pollen evidence for Droseraceae growing in East Antarctica during the Late Eocene. Association of Australasian Palaeontologists Memoirs 29: 8597.Google Scholar
Truswell, E.M., Kershaw, A.P. & Sluiter, I.R. 1987. The Australia–south-east Asian connection: evidence from the palaeobotanic record. Pp 3249 in Whitomer, T.C. (ed.), Biogeographical Evolution of the Malay Archipelago. Oxford: Oxford University Press.Google Scholar
Tryon, A.F. & Vida, G. Platyzoma: a new look at an old link in ferns. Science 156: 11091110.CrossRefGoogle Scholar
Tryon, R. 1960. A glossary of some terms relating to the fern leaf. Taxon 9: 104109.CrossRefGoogle Scholar
Tucker, N.I.J. 2000. Linkage restoration: interpreting fragmentation theory for the design of a rainforest linkage in the humid wet tropics of north-eastern Queensland. Ecological Management Restoration 1: 3541.CrossRefGoogle Scholar
Tuetken, T., Hummmel, J. & Sander, M. 2006. The diet of sauropod dinosaurs: carbon isotope compositions of fossil bones and potential food plants. Journal of Vertebrate Paleontology 26.Google Scholar
Turner, C.E. & Peterson, F. 2004. Reconstruction of Upper Jurassic Morrison Formation extinct ecosystems: a synthesis. Sedimentary Geology 167: 309355.CrossRefGoogle Scholar
Turner, D.P. 1985. Successional relationships and comparison of biological characteristics between six north western conifers. Bulletin of the Torrey Botanical Club 112: 421428.CrossRefGoogle Scholar
Turner, M.G., Turner, D.M., Romme, W.H. & Tinker, D.B. 2007. Cone production in young post-fire Pinus contorta stands in Greater Yellowstone (USA). Forest Ecology and Management 242: 119126.CrossRefGoogle Scholar
Turreson, G. 1922. The genotypical response of the plant species to the habitat. Hereditas 3: 211350.CrossRefGoogle Scholar
Turreson, G. 1925. The plant species in relation to habitat and climate. Hereditas 6: 147236.CrossRefGoogle Scholar
Turrill, W.B. 1934. The correlation of morphological variation with distribution in some species of Ajuga. New Phytologist 33: 218230.CrossRefGoogle Scholar
Turrill, W.B. 1942. Taxonomy and phylogeny. Botanical Review 8: 247270, 473–532, 655–707.CrossRefGoogle Scholar
Turrill, W.B. 1946. The ecotype concept: a consideration with appreciation and criticism especially of recent trends. New Phytologist 45: 3443.CrossRefGoogle Scholar
Tutin, T.G. 1953. The vegetation of the Azores. Journal of Ecology 41: 5361.CrossRefGoogle Scholar
US Fish and Wildlife Service 1986. Florida Torreya (Torreya taxifolia) Recovery Plan. Atlanta, GA: US Fish and Wildlife Service.Google Scholar
Ueda, Y. & Nishida, M. 1982. On petrified pine leaves from the Upper Cretaceous of Hokkaido. Journal of Japanese Botany 57: 133145.Google Scholar
Ueno, J. 1959. Studies on pollen grains of Gymnospermae, concluding remarks to the relationships between Coniferae. Journal of the Institute of Polytechnics, Osaka City University (ser D) 11: 109136.Google Scholar
Unger, F. 1850. Die Gattung Glyptostrobus in der Tertiar-Formation. Sitzungsberichte der Kaiserlichen Akademie der Wissenschaten in Wein, Mathematische-Naturwissenschaftliche lasse 5: 434435.Google Scholar
Unger, F. 1850. Genera et Species Plantarum Fossilium. Vienna: Wilhelm Braumuller.Google Scholar
Upchurch, G.R. & Doyle, J.A. 1981. Paleoecology of the conifers Frenlopsis and Pseudofrenlopsis (Cheirolepidiaceae) from the Cretaceous Potomac Group of Maryland and Virginia. In Romans, R.C. (ed.) Geobotany, Vol 2. New York: Plenum Press.Google Scholar
Upchurch, P. 1998. The phylogenetic relationships of sauropod dinosaurs. Zoological Journal of the Linnean Society 124: 43103.CrossRefGoogle Scholar
Upchurch, P. & Barrett, P.M. 2000. The evolution of sauropod feeding mechanisms. Pp 79122 in Sues, H.D. (ed.), Evolution of Herbivory in Terrestrial Vertebrates. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Upchurch, P., Barrett, P.M. & Dodson, P. 2004. Sauropsida. Pp 259324 in Weishampel, D.B., Dodson, P. & Osmolska, H. (eds.), The Dinosauria. Berkley, CA: University of California Press.CrossRefGoogle Scholar
Urretavizcaya, M.F. & Defosse, G.E. 2004. Soil seed bank of Austrocedrus chilensis (D.Don) Pic. Serm. et Bizarri related to different degrees of fire disturbance in two sites of southern Patagonia, Argentina. Forest Ecology and Management 187: 361372.CrossRefGoogle Scholar
Vaario, L.-M., Lu, J., Koistinen, A., Terahauta, A. & Aronen, T. 2015. Variation among matsutake ectomycorrhizae in four clones of Pinus sylvestris. Mycorrhiza 25: 195204.CrossRefGoogle ScholarPubMed
Vajda, V. & McLoughlin, S. 2004. Fungal proliferation at the Cretaceous–Tertiary boundary. Science 303: 1489.CrossRefGoogle ScholarPubMed
Vakhrameev, V.A. 1978. The climates of the Northern Hemisphere in the Cretaceous in the light of palaeobotanic data. Palaeontological Zhurnal 2: 312.Google Scholar
Vakhrameev, V.A. 1980. Classopollis pollen as an indicator of Jurassic and Cretaceous climates. Soviet Geology 8: 4856 (in Russian).Google Scholar
Vakhrameev, V.A. 1988. Jurasic and Cretaceous Floras and Climates of the Earth. Cambridge: Cambridge University Press.Google Scholar
Van der Ham, R.W.J.M., van Konijnenburg-van Cittert, J.H.A. & van den Burgh, J. 2001. Taxodiaceaous conifers from the Maastrichtian type area (Late Cretaceous, NE Belgium, SE Netherlands). Review of Palaeobotany and Palynology 116: 233250.CrossRefGoogle Scholar
Van der Ham, R.W.J.M., van Konijnenburg-van Cittert, J.H.A., Dortangs, R.W., Herngreen, G.F.W. & van den Burgh, J. 2003. Brachyphyllum patens (Miquel) comb. nov. (Cheirolepidiaceae?): remarkable conifer foliage from the Maastrichtian type area (Late Cretaceous, NE Belgium, SE Netherlands). Review of Palaeobotany and Palynology 127: 7797.CrossRefGoogle Scholar
Van der Hammen, T. & Hooghiemstra, H. 2001. Historia y paleoecología de los bosques montanos andinos neotropicales. Pp 6384 in Kappelle, M. & Brown, A.D. (eds.), Bosques Nublados del Neotropico. Santiago Domingo de Heredia: Instituto Nacional de Biodiversidad.Google Scholar
Van der Heijden, M.G.A., Boller, T., Wiemken, A. & Sanders, I.R. 1998. Different arbuscular mycorrhizal fungal species are potential determinant of plant community structure. Ecology 79: 20822091.CrossRefGoogle Scholar
Van der Heijden, M.G.A., Klronomos, J.N., Ursic, M., et al. 1998. Mycorrhizal fungal diversity determines plant biodiversity, ecosystems variability and productivity. Nature 396: 6972.CrossRefGoogle Scholar
Van der Heijden, M.G.A., Wiemken, A. & Sanders, I.R. 2003. Different arbuscular mycorrhizal fungi alter coexistence and resource distribution between co-occurring plants. New Phytologist 157: 569578.CrossRefGoogle Scholar
Van Hees, P.A.W., Godbold, D.L., Jentshke, G. & Jones, D.L. 2003. Impact of ectomycorrhizas on the concentration of simple organic acids in a forest soil. European Journal of Soil Science 54: 697706.CrossRefGoogle Scholar
Van Hees, P.A.W., Rosling, A., Lundström, U.S. & Findlay, R.D. 2006. The biogeochemical impact of ectomycorrhizal conifers on major soil elements (Al, Fe, K & Si). Geoderma 136: 364377.CrossRefGoogle Scholar
Van Konijnenburg-Van Cittert, J.H.A. 1971. New data of Pagiophyllum maculosum Kendall and its male cone from the Jurassic of North Yorkshire. Review of Palaeobotany and Palynology 51: 95105.CrossRefGoogle Scholar
Van Koninjnenburg-van Cittert, J.H.A. 2010. The Early Jurassic male ginkgoalean inflorescence Stachyopitys preslii Schenk and its in situ pollen. Scripta Geologica 7: 141149.Google Scholar
Van Mantgem, P. & Schwartz, M. 2003. Bark heat resistance of small trees in Californian mixed conifer forest: testing some model assumptions. Forest Ecology and Management 178: 341352.CrossRefGoogle Scholar
Van Mantgem, P.J. & Stephenson, N.L. 2005. The accuracy of matrix population model projection for coniferous trees in the Sierra Nevada, California. Journal of Ecology 93: 737747.CrossRefGoogle Scholar
Van Nieuwstadt, M.G.L. & Sheil, D. 2005. Drought, fire and tree survival in a Borneo rain forest, East Kalimantan, Indonesia. Journal of Ecology 93: 191201.CrossRefGoogle Scholar
Van Pelt, R. 2001. Forest Giants of the Pacific Coast. Vancouver: University of Washington Press in association with Global Forestry Society.Google Scholar
Van Steenis, C.G.G.J. 1957. Outline of the vegetation types in Indonesia and some adjacent regions. Proceedings of the Pacific Science Congress 8: 6197.Google Scholar
Van Steenis, C.G.G.J. 1961. Plant geography of the mountain flora of Kinabalu. Proceedings of the Royal Society of London B 161: 738.Google Scholar
Van Steenis, C.G.G.J. 1972. The Mountain Flora of Java. Leiden: E.J.Brill.Google Scholar
Van Valkenburg, J.L.C.H. & Ketner, P. 1994. Vegetation changes following human disturbance of mid-montane forest in the Wau area, Papua New Guinea. Journal of Tropical Ecology 10: 4154.CrossRefGoogle Scholar
Vander Wall, S.B. 1982. An experimental analysis of cache recovery in Clark’s nutcracker. Animal Behaviour 30: 8494.CrossRefGoogle Scholar
Vander Wall, S.B. 1992. The role of animals in dispersing a ‘wind dispersed’ pine. Ecology 73: 614621.CrossRefGoogle Scholar
Vander Wall, S.B. 2002. Masting in animal-dispersed pines facilitates seed dispersal. Ecology 83: 35083516.CrossRefGoogle Scholar
Vander Wall, S.B. & Hutchins, H.E. 1983. Dependence of Clark’s nutcracker Nucifraga columbiana, on conifer seeds during the postfledging period. Canadian Field Naturalist 97: 208214.CrossRefGoogle Scholar
Vankat, J.L. & Major, J. 1978. Vegetation changes in Sequoia National Park, California. Journal of Biogeography 5: 377402.CrossRefGoogle Scholar
Varner, J.M. III & Rush, J.S. 2004. Remnant old-growth longleaf pine (Pinus palustris Mill.) savannas and forests of the southeastern USA: status and threats. Natural Areas Journal 24: 141149.Google Scholar
Vasek, F.C. 1966. The distribution and taxonomy of three western junipers. Brittonia 18: 350372.CrossRefGoogle Scholar
Vaudois, N. & Privé, C. 1971. Révision des bois fossils de Cupressaceae. Palaeontographica B 134: 6186.Google Scholar
Veblen, T.T. 1992. Regeneration dynamics. Pp 152187 in Glen-lewin, D.C., Peet, R.K. & Veblen, T.T. (eds.), Plant Succesion: Theory and Prediction. London: Chapman & Hall.Google Scholar
Veblen, T.T. & Lorenz, D.C. 1987. Post-fire stand development of AustrocedrusNothofagus forests in northern Patagonia. Plant Ecology 71: 113126.CrossRefGoogle Scholar
Veblen, T.T. & Lorenz, D.C. 1988. Recent vegetation changes along the forest/steppe ecotone in northern Patagonia. Annals of the Association of American Geographers 78: 93111.CrossRefGoogle Scholar
Veblen, T.T., Schlegel, E.M. & Escobar, B. 1980. Structure and dynamics of old-growth Nothofagus forest in the Valdivian Andes, Chile. Journal of Ecology 68: 131.CrossRefGoogle Scholar
Veblen, T.T., Kitzbeerger, T. & Antonio, L. 1992. Disturbance and forest dynamics along a transect from Andean rain forest to Patagonian shrubland. Journal of Vegetation Science 3: 507520.CrossRefGoogle Scholar
Veevers, J.J., Powell, C.M. & Roots, S.R. 1991. Review of seafloor spreading around Australia. I. Synthesis of the pattern of spreading. Australian Journal of Earth Science 38: 373389.CrossRefGoogle Scholar
Veitch, J. 1906. Hortus Veitchii. London: James Veitch & Sons.Google Scholar
Velala, S.M., Rajala, T., Heinonsala, J., Taylor, A.F.S. & Pennamen, T. 2014. Profiling functions of ectomycorrhizal diversity and root structuring in seedlings of Norway spruce (Picea abies) with fast- and slow-growing phenotypes. New Phytologist 201: 610622.CrossRefGoogle Scholar
Verboom, G.A., Linder, H.P., Forest, F., et al. 2014. Cenozoic assembly of the Greater Cape Flora. Pp 93118 in Allsopp, N., Colville, J.F. & Verboom, G.A. (eds.), Fynbos Ecology, Evolution and Conservation of a Megadiverse Region. Oxford: Oxford University Press.CrossRefGoogle Scholar
Verkaik, I. & Espelta, J.M. 2006. Post-fire regeneration thinning, cone production, serotiny and regeneration age in Pinus halepensis. Forest Ecology and Management 231: 155163.CrossRefGoogle Scholar
Verkaik, E., Gardner, R.O. & Braakhekke, W.G. 2006. Site conditions affect seedling distribution below and outside the crown of kauri trees (Agathis australis). New Zealand Journal of Ecology 31: 1321.Google Scholar
Viani, R.A.G., Rodrigues, R.R., Dawson, T.E. & Rafael, S.O. 2011. Savanna soil fertility limits growth but not survival of tropical forest tree seedlings. Plant and Soil 349: 341353.CrossRefGoogle Scholar
Vilar de Seoane, L., Cúneo, N.R., Eszepa, I., Wilf, P. & Gandolfo, M.A. 2015. Ginkgoites patagonica (Berry) comb.nov. from the Eocene of Patagonia, the last Ginkgoleans record in South America. International Journal of Plant Science 176: 346363.CrossRefGoogle Scholar
Villa-Castillo, J. & Wagner, M.R. 2002. Ground beetle (Coleoptera: Carabidae) species assemblage as an indicator of forest condition in northern Arizona Ponderosa pine forests. Environmental Entomology 31: 242252.CrossRefGoogle Scholar
Villagran, C. 1988. Late Quaternary vegetation of southern Isla Grande de Chiloé, Chile. Quaternary Research 29: 294306.CrossRefGoogle Scholar
Villalba, R. 1990. Latitude of the surface high-pressure belt over western South America during the last 500 years as inferred from tree-ring analysis. Quaternary of South America and Antarctic Peninsula 7: 273303.Google Scholar
Villalba, R. 1990. Climatic fluctuations in northern Patagonia during the last 1,000 years as inferred from tree-ring records. Quaternary Research 34: 346360.CrossRefGoogle Scholar
Villalba, R. & Veblen, T.T. 1997. Regional patterns of tree population age structure in northern Patagonia: climatic and disturbance influences. Journal of Ecology 85: 113124.CrossRefGoogle Scholar
Vincent, L., Piere, G., Michel, S. Robert, N. & Masson-Delmotte, V. 2007. Tree-rings and the climate of New Caledonia (SW Pacific): preliminary results from Araucariaceae. Palaeogeography, Palaeoclimatology, Palaeoecology 253: 477489.CrossRefGoogle Scholar
Vitousek, P.M. & Howarth, R.W. 1991. Nitrogen limitation on land and in the sea: how can it occur? Biogeochemistry 13: 87115.CrossRefGoogle Scholar
Viveros-Viveros, H., Sáenz-Romero, C., López-Upton, J. & Vargas-Hernández, J. 2007. Growth and frost damage variation among Pinus pseudostrobus, P. montezumae and P. hartwegii tested in Michoacan, Mexico. Forest Ecology and Management 253: 8188.CrossRefGoogle Scholar
Vogel, J.C., Rumsey, F.J., Schneller, J.J., Barrett, J.A. & Gibby, M. 1999. Where are the glacial refugia in Europe? Evidence from pteridophytes. Botanical Journal of the Linnean Society 66: 2337.CrossRefGoogle Scholar
Volkova, P.A., Travnicek, P. & Brochmann, C. 2010. Evolutionary dynamics across discontinuous freshwater ecosystems: rapid expansions and repeated allopolyploid origins in Palearctic white water-lilies (Nymphaceae). Taxon 59: 483494.CrossRefGoogle Scholar
Vozenin-serra, C. & Salard-Cheboldaeff, M. 1992. Les bois minéralisés permo-triasiques de Nouvelle-Calédonie. Implications phylogénetique et paléogéographique. Palaeontographica B 225: 125.Google Scholar
Wace, N.M. 1972. Discussion on the plant geography around Torres Strait. Pp 197211 in Walker, D. (ed.), Bridge and Barrier: The Natural and Cultural History of Torres Strait. Canberra: Australian National University Press.Google Scholar
Wagg, C., Pautler, M., Masicotte, H.B. & Peterson, R.L. 2008. The co-occurrence of ectomycorrhizal and dark separate fungi in seedlings of four members of the Pinaceae. Mycorrhiza 18: 103110.CrossRefGoogle Scholar
Wagstaff, S.J. & Garnock-Jones, P.J. 1998. Evolution and biogeography of the Hebe complex (Scrophulariaceae) inferred from ITS sequences. New Zealand Journal of Botany 36: 425437.CrossRefGoogle Scholar
Waldrop, T.A. & Brose, P.H. 1999. A comparison of fire intensity levels for stand replacement of table mountain pine (Pinus pungens Lamb.). Forest Ecology and Management 113: 155166.CrossRefGoogle Scholar
Walker, C. 1986 Mycorrhizal growth enhancement in Sitka spruce seedlings. P 213 in Molina, R. (ed.,) Proceedings of the 6th North American Conference on Mycorrhizae. Corvallis, OR: Forest Research Laboratory.Google Scholar
Walker, C. 1987. Sitka spruce mycorrhizas. Proceedings of the Royal Society of Edinburgh 93B: 117119.Google Scholar
Walker, D. 1982. Speculations on the origin and evolution of Sunda and Sahul rain forests. Pp. 554575 in Prance, G.T. (ed.), Biological Diversification in the Tropics. New York: Columbia University Press.Google Scholar
Walker, J.W. 1972. Chromosome numbers, phylogeny, phytogeography of the Annonaceae and their bearing on the (original) basic chromosome numbers of angiosperms. Taxon 21: 5765.CrossRefGoogle Scholar
Walker, J.W. & Walker, A.G. 1984. Ultrastructure of Lower Cretaceous angiosperm pollen and the origin and early evolution of flowering plants. Annals of the Missouri Botanical Garden 71: 464521.CrossRefGoogle Scholar
Walker, L.R. & Shiels, A.B. 2008. Post-disturbance erosion impacts carbon fluxes and plant succession on recent tropical landslide. Plant and Soil 313: 205216.CrossRefGoogle Scholar
Walker, T.G. 1958. Hybridisation in some species of Pteris. Evolution 12: 8292.CrossRefGoogle Scholar
Walkom, A.B. 1919. Mesozoic floras of Queensland. Parts 3 & 4. The floras of the Burrum and Styx River Series. Geological Survey of Queensland Publication 263: 177.Google Scholar
Walters, S.M. The shaping of angiosperm taxonomy. New Phytologist 60: 7484.CrossRefGoogle Scholar
Wang, C. 1968. The coniferous forests of Taiwan. Tunghai University Biological Bulletin 34: 152.Google Scholar
Wang, C.S. & Ding, X.L. 1998. New progress of the study on Himalaya uplift and its sedimentary response. Geological Science and Technology Information 17: 17.Google Scholar
Wang, D.L., Li, Z.C., Hao, G. & Ge, X.J. 2004. Genetic diversity of Calocedrus macrolepis (Cupressaceae) in southwestern China. Biochemical Systematics and Ecology 32: 797807.CrossRefGoogle Scholar
Wang, D.L., Li, Z.C. & Ge, X.J. 2005. Genetic diversity of the relict plant Amentotaxus yunnanensis. Journal of Subtropical Botany 13: 143148.Google Scholar
Wang, F.H. 1948. The early embryogeny of Glyptostrobus. Botanical Bulletin Academica Sinica 2: 110.Google Scholar
Wang, F.H., Chen, Z.K. & Hu, Y.S. 1979. On the systematic position of Taxaceae from the embryological and anatomical studies. Acta Phytotaxonomica Sinica 17: 17 (in Chinese with English summary).Google Scholar
Wang, F.-X. (ed.) 1990. The Biology of Cathaya argyrophylla. Beijing: Science Press.Google Scholar
Wang, H.-Y., Hu, Y.-S. & Wang, F.-H. 1995. Anatomy of gymnosperms endemic to China. III. Pseudotaxus chienii Cheng. Cathaya 7: 147163.Google Scholar
Wang, T., Su, Y.J., Huang, C. & Zhu, J.M. 1999. RAPD analyses of Nageioids. Acta Botanica Yunnanica 21: 144148.Google Scholar
Wang, W.P. 2003. New species of Cupressus. Plant Systematics and Evolution 241: 1328.CrossRefGoogle Scholar
Wang, X., Wang, Z., Xiao, Y. Duan, R. & Zhao., X. 2005. Ecological plasticity of Larix chinensis population cones and seeds in Quinlin Mountain. Chinese Journal of Applied Ecology 16: 2932.Google Scholar
Wang, X.Q. & Shu, Y.Q. 2000. Chloroplaast mat K gene phylogeny of Taxaceae and Cephalotaxaceae, with additional reference to the systematic position of Nageia. Acta Phytotaxonomica Sinica 38: 201210.Google Scholar
Wang, X.-R., Tsumura, Y., Yoshimaru, H., Nagasaka, K. & Szmidt, A.E. 1999. Phylogenetic relationships of European pines (Pinus, Pinaceae) based on chloroplast rbcL, matK, rpl20-rpsl8 spacer, and trnV intron sequences. American Journal of Botany 86: 17421753.CrossRefGoogle Scholar
Wang, Y.D., Guignard, G., Thevenard, F., et al. 2005. Cuticular anatomy of Sphenobaiera huangii (Ginkgoales) from the Lower Jurassic of Hubei, China. American Journal of Botany 92: 709721.CrossRefGoogle Scholar
Wang, Z., Wang, N., Wu, L., Zhang, J. & Chen, Y. 1995. Measurement of available B content and the effect on application of B fertiliser in Chinese fir seed orchard. Forest Research 8: 634640.Google Scholar
Warburg, O. 1900. Beitrage zur Kenntnissder Vegetation des Sud- und Ostasiatischen Monsungebietes. Monsumia 1: 1189.Google Scholar
Wardle, P. 1963. Growth habits of New Zealand subalpine shrubs and trees. New Zealand Journal of Botany 1: 1847.CrossRefGoogle Scholar
Wardle, P. 1969. Biological flora of New Zealand. 4. Phyllocladus alpinus Hook. F. (Podocarpaceae). New Zealand Journal of Botany 7: 7695.CrossRefGoogle Scholar
Wardle, P. 1977. Plant communities of Westland National Park (New Zealand) and neighbouring coastal and lowland areas. New Zealand Journal of Botany 15: 323398.CrossRefGoogle Scholar
Wardle, D.A., Bardgett, R.D., Klironomos, J.N., et al. 2004. Ecological linkages between aboveground and belowground biota. Science 304: 16341637.CrossRefGoogle ScholarPubMed
Wardle, D.A., Walker, L.R. & Bardgett, R.D. 2004. Ecosystem properties and forest decline in contrasting long-term chronosequences. Science 305: 509513.CrossRefGoogle ScholarPubMed
Waring, R.H. 1969. Forest plants of the eastern Siskiyous: their environmental and vegetational distribution. Northwest Science 43: 117.Google Scholar
Waring, R.H. & Major, J. 1964. Some vegetation of the California coastal redwood region in relation to gradients of moisture, nutrients, light and temperature. Ecological Monographs 34: 167215.CrossRefGoogle Scholar
Warnaar, J., Bijl, P.K. & Huber, M. Orbitally forced climate changes in the Tasman sector during the Middle Eocene. Palaeogeography, Palaeoclimatology, Palaeoecology 280: 361370.CrossRefGoogle Scholar
Wasscher, J.A. 1941. The genus Podocarpus in the Netherlands Indies. Blumea 4: 359481.Google Scholar
Watling, R. 1990. Mycorrhizal fungi. Bulletin of the British Mycological Society 14: 5962.CrossRefGoogle Scholar
Watling, R. & Dobbie, L. 1991. Endomycorrhizae in glasshouse grown conifers. Botanical Journal of Scotland 46: 145151.CrossRefGoogle Scholar
Watson, E. & Luckmann, B.H. 2002. The dendroclimatic signal in Douglas-fir and ponderosa pine tree-ring chronologies from the southern Canadian Cordillera. Canadian Journal of Forest Research 32: 18581874.CrossRefGoogle Scholar
Watson, J. 1974. Manica: a new fossil conifer genus. Taxon 23: 428.CrossRefGoogle Scholar
Watson, J. 1977 Some Lower Cretaceous conifers of the Cheirolepiiaceae from the U.S.A. and England. Palaeontology 20: 715749.Google Scholar
Watson, J. & Alvin, K. 1996. An English Wealden floral list, with comments on possible environmental indicators. Cretaceous Research 17: 526.CrossRefGoogle Scholar
Watson, J. & Fischer, H.L. 1984. A new conifer genus from the Lower Cretaceous Glen Rose Formation, Texas. Palaeontology 27: 719727.Google Scholar
Watson, J. & Sincock, C.A. 1992. Bennettitales of the English Wealden. Monograph of the Palaeontographical Society of London 145: 1288.Google Scholar
Watson, J., Fischer, H.L. & Hall, N.A. 1987. A new species of Brachyphyllum from the English Wealden and its probable female cone. Review of Palaeobotany and Palynology 51: 169187.CrossRefGoogle Scholar
Weathers, K.C. 1999. The importance of cloud and fog in the maintenance of ecosystems. Trends in Ecology and Evolution 14: 214215.CrossRefGoogle Scholar
Weathers, K.C., Lovett, G.M. & Likens, G.E. 1995. Cloud deposition to a spruce forest edge. Atmosphere and Environment 29: 665672.CrossRefGoogle Scholar
Weathers, K.C., Lovett, G.M., Likens, G.E. & Caraco, M.F.M. 2000. Cloudwater inputs of nitrogen to forest ecosystems in southern Chile: forms, fluxes and sources. Ecosystems 3: 590595.CrossRefGoogle Scholar
Weaver, J.C. 1983. The improbable endotherm: the energetics of the sauropod dinosaur Brachiosaurus. Paleobiology 9: 173182.CrossRefGoogle Scholar
Webb, L.J. 1969. Edaphic differentiation of some forest types in eastern Australia. II. Soil chemical factors. Journal of Ecology 57: 817830.CrossRefGoogle Scholar
Webb, L.J. & Tracey, J.G. 1975. The Cooloola rain forests. Proceedings of the Ecological Society of Australia 9: 317321.Google Scholar
Webb, L.J. & Tracey, J.G. 1981. The rainforests of northern Australia. Pp 67101 in Groves, R.H. (ed.), Australian Vegetation. Cambridge: Cambridge University Press.Google Scholar
Webb, L.J., Tracey, J.G. & Williams, W.T. 1972. Regeneration and pattern in the subtropical rainforest. Journal of Ecology 60: 675695.CrossRefGoogle Scholar
Webb, L.J., Tracey, J.G. & Williams, W.T. 1984. A floristic framework of Australian rainforests. Australian Journal of Ecology 9: 169198.CrossRefGoogle Scholar
Webb, L.J., Tracey, J.G. & Jessup, L.W. 1986. Recent evidence for the autochthony of Australian tropical and subtropical rainforest floristic elements. Telopea 2: 575590.CrossRefGoogle Scholar
Wells, A., Stewart, G.G. & Duncan, R.P. 1998. Evidence for widespread, synchronous, disturbance-initiated forest establishment in Westland, New Zealand. Journal of the Royal Society of New Zealand 28: 333345.CrossRefGoogle Scholar
Wells, B.W. 1928. Plant communities of the coastal plain of North Carolina and their successional relationships. Ecology 9: 230242.CrossRefGoogle Scholar
Wen, Z., Murata, M., Xu, Z., Chen, Y. & Nara, K. 2015. Ectomycorrhizal fungal communities on the endangered Chinese Douglas-fir (Pseudotsuga chinensis) indicating regional fungal sharing overrides host conservatism across geographical regions. Plant and Soil 387: 189199.CrossRefGoogle Scholar
Wessel, P. & Kroenke, L.W. 2007. Reconciling late Neogene Pacific absolute and relative plate motion changes. Geochemistry, Geophysics and Geosystems 8: Q08001.CrossRefGoogle Scholar
Westhoff, M., Zimmermann, D., Zimmerman, G., et al. 2009. Distribution and function of epistomatal mucilage plugs. Protoplasma 235: 101105.CrossRefGoogle ScholarPubMed
Whang, S.S., Park, J.-H., Hil, R.S. & Kim, K. 2001. Cuticle micromorphology of leaves of Pinus (Pinaceae) from Mexico and Central America. Botanical Journal of the Linnean Society 135: 349373.CrossRefGoogle Scholar
White, C.T. 1947. Note on two species of Araucaria in New Guinea and a proposed new section of the genus. Journal of the Arnold Arboretum 28: 259260.CrossRefGoogle Scholar
White, P.S., MacKenzie, M.D. & Busing, R.T. 1985. A critique of overstory/understory comparisons based on transition probability analysis of an old growth spruce-fir stand in the Appalachians. Vegetatio 64: 3745.CrossRefGoogle Scholar
Whitmore, T.C. 1966. The social status of Agathis in rainforest in Melanesia. Journal of Ecology 54: 285301.CrossRefGoogle Scholar
Whitmore, T.C. 1980. Utilisation, potential and conservation of Agathis, a genus of tropical Asian conifers. Economic Botany 34: 112.CrossRefGoogle Scholar
Whitmore, T.C. 1982. On pattern and process in forest. Pp 4249 in Newman, E.I. (ed.), The Plant Community as a Working Mechanism. Oxford: Blackwell.Google Scholar
Whitmore, T.C. 1984. Tropical Rainforests of the Far East, 2nd edn. Oxford: Clarendon Press.Google Scholar
Whitmore, T.C. 1984. A vegetation map of Malesia at a scale of 1:5 million. Journal of Biogeography 11: 461471.CrossRefGoogle Scholar
Whitmore, T.C. 1990. An Introduction to Tropical Rainforests. Oxford: Clarendon Press.Google Scholar
Whitmore, T.C. & Bowen, M.R. 1983. Growth analyses of some Agathis species. Malaysian Forester 46: 186196.Google Scholar
Whittaker, R.H. 1954. The vegetation response to serpentine soils. Ecology 35: 275288.Google Scholar
Whittaker, R.H. 1956. Vegetation of the Great Smokey Mountains. Ecological Monographs 26: 180.CrossRefGoogle Scholar
Whittaker, R.H. 1975. Communities and Ecosystems, 2nd edn. New York: MacMillan.Google Scholar
Whittaker, R.J., Triantis, K.A. & Ladle, R.J. 2008. A general dynamic theory of island biogeography. Journal of Biogeography 35: 977994.CrossRefGoogle Scholar
Whittlake, E.B. 1972. Taxonomic note on fossil Glyptostrobus in northeastern Arkansas. Proceedings of the Arkansas Academy of Science 26: 13.Google Scholar
Wieland, G.R. 1929. The world’s two greatest petrified forests. Science 69: 6063.CrossRefGoogle ScholarPubMed
Wieland, G.R. 1935. The Cerro Cuadrado Petrified Forest. Washington, DC: Carnegie Institute of Washington.Google Scholar
Wierman, C.A. & Oliver, C.D. 1979. Crown stratification by species in even-aged mixed stands of Douglas-fir-western Hemlock. Canadian Journal of Forest Research 9: 19.CrossRefGoogle Scholar
Wilczynaski, S. & Shrzyszewsky, J. 2002. Das Klimatische Signal in den Jahrringen von Kiefern (Pinus sylvestris L.) im Sudeten-Vorland (Sud-Polen). Forstwissenschaftenschaftliches Centralblatt 121: 1524.CrossRefGoogle Scholar
Wild, H. & Bradshaw, A.D. 1977. The evolutionary effects of metalliferous and other anomalous soils in south-central Africa. Evolution 31: 282293.CrossRefGoogle ScholarPubMed
Wilde, M.H. 1944. A new interpretation of coniferous cones. I. Podocarpaceae Podocarpus. Annals of Botany n.s. 8: 141.CrossRefGoogle Scholar
Wilde, M.H. & Eames, A.J. 1952. The ovule and ‘seed’ of Araucaria bidwillii, with discussion of the taxonomy of the genus. II. Taxonomy. Annals of Botany 16: 2747.CrossRefGoogle Scholar
Wilf, P. & Escapa, I. 2015. Green web or megabiased clock? Plant fossils from Gondwanan Patagonia speak on evolutionary radiations. New Phytologist 207: 283290.CrossRefGoogle ScholarPubMed
Wilf, P., Labandeira, C.C., Johnson, K.R. & Cuneo, N.R. 2005. Richness of plant–insect associations in Eocene Patagonia: a legacy for South American biodiversity. Proceedings of the National Academy of Sciences, USA 102: 89448948.CrossRefGoogle ScholarPubMed
Wilf, P., Cúneo, N.R., Escapa, I.H., Pol, D. & Woodburne, M.O. 2013. Splendid and seldom isolated: the palaeobiogeography of Patagonia. Annual Review of Earth and Planetary Science 41.CrossRefGoogle Scholar
Wilkinson, D.M. 1997. The role of seed dispersal in the evolution of mycorrhizae. Oikos 78: 394396.CrossRefGoogle Scholar
Wilkinson, D.M. & Ruxton, G.D. 2013. High C/N ratio (not low energy content) of vegetation may have driven gigantism in sauropod dinosaurs and perhaps omnivory and/or endothermy in their juveniles. Functional Ecology 27: 131135.CrossRefGoogle Scholar
Williams, A., Ridgway, H.J. & Norton, D.A. 2011. Growth and competitiveness of the New Zealand tree species of Podocarpus cunninghamii is reduced by ex-agricultural AMF but enhanced by forest AMF. Soil Biology and Biochemistry 43: 339345.CrossRefGoogle Scholar
Williams, C. 1999. Species accounts: Kremp’s pine (Pinus krempfii Lecomte). Pp 105106 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Williams, C.G., LaDeau, S.L., Oren, R. & Katul, G.G. 2006. Modeling seed dispersal distances: implications for transgenic Pinus taeda. Ecological Applications 16: 117124.CrossRefGoogle ScholarPubMed
Williams, M.A.J. 1984. Quaternary environments. Pp 4247 in Veevers, J.J. (ed.), Phanerozoic Earth History of Australia. Oxford: Clarendon Press.Google Scholar
Williamson, G.B. & Black, E.M. 1981. High temperature of forest fires under pines as a selective advantage over oaks. Nature 293: 643644.CrossRefGoogle Scholar
Williamson, R. & Williamson, K. 1973. The bird community of yew woodland at Kingley Vale, Sussex. British Birds 66: 1223.Google Scholar
Willig, M.R., Kaufman, D.M. & Stevens, R.D. 2003. Latitudinal gradients of biodiversity: pattern, process, scale and synthesis. Annual Review of Ecology, Evolution and Systematics 34: 273309.CrossRefGoogle Scholar
Wilson, E.H. 1926. The taxads and conifers of Yunnan. Journal of the Arnold Arboretum 7: 3768.CrossRefGoogle Scholar
Wilson, M.F., Sabag, C., Figueroa, J. & Armesto, J.J. 1996. Frugivory and seed dispersal of Podocarpus nubigena in Chiloé, Chile. Revista Chilena de Historia Natural 69: 343349.Google Scholar
Wings, O. & Sander, P.M. 2007. No gastric mill in sauropod dinosaurs: new evidence from analysis of gastrolith mass and function in ostriches. Proceedings of the Royal Society of London B 274: 635640.Google ScholarPubMed
Winkworth, R.C., Wagstaff, S.J., Glenny, D. & Lockhardt, P. 2002. Plant dispersal news from New Zealand. Trends in Ecology and Evolution 17: 514520.CrossRefGoogle Scholar
Wolf, C.B. 1948. Taxonomic and distributional studies of the New World cypresses. El Aliso 1: 1250.CrossRefGoogle Scholar
Wolfe, J.A. 1971. Tertiary climatic fluctuations and methods of analysis of Tertiary floras. Palaeogeography, Palaeoclimatology, Palaeoecology 9: 2757.CrossRefGoogle Scholar
Wolfe, J.A. 1980. Tertiary climates and floristic relationships at high latitudes in the northern hemisphere. Palaeogeography, Palaeoclimatology, Palaeoecology 30: 313323.CrossRefGoogle Scholar
Wolfe, J.A. 1990. Palaeobotanical evidence for a marked temperature increase following the Cretaceous/Tertiary boundary. Nature 343: 153156.CrossRefGoogle Scholar
Wolfe, J.A. 1991. Palaeobotanical evidence for a June ‘impact winter’ at the Cretaceous/Tertiary boundary. Nature 352: 420423.CrossRefGoogle Scholar
Wolfe, J.A. & Upchurch, G.R. Jr. 1986. Vegetation, climatic and floral changes at the Cretaceous–Tertiary boundary. Nature 324: 148152.CrossRefGoogle Scholar
Wolfe, J.A., Schorn, H.E., Forest, C.E. & Molnar, P. 1997. Paleobotanical evidence for high altitudes in Nevada during the Miocene. Science 276: 16721675.CrossRefGoogle Scholar
Woltz, P. 1969. Une nouvelle espéce de Podocarpus de Madagascar: P. gaussenii. Travaux Laboratoire Forestiere Toulouse 8(2): 18.Google Scholar
Woltz, P. 1969. Morphologies et anatomie de quelgues plantules de Podocarpus malagaches. Bulletin Société Botanique de France 116(7–8): 349355.CrossRefGoogle Scholar
Woltz, P. 1985. Place des gymnosperms endémiques des Andes méridionales dans la végétation du Chili. Lazaroa 8: 293314.Google Scholar
Woltz, P. 1987. Les Podocarpineae, étude des plantules et évolution: comparison avec les familles de Conifères de l’hémisphère sud: Arauariaceae, Cupressaceae, Taxodiaceae, Taxaceae. Bulletin de la Société Botanique de France 134: 141151.CrossRefGoogle Scholar
Woltz, P. & Baily, Y. 1982. Austrotaxus spicata Compton de Nouvelle Calédonie: aspects anatomiques et évolution de rappares conducteur de la plante. Bulletin Société Botanique de France 126(3): 223230.CrossRefGoogle Scholar
Woltz, P. & Rouane, M.L. 1977. Phyllocladus trichomanoides D.Don de Nouvelle Zéland: aspects anatomique et evolution de l’appareil conducteur de la plantule. Bulletin Société Botanique de France 124: 6167.CrossRefGoogle Scholar
Woltz, P. & Rouane, M.L. 1980. A propos du Microcachrys tetragona Hook. F. (Podocarpacées) et de evolution vasculaire de la plantule. Bulletin Société Botanique de France, Lettres Botanique 127: 151158.CrossRefGoogle Scholar
Woltz, P. & Rouane, M.L. 1987. Stachycarpus amara (Blume) Gaussen, Podocarpaceae d’Australasie: Aspects anatomiques et évolution de l’appareil conducteur de la plantule. Gaussenia 3: 311.Google Scholar
Woltz, P., Gajardo, R. & Provane, M.L. 1984. A propos du Saxegothaea conspicua Lindl. Podocarpine endémique des Anders meridionales et de l’évolution vasculaire de la plantule. Gaussenia 1: 518.Google Scholar
Woltz, P., Rouane, M.L. & Gondran, M. 1993. Apport des cotyledons dans l’evolution des Podocarpineae. Gaussenia 8: 615.Google Scholar
Woltz, P., Stockey, R.A., Gondran, M. & Cherrier, J.-F. 1995. Interspecific parasitism in the gymnosperms: unpublished data on two endemic New Caledonian Podocarpaceae using scanning electron microscopy. Acta Botanica Gallica 141: 731746.CrossRefGoogle Scholar
Woltz, P., Gondran, M., Marguerier, J. & Gajardo, R. 1998. Xylogie des conifèrs endemiques des Andes meridiaonales au MEB III. Podocarpaceae. Revue de Cytologie et Biologie Végétales – Le Botaniste 21: 314.Google Scholar
Woodburne, M.O. & Case, J.A. 1996. Dispersal, vicariance, and the late Cretaceous to early Tertiary land mammal biogeography from South America to Australia. Journal of Mammal Evolution 3: 212–161.CrossRefGoogle Scholar
Woods, D. 1976. Remarks on Widdringtonia. Veld and Flora 62: 32.Google Scholar
Woods, J.H., Blake, G.M. & Allendorf, F.W. 1983. Amount and distribution of isozyme variation in Ponderosa pine from eastern Montana. Silvae Genetica 32: 151157.Google Scholar
Woodward, F.I., Lomas, M.R. & Kelly, C.K. 2004. Global climate and the distribution of plant biomes. Philosophical Transactions of the Royal Society London B 359: 14651476.CrossRefGoogle ScholarPubMed
Worsdell, W.C. 1901. The morphology of the ‘flowers’ of Cephalotaxus. Annals of Botany 15: 637652.CrossRefGoogle Scholar
Wright, I.J., Reich, P.B. & Westoby, M. 2001. Strategy shifts in leaf physiology, structure and nutrient content between species of high- and low-rainfall and high- and low-nutrient habitats. Functional Ecology 15: 423434.CrossRefGoogle Scholar
Wright, J.A., Marin, A.M. & Dvorak, W.S. 1996. Conservation and use of the Pinus chiapensis genetic resource in Colombia. Forest Ecology and Management 88: 283288.CrossRefGoogle Scholar
Wright, J.W. 1959. Species hybridisation in the white pines. Forest Science 5: 210222.Google Scholar
Wright, S., Berch, S. & Berbee, M. 2009. The effect of fertilization on the below-ground diversity and community composition of ectomycorrhizal fungi associated with western hemlock (Tsuga heterophylla). Mycorrhiza 19: 267276.CrossRefGoogle Scholar
Wright, S.J. 2002. Plant diversity in tropical forests: a review of mechanisms and species coexistence. Oecologia 130: 114.CrossRefGoogle ScholarPubMed
Wright, S.J., Muller-Landau, H.C., Condit, R. & Hubbell, S.P. 2003. Gap-dependent recruitment, realised vital rates, and size distributions of tropical trees. Ecology 84: 31743185.CrossRefGoogle Scholar
Wu, C.L. 1956. The taxonomic revision and phytogeographical study of Chinese pines. Acta Phytogeographical Sinica 5: 131163 (in Chinese).Google Scholar
Wu, J.J., Liao, W.B., Cui, D.F. & Chen, Z.M. Research on the communities with Taxus mairei in north Guangdong province: species diversity and population pattern. Guihaia 22: 6166.Google Scholar
Wu, S.-C. & Chern, J.-H. 1995. Group analysis as applied to wood anatomy of Taxodiaceae members. Taiwan University Research Reports 35: 360374.Google Scholar
Wu, Z., Sun, H., Zhou, Z., Peng, H. & Li, D. 2005. Origin and differentiation of endemism in the flora of China. Acta Botanica Yunnanica 27: 577604.Google Scholar
Wubet, T., Kottke, I., Teketay, D. & Oberwinkleer, F. 2003. Mycorrhizal status of indigenous trees in dry Afromontane forests of Ethiopia. Forest Ecology and Management 179: 387399.CrossRefGoogle Scholar
Wubet, T., Weiss, M., Kottke, I., Teketay, D. & Oberwinkleer, F. 2006. Phylogenetic analysis of nearly complete SSU rDNA sequences reveals that the endangered African pencil cedar, Juniperus procera Hochst. Ex Endl., is associated with distinct Glomus taxa. Mycological Research 110: 10591069.CrossRefGoogle Scholar
Xi, Y.-Z. 1986. Studies on pollen morphology of Taxaceae from China. Acta Phytotaxonomica Sinica 24: 247252 (in Chinese with English summary).Google Scholar
Xiang, Q.Y. 2000. Timing the eastern Asian–eastern North American floristic disjunction: molecular clock corroborates paleontological estimates. Molecular Phylogenetics and Evolution 15: 462472.CrossRefGoogle Scholar
Xiang, S.Y., Li, S.M., Wang, F., Han, K.J. & Wang, L. 2007. Comparative morphology and its systematic implication on epiphylous microsporangia from Ginkgo biloba L. Acta Horticultura Sinica 34: 805812 (in Chinese with English abstract).Google Scholar
Xie, Y.Q. 1993. Cathaysian paleocontinent and Cathaysian flora. Bulletin of Botanical Research 13: 202209.Google Scholar
Ya, T. 1999. Species accounts: dawn redwood (Metasequoia glyptostroboides Hu & W.C. Cheng). Pp 103104 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Yadav, R.R. & Battacharya, A. 1994. Tree ring evidence of the 1991 earthquake of Uttarkashi, Western Himalaya. Current Science 66: 862864.Google Scholar
Yaldwym, J.C. 1956. A preliminary account of the sub-fossil avifauna of the Martinborough caves. Records of the Dominion Museum 3: 17.Google Scholar
Yamada, I. 1977. Forest ecological studies on the montane forest of Mt. Pangrango, West Java. IV. Floristic composition along the altitude. South East Asian Studies 15: 226251.Google Scholar
Yamada, I. 1990. The changing pattern of vertical stratification along an altitudinal gradient of the forests of Mt. Pangrango, West Java. Pp in Baas, P., Kalkman, K. & Geesink, R. (eds.), The Plant Diversity of Malesia. Proceedings of the Fourth Malesiana Symposium. Leiden: Kluwer Academic Publishers.Google Scholar
Yamamoto, S. 1993. Seedling establishment of Chamaecyparis obtusa in different microenvironments in the Akasawa Forest Reserve, Central Japan. Journal of the Japanese Forestry Society 75: 519527.Google Scholar
Yamamoto, S. 1993. Gap characteristics and gap regeneration in a subalpine coniferous forest on Mt Ontake, Central Honshu. Japanese Ecological Research 8: 277285.CrossRefGoogle Scholar
Yamamoto, S. 1994. Seedling establishment of Chamaecyparis pisifera on different substrata in an old-growth C. pisifera stand, Akasawa Forest Reserve, central Japan. Journal of the Japanese Forestry Society 76: 553559.Google Scholar
Yamamoto, S.-I. & Moriyama, Y. 1997. Stand structure and the regeneration of Chamaecyparis pisifera (Sieb. & Zucc.) Endl. on sites with different soil development in an old-growth coniferous forest, Central Japan. Journal of Forest Research 2: 133140.CrossRefGoogle Scholar
Yamamoto, S., Moriyama, Y. & Kobayashi, M. 1994. Two types of vegetative reproduction of Chamaecyparis pisifera (Sieb. Et Zuc.) Endlicher. Japan Journal of Forest and Environment 36: 5759.Google Scholar
Yamasaki, S. & Dillenburg, L.R. 1999. Measurements of leaf relative water content in Araucaria angustifolia. Revista Brasileira de Fisiologia Vegetal 11: 6975.Google Scholar
Yamazaki, T. & Takeoka, M. 1956. On the identification of the pollen of Taxodiaceae. Saikyo University Agriculture Science Reports 8: 1016.Google Scholar
Yang, H. & Jin, J.-H. 2000. Phytogeographic history and evolutionary stasis of Metasequoia: geological and genetic information contrasted. Acta Palaeontologica Sinica 39: 288307.Google Scholar
Yang, H., Huang, Y., Leng, Q., LePage, B.A. & Williams, C.J. 2005. Biomolecular preservation of Tertiary Metasequoia fossil Lagerstatten revealed by comparative pyrolysis analysis. Review of Palaeobotany and Palynology 134: 237256.CrossRefGoogle Scholar
Yang, J., Yang, X. & Liang, H. 2004. The discovery of buried Metasequoia wood in Lichuan, Hubei, China, and its significance. Acta Paleontologica Sinica 43: 124131.Google Scholar
Yang, Q.-J., Xu, H., Yan, Z.-G., et al. 2006. Natural resource and conservation of Taiwania cryptomerioides in Hubei Province. Guihaia 26: 551556.Google Scholar
Yang, Q.-J., Chen, G.-F., Liu, X.-Q. & Chen, L.-Q. 2009. Analysis of genetic diversity of Taiwania cryptomerioides in Xingdoushan, Hubei province. Guihaia 29: 450454.Google Scholar
Yang, Z.-Y., Ran, J.-H. & Wang, X.-Q. 2012. Three genome-based phylogeny of Cupressaceae s.l.: further evidence for the evolution of gymnosperms and Southern Hemisphere biogeography. Molecular Phylogenetics and Evolution 64: 452470.CrossRefGoogle Scholar
Yao, X., Taylor, T.N. & Taylor, E.L. 1991. Silicified dipterid ferns from the Jurassic of Antarctica. Review of Palaeobotany and Palynology 67: 353362.CrossRefGoogle Scholar
Yates, A.M. & Kitching, J.W. 2003. The earliest known sauropod dinosaur and the first steps towards sauropod locomotion. Proceedings of the Royal Society of London B 270: 17531758.CrossRefGoogle ScholarPubMed
Yates, A.M., Bonnan, M.F., Neveling, J., Chinsamy, A. & Blackbeard, M.G. 2000. A new transitional sauropodomorph dinosaur from the Early Jurassic of South Africa and the evolution of sauropod feeding and quadrupedalism. Proceedings of the Royal Society, Biological Science B 277: 787794.CrossRefGoogle Scholar
Yates, D.I., Earp, B.L., Levy, F. & Walker, E.S. 2006. Propagation of Sciadopitys verticillata (Thumb.) Sieb. & Zucc. by stem cuttings and properties of its latex-like sap. Horticultural Science 41: 16621666.Google Scholar
Yates, A.M., Bonnan, M.F. & Neveling, J. 2011. A new basal sauropod dinosaur from the early Jurassic of South Africa. Journal of Vertebrate Paleontology 31: 610625.CrossRefGoogle Scholar
Yeh, F.C. 1988. Isozyme variation of Thuja plicata (Cupressaceae) in British Columbia. Biochemical Systematics and Ecology 16: 373377.CrossRefGoogle Scholar
Yi, S., Saito, Y., Zhao, Q.-H. & Wang, P.-X. 2003. Vegetation and climate changes in the Cjangliang (Yangtze River) Delta, China, during the past 13,000 years inferred from pollen records. Quaternary Science Reviews 22: 15011519.CrossRefGoogle Scholar
Yin, A. & Harrison, M. (eds.). 2007. The Tectonic Evolution of Asia. Cambridge: Cambridge University Press.Google Scholar
York, R.A., Heald, R.C., Battles, J.J. & York, J.D. 2004. Group selection management in conifer forests: relationships between opening size and tree growth. Canadian Journal of Forest Research 34: 630641.CrossRefGoogle Scholar
Yoshida, N., Son, J., Matsushita, N., Iwamoto, K. & Hogetsu, T. 2014. Fine-scale distribution of ectomycorrhizal fungi colonising Tsuga diversifolia seedling growing on rocks in a subalpine Abies veitchii forest. Mycorrhiza 24: 247257.CrossRefGoogle Scholar
Young, A. (ed). 2000. Forest Conservation: Genetics. Collingwood: CSIRO Publishing.CrossRefGoogle Scholar
Yu, C.H. 1948. The wood structure of Metasequoia disticha. Botanical Bulletin Academica Sinica 2: 227230.Google Scholar
Yu, C.-H. 1956. The wood structure of TsugaKeteleeria with special reference to its taxonomy. Acta Botanica Sinica 5: 243248 (in Chinese).Google Scholar
Zachos, J.C., Dickens, G.R. & Zeebe, R.E. 2008. An early Cenozoic perspective on greenhouse warming and carbon-cycle dynamics. Nature 451: 279283.CrossRefGoogle Scholar
Zamuner, A.B., & Falaschi, P. 2005. Agathoxylon matildense n. sp., leno araucariaceao del Bosque Petrificado del Cerro Madre e Hilga Formacion La Matilde (Jurassico Medio), Provincia de Santa Cruz, Argentina. Ameghiniana 42: 339346.Google Scholar
Zangaroo, W, Bononi, V.L.R. & Truffen, S.B. 2000. Mycorrhizal dependency, inoculum potential and habit preference of native woody species in South Brazil. Journal of Tropical Ecology 16: 603622.CrossRefGoogle Scholar
Zanoni, T.A. 1980. Notes on Cupressus in Mexico. Boletín de la Sociedad Botánica de México 39: 128133.Google Scholar
Zanoni, T.A. 1999. Regional action plan: Caribbean conifers – current status. Pp 5962 in Farjon, A. & Page, C.N. (eds.), Conifers: Status Survey and Conifer Action Plan. Gland: IUCN.Google Scholar
Zanoni, T.A. & Adams, R.P. 1979. The genus Juniperus (Cupressaceae) in Mexico and Guatemala: synonymy, key, and distributions of the taxa. Boletín de la Sociedad Botánica de México 38: 83121.Google Scholar
Zanzi, A., Pelfini, M., Muttoni, G., Santilli, M. & Leonelli, G. 2007. Spectral analysis on mountain pine tree-ring chronologies. Dendrochronologia 24: 145154.CrossRefGoogle Scholar
Zastawniak, E. 1981. Tertiary leaf flora from the Point Henniquin Group of King George island (South Shetland Islands), Antarctica. Preliminary report. Geol. Sudet. (Polska) 72: 97108.Google Scholar
Zavarin, E.L., Smitii, L.V. & Bicho, J.G. 1967. Tropolones of Cupressaceae: III. Phytochemistry 6: 13871394.CrossRefGoogle Scholar
Zedler, P.H. 1986. Closed-cone conifers of the chaparral. Freemontia 14: 1417.Google Scholar
Zhang, H.J., Xie, M.S. & Yang, L.C. 1994. Study on the roughness coefficient of slope lace in the granitic region of Three-Gorges of the Yangtze River. Journal of Soil and Water Conservation 1: 3338.Google Scholar
Zhang, P., Zhou, P.-P., Jiang, C., Yu, H. & Yu, L-J. 2008. Screening of taxol-producing fungi based on PCR amplification from Taxus. Biotechnology Letters 30: 21192123.CrossRefGoogle ScholarPubMed
Zhang, Q., Chiang, T.Y., George, M., Liu, J.Q. & Abbott, R.J. 2005. Phylogeography of the Qinghai–Tibetan Plateau endemic Juniperus prewalskii (Cupressaceae) inferred from chloroplast DNA sequence variation. Molecular Ecology 14: 35133524.CrossRefGoogle ScholarPubMed
Zhang, Z.-Y., Yang, J.B. & Li, D.-Z. 2003. Phylogenetic relationship of an extremely endangered species, Pinus squamata (Pinaceae) inferred from four sequences of the chloroplast genome and ITS of the nuclear ribosomal DNA. Acta Botanica Sinica 45: 530535.Google Scholar
Zhao, L. 1980. Species nova generis Cupressi. Acta Phytotaxonomica Sinica 18: 210.Google Scholar
Zheng, W., Morris, E.K. & Rillig, M.C. 2014. Ectomycorrhizal fungi in association with Pinus sylvestris seedlings promote soil aggregation and soil water repellency. Soil Biology and Biochemistry 78: 326331.CrossRefGoogle Scholar
Zheng, Y.-Q. 2003. Progress of hybrid breeding of Pinus caribaea and strategy for future development. Forest Research 16: 110116.Google Scholar
Zheng, Z. 1991. Distribution of Dacrydium in south China during late Quaternary. Acta Botanica Sinica 33: 130139.Google Scholar
Zhilin, S. 1989. History of the development of the temperate forest flora in Kazakhstan from the Oligocene to the Early Miocene. Botanical Review 55: 205330.CrossRefGoogle Scholar
Zhou, K., Liang, X., Yan, F. & Ye, Y. 1983. Some remarks on the Nangou cold period based on the analysis of the pollen from the Nihewan stratum. Scientia Geologica Sinica 1: 8292.Google Scholar
Zhou, Z. 1983. Stalagmata samara, a new podocarpaceous conifer with monocolpate pollen from the Upper Triassic of Hunan, China. Paleontographica B 185: 5678.Google Scholar
Zhou, Z. 1987. Elatides harrisii, sp. nov., from the lower Cretaceous of Liaoning, China. Review of Palaeobotany and Palynology 131: 91103.Google Scholar
Zhou, Z. & Li, H. 1994. Early Tertiary gymnosperms from Fildes Peninsula, King George Island, Antarctica. Pp 191221 in Shen, Y. (ed.), Stratigraphy and Paleontology of Fildes Peninsula, King George Island, Antarctica. Beijing: Science Press.Google Scholar
Zhou, Z.Y., Zhang, B.L., Wang, Y.D. & Guignard, G. 2002. A new Karkenia (Ginkgoales) from the Upper Jurassic Yima formation, Henan, China and its megaspore membrane ultrastructure. Review of Palaeobotany and Palynology 120: 91105.CrossRefGoogle Scholar
Zhou, Z.Y., Quon, C. & Liu, Y.-S. 2012. Tertiary Ginkgo ovulate organs with assorted leaves from North Dakota, USA, and their evolutionary significance. International Journal of Plant Science 173: 6780.CrossRefGoogle Scholar
Ziegler, A.M., Rees, P.M., Rowley, D.B., et al. 1996. Mesozoic assembly of Asia: constraints from fossil floras, tectonics and paleomagmatism. Pp 371374 in Yin, A. & Harrison, M. (eds.), The Tectonic Evolution of Asia, Cambridge: Cambridge University Press.Google Scholar
Zimmer, E.A., Hamby, R.K., Arnold, M.R., Leblanc, M.L. & Theriot, C. 1989. Ribosomal RNA phylogenies and flowering plant evolution. Pp 195204 in Fernholm, B., Bremneer, K. & Jornvall, H. (eds.), The Hierarchy of Life. Amsterdam: Elsevier.Google Scholar
Zimmerman, D., Westhoff, M., Zimmermann, G., et al. 2007. Foliar water supply of tall trees: evidence for mucilage-facilitated moisture uptake from the atmosphere and the impact on pressure bomb measurements. Protoplasma 22: 1134.CrossRefGoogle Scholar
Zinke, P.J. & Stangenberger, A.G. 1994. Soil and nutrient element aspects of Sequoiadendron giganteum. Pp 6977 in Aune, P.S. (ed.), Proceedings of the Symposium on Giant Sequoias: Their Place in the Ecosystem and Society. Washington, DC: USDA.Google Scholar
Zinsmeister, W.J. 1987. Cretaceous paleogeography of Antarctica. Palaeogeography, Palaeoclimatology, Palaeoecology 59: 197206.CrossRefGoogle Scholar
Zobel, B.J. 1953. Are there natural loblolly–shortleaf pine hybrids? Journal of Forestry 51: 494502.Google Scholar
Zurawski, G. & Clegg, M.T. 1987. Evolution of higher plant chloroplast DNA-coded genes: implications for structure–function and phylogenetic studies. Annual Review of Plant Physiology 38: 391418.CrossRefGoogle Scholar

Save book to Kindle

To save this book to your Kindle, first ensure no-reply@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.

  • Selected Bibliography
  • Christopher N. Page, University of Exeter
  • Book: Evolution of the Arborescent Gymnosperms
  • Online publication: 11 November 2024
  • Chapter DOI: https://doi.org/10.1017/9781009263108.039
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.

  • Selected Bibliography
  • Christopher N. Page, University of Exeter
  • Book: Evolution of the Arborescent Gymnosperms
  • Online publication: 11 November 2024
  • Chapter DOI: https://doi.org/10.1017/9781009263108.039
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.

  • Selected Bibliography
  • Christopher N. Page, University of Exeter
  • Book: Evolution of the Arborescent Gymnosperms
  • Online publication: 11 November 2024
  • Chapter DOI: https://doi.org/10.1017/9781009263108.039
Available formats
×