Hostname: page-component-848d4c4894-pftt2 Total loading time: 0 Render date: 2024-05-06T15:00:51.033Z Has data issue: false hasContentIssue false

Dark and disturbed: a new image of early angiosperm ecology

Published online by Cambridge University Press:  08 April 2016

Taylor S. Feild
Affiliation:
Department of Botany, University of Toronto, Toronto, Ontario M5S 3B2, Canada. E-mail: feild@botany.utoronto.ca
Nan Crystal Arens
Affiliation:
Department of Geoscience, Hobart and William Smith Colleges, Geneva, New York 14456
James A. Doyle
Affiliation:
Section of Evolution and Ecology, University of California, Davis, California 95616
Todd E. Dawson
Affiliation:
Department of Integrative Biology, University of California, Berkeley, California 94720-3140
Michael J. Donoghue
Affiliation:
Department of Ecology and Evolution, Yale University, New Haven, Connecticut 06520

Abstract

Better understanding of the functional biology of early angiosperms may clarify ecological factors surrounding their origin and early radiation. Phylogenetic studies identify Amborella, Nymphaeales (water lilies), Austrobaileyales, and Chloranthaceae as extant lineages that branched before the radiation of core angiosperms. Among living plants, these lineages may represent the best models for the ecology and physiology of early angiosperms. Here we combine phylogenetic reconstruction with new data on the morphology and ecophysiology of these plants to infer early angiosperm function. With few exceptions, Amborella, Austrobaileyales, and Chloranthaceae share ecophysiological traits associated with shady, disturbed, and wet habitats. These features include low and easily light-saturated photosynthetic rates, leaf anatomy related to the capture of understory light, small seed size, and clonal reproduction. Some Chloranthaceae, however, possess higher photosynthetic capacities and seedlings that recruit in canopy gaps and other sunny, disturbed habitats, which may have allowed Cretaceous Chloranthaceae to expand into more diverse environments. In contrast, water lilies possess ecophysiological features linked to aquatic, sunny habitats, such as absence of a vascular cambium, ventilating stems and roots, and floating leaves tuned for high photosynthetic rates in full sun. Nymphaeales may represent an early radiation into such aquatic environments. We hypothesize that the earliest angiosperms were woody plants that grew in dimly lit, disturbed forest understory habitats and/or shady streamside settings. This ecology may have restricted the diversity of pre-Aptian angiosperms and living basal lineages. The vegetative flexibility that evolved in the understory, however, may have been a key factor in their diversification in other habitats. Our inferences based on living plants are consistent with many aspects of the Early Cretaceous fossil record and can be tested with further study of the anatomy, chemistry, and sedimentological context of Early Cretaceous angiosperm fossils.

Type
Articles
Copyright
Copyright © The Paleontological Society 

Access options

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

References

Literature Cited

Ackerly, D. D., and Donoghue, M. J. 1998. Leaf size, sapling allometry, and Corner's rules: phylogeny and correlated evolution in maples (Acer). American Naturalist 152:767791.Google Scholar
Alvin, K. L. 1974. Leaf anatomy of Weichselia based on fusainized material. Palaeontology 17:587598.Google Scholar
Arber, E. A. N., and Parkin, J. 1907. On the origin of angiosperms. Journal of the Linnean Society (Botany) 38:2980.Google Scholar
Archangelsky, S., and Gamerro, J. C. 1967. Spore and pollen types of the Lower Cretaceous in Patagonia (Argentina). Review of Palaeobotany and Palynology 1:211217.CrossRefGoogle Scholar
Archangelsky, S., and Taylor, T. N. 1993. The ultrastructure of in situ Clavatipollenites pollen from the Early Cretaceous of Patagonia. American Journal of Botany 80:879885.Google Scholar
Arens, N. C. 1997. Responses of leaf anatomy to light environment in the tree fern Cyathea caracasana (Cyatheaceae) and its application to some ancient seed ferns. Palaios 12:8494.CrossRefGoogle Scholar
Arens, N. C. 2001. Variation in performance of the tree fern Cyathea caracasana (Cyatheaceae) across the successional mosaic of an Andean cloud forest. American Journal of Botany 88:545551.Google Scholar
Arens, N. C., and Jahren, A. H. 2000. Carbon isotope excursion in atmospheric CO2 at the Cretaceous-Tertiary boundary: evidence from terrestrial sediments. Palaios 15:314322.Google Scholar
Arens, N. C., and Jahren, A. H. 2002. Chemostratigraphic correlation of four fossil-bearing sections in southwestern North Dakota. Geological Society of America Special Paper 361:7593.Google Scholar
Arens, N. C., Jahren, A. H., and Amundson, R. 2000. Can C3 plants faithfully record the carbon isotopic composition of atmospheric carbon dioxide? Paleobiology 26:137164.2.0.CO;2>CrossRefGoogle Scholar
Axelrod, D. I. 1952. A theory of angiosperm evolution. Evolution 6:2960.Google Scholar
Axelrod, D. I. 1970. Mesozoic paleogeography and early angiosperm history. Botanical Review 36:277319.CrossRefGoogle Scholar
Bakker, R. T. 1978. Dinosaur feeding behaviour and the origin of flowering plants. Nature 274:661663.Google Scholar
Barkman, T. J., Chenery, G., McNeal, J. R., Lyons-Weiler, J., Ellisens, W. J., Moore, G., Wolfe, A. D., and de Pamphilis, C. W. 2000. Independent and combined analyses of sequences from all three genomic compartments converge on the root of flowering plant phylogeny. Proceedings of the National Academy of Sciences USA 97:1316613171.Google Scholar
Barrat-Segretain, M.-H. 1996. Germination and colonization dynamics of Nuphar lutea (L.) Sm. in a former river channel. Aquatic Botany 55:3138.Google Scholar
Barrett, P. M., and Willis, K. J. 2001. Did dinosaurs invent flowers? Dinosaur-angiosperm coevolution revisited. Biological Reviews 76:411447.CrossRefGoogle ScholarPubMed
Barron, E. J., Fawcett, P. J., Pollard, D., and Thompson, S. L. 1994. Model simulations of Cretaceous climates: the role of geography and carbon dioxide. Pp. 99108in Allen, J. R. L., Hoskins, B. J., Sellwood, B. W., Spicer, R. A., and Valdes, P. J., eds. Palaeoclimates and their modelling: with special reference to the Mesozoic Era. Chapman and Hall, London.CrossRefGoogle Scholar
Beerling, D. J., and Royer, D. L. 2002. Fossil plants as indicators of the Phanerozoic global carbon cycle. Annual Review of Earth and Planetary Sciences 30:527556.Google Scholar
Beerling, D. J., and Woodward, F. I. 2001. Vegetation and the terrestrial carbon cycle: modelling the first 400 million years. Cambridge University Press, Cambridge.Google Scholar
Behrensmeyer, A. K., Kidwell, S. M., and Gastaldo, R. A. 2000. Taphonomy and paleobiology. Paleobiology 26:103147.Google Scholar
Bernhardt, P., Sage, T. L., Weston, P., Azuma, H., Lam, M., Thien, L. B., and Bruhl, J. 2003. The pollination of Trimenia moorei (Trimeniaceae): floral volatiles, insect/wind pollen vectors, and stigmatic self-incompatibility in a basal angiosperm. Annals of Botany 92:445458.Google Scholar
Berry, E. W. 1911. Systematic paleontology, Lower Cretaceous, Pteridophyta-Dicotyledonae. Pp. 214508in Clark, W. B., ed. Maryland Geological Survey, Lower Cretaceous. Johns Hopkins University Press, Baltimore.Google Scholar
Bews, J. W. 1927. Studies in the ecological evolution of angiosperms. New Phytologist 26:121.CrossRefGoogle Scholar
Bilger, W., Schreiber, U., and Bock, M. 1995. Determinants of the quantum efficiency of photosystem II and nonphotochemical quenching of chlorophyll fluorescence in the field. Oecologia 102:425432.Google Scholar
Björkman, O., and Demmig, B. 1987. Photon yield of O2 evolution and chlorophyll fluorescence characteristics at 77 K among vascular plants of diverse origins. Planta 170:489504.Google Scholar
Blanc, P. 1986. Edification d'arbres par croissance d'établissement de type monocotylédonien: l'exemple de Chloranthaceae. Pp. 101123in Colloque international sur l'arbre 1986. Naturalia Monspeliensia, numéro hors série, Montpellier.Google Scholar
Bond, W. J., and Midgely, J. J. 1988. Allometry and sexual differences in leaf size. American Naturalist 131:901910.Google Scholar
Bond, W. J., and Midgely, J. J. 2001. Ecology of sprouting in woody plants: the persistence niche. Trends in Ecology and Evolution 16:4551.Google Scholar
Bond, W. J., and Midgely, J. J. 2003. The evolutionary ecology of sprouting. International Journal of Plant Sciences 164(Suppl.):S103S114.Google Scholar
Bongers, F., and Popma, J. 1990. Leaf characteristics of the tropical rain forest flora of Los Tuxtlas, Mexico. Botanical Gazette 151:354365.Google Scholar
Brenner, G. J. 1963. The spores and pollen of the Potomac Group of Maryland. Maryland Department of Geology, Mines and Water Resources Bulletin 27:1215.Google Scholar
Brenner, G. J. 1976. Middle Cretaceous floral provinces and early migrations of angiosperms. Pp. 2347in Beck, C. B., ed. Origin and early evolution of angiosperms. Columbia University Press, New York.Google Scholar
Brenner, G. J. 1996. Evidence for the earliest stage of angiosperm pollen evolution: a paleoequatorial section from Israel. Pp. 91115in Taylor, D. W. and Hickey, L. J., eds. Flowering plant origin, evolution and phylogeny. Chapman and Hall, New York.CrossRefGoogle Scholar
Brodribb, T. J., and Feild, T. S. 2000. Stem hydraulic supply is linked to leaf photosynthetic capacity: evidence from New Caledonian and Tasmanian rainforests. Plant, Cell and Environment 23:13811388.Google Scholar
Brodribb, T., and Hill, R. S. 1997. Light response characteristics of a morphologically diverse group of southern hemisphere conifers as measured by chlorophyll fluorescence. Oecologia 110:1017.Google Scholar
Brodribb, T. J., Holbrook, N. M., and Gutierrez, M. V. 2002. Hydraulic and photosynthetic coordination in seasonally dry tropical forest. Plant, Cell and Environment 25:14351444.CrossRefGoogle Scholar
Burger, D. 1990. Early Cretaceous angiosperms from Queensland, Australia. Review of Palaeobotany and Palynology 65:153163.CrossRefGoogle Scholar
Burger, D. 1993. Early and middle Cretaceous angiosperm pollen grains from Australia. Review of Palaeobotany and Palynology 78:183234.CrossRefGoogle Scholar
Calder, J. H., Gibling, M. R., Eble, C. F., Scott, A. C., and MacNeil, D. J. 1996. The Westphalian D fossil lepidodendrid forest at Table Head, Sydney Basin, Nova Scotia: sedimentology, paleoecology and floral response to changing edaphic conditions. International Journal of Coal Geology 31:277313.Google Scholar
Canham, C. D., Kobe, R. K., Latty, E. F., and Chazdon, R. L. 1999. Interspecific and intraspecific variation in tree seedling survival: effects of allocation to roots versus carbohydrate reserves. Oecologia 121:111.CrossRefGoogle ScholarPubMed
Cantrill, D. J., and Nichols, G. J. 1996. Taxonomy and paleoecology of Early Cretaceous (Late Albian) angiosperm leaves from Alexander Island, Antarctica. Review of Palaeobotany and Palynology 92:128.Google Scholar
Carlquist, S. 1975. Ecological strategies of xylem evolution. University of California Press, Berkeley.Google Scholar
Carlquist, S. 1984. Wood anatomy of Trimeniaceae. Plant Systematics and Evolution 144:103118.Google Scholar
Carlquist, S. 1987. Presence of vessels in wood of Sarcandra (Chloranthaceae): comments on vessel origin in angiosperms. American Journal of Botany 74:17651771.Google Scholar
Carlquist, S. 1992a. Wood anatomy and stem of Chloranthus; summary of wood anatomy of Chloranthaceae, with comments on relationships, vessellessness, and the origin of monocotyledons. International Association of Wood Anatomists Bulletin 13:316.Google Scholar
Carlquist, S. 1992b. Wood anatomy of Hedyosmum (Chloranthaceae) and the tracheid-vessel element transition. Aliso 13:447462.Google Scholar
Carlquist, S. 1999. Wood and bark anatomy of Schisandraceae: implications for phylogeny, habit, and vessel evolution. Aliso 18:4555.Google Scholar
Carlquist, S., and Schneider, E. L. 2002. The tracheid-vessel element transition in angiosperms involves multiple independent features: cladistic consequences. American Journal of Botany 89:185195.Google Scholar
Chandrasekharam, A. 1972. Spongy mesophyll remains in fossil leaf compressions. Science 177:354356.Google Scholar
Chazdon, R. L. 1988. Sunflecks and their importance to forest understory plants. Advances in Ecological Research 18:163.Google Scholar
Chazdon, R. L., Pearcy, R. W., Lee, D. W., and Fetcher, N. 1996. Photosynthetic responses of tropical forest plants to contrasting light environments. Pp. 555in Mulkey, S. S., Chazdon, R. L., and Smith, A. P., eds. Tropical forest plant ecophysiology. Chapman and Hall, London.CrossRefGoogle Scholar
Clark, D. B., and Clark, D. A. 1988. Leaf production and the cost of reproduction in the neotropical rain forest cycad, Zamia skinneri. Journal of Ecology 76:11531163.Google Scholar
Clark, D. B., Clark, D. A., and Grayum, M. H. 1992. Leaf demography of a neotropical rain forest cycad, Zamia skinneri (Zamiaceae). American Journal of Botany 79:2833.Google Scholar
Condon, A. G., Richards, R. A., and Farquhar, G. D. 1993. Relationships between carbon isotopic discrimination, water use efficiency, and transpiration efficiency for dryland wheat. Australian Journal of Agricultural Research 44:16931711.CrossRefGoogle Scholar
Cook, R. E. 1983. Clonal plant populations. American Scientist 71:244253.Google Scholar
Crane, P. R. 1985. Phylogenetic analysis of seed plants and the origin of angiosperms. Annals of the Missouri Botanical Garden 72:716793.Google Scholar
Crane, P. R. 1987. Vegetational consequences of the angiosperm diversification. Pp. 107144in Friis, E. M., Chaloner, W. G., and Crane, P. R., eds. The origins of angiosperms and their biological consequences. Cambridge University Press, Cambridge.Google Scholar
Crane, P. R., and Lidgard, S. 1989. Angiosperm diversification and paleolatitudinal gradients in Cretaceous floristic diversity. Science 246:675678.Google Scholar
Crane, P. R., and Upchurch, G. R. 1987. Drewria potomacensis gen. et sp. nov., an Early Cretaceous member of Gnetales from the Potomac Group of Virginia. American Journal of Botany 74:17221736.Google Scholar
Crane, P. R., Pedersen, K. R., Friis, E. M., and Drinnan, A. N. 1993. Early Cretaceous (early to middle Albian) platanoid inflorescences associated with Sapindopsis leaves from the Potomac Group of eastern North America. Systematic Botany 18:328344.Google Scholar
Crane, P. R., Friis, E. M., and Pedersen, K. R. 1994. Paleobotanical evidence on the early radiation of magnoliid angiosperms. Plant Systematics and Evolution 8(Suppl.):5172.Google Scholar
Crane, P. R., Friis, E. M., and Pedersen, K. R. 1995. The origin and early diversification of angiosperms. Nature 374:2733.Google Scholar
Cronquist, A. 1988. The evolution and classification of flowering plants, 2d ed.New York Botanical Garden, New York.Google Scholar
Dacey, J. W. H. 1980. Internal winds in water lilies: an adaptation for life in anaerobic sediments. Science 210:10171019.Google Scholar
Dai, N., Kenji, S., and Sakai, A. 2002. Seedling establishment of deciduous trees in various topographic positions. Journal of Vegetation Science 13:3544.Google Scholar
Davies-Vollum, K. S. 1999. The formation of beds underlying carbonaceous shales in aquic paleosols: examples from the Bighorn Basin of Wyoming. International Journal of Coal Geology 41:239255.CrossRefGoogle Scholar
Davies-Vollum, K. S., and Wing, S. L. 1998. Sedimentological, taphonomic, and climatic aspects of Eocene swamp deposits (Willwood Formation, Bighorn Basin, Wyoming). Palaios 13:2840.Google Scholar
Demmig-Adams, B., Adams, W. W., Logan, B., and Verhoeven, A. 1995. Xanthophyll cycle dependent energy dissipation and flexible photosystem II efficiency in plants acclimated to light stress. Australian Journal of Plant Physiology 22:249260.Google Scholar
Dettmann, M. E. 1994. Cretaceous vegetation: the microfossil record. Pp. 143188in Hill, R. S., ed. History of the Australian vegetation. Cambridge University Press, Cambridge.Google Scholar
Donoghue, M. J., and Doyle, J. A. 1989. Phylogenetic analysis of angiosperms and the relationships of Hamamelidae. Pp. 1445in Crane, P. R. and Blackmore, S., eds. Evolution, systematics, and fossil history of the Hamamelidae, Vol. 1. Clarendon, Oxford.Google Scholar
Doyle, J. A. 1969. Cretaceous angiosperm pollen of the Atlantic Coastal Plain and its evolutionary significance. Journal of the Arnold Arboretum 50:135.Google Scholar
Doyle, J. A. 1977. Patterns of evolution in early angiosperms. Pp. 501546in Hallam, A., ed. Patterns of evolution. Elsevier, Amsterdam.Google Scholar
Doyle, J. A. 1978. Origin of angiosperms. Annual Review of Ecology and Systematics 9:365392.Google Scholar
Doyle, J. A. 1999. The rise of angiosperms as seen in the African Cretaceous pollen record. in Heine, K., ed. Palaeoecology of Africa 26:329.Google Scholar
Doyle, J. A. 2001. Significance of molecular phylogenetic analyses for paleobotanical investigations on the origin of angiosperms. Palaeobotanist 50:167188.Google Scholar
Doyle, J. A., and Donoghue, M. J. 1986. Seed plant phylogeny and the origin of angiosperms: an experimental cladistic approach. Botanical Review 52:321431.CrossRefGoogle Scholar
Doyle, J. A., and Donoghue, M. J. 1993. Phylogenies and angiosperm diversification. Paleobiology 19:141167.Google Scholar
Doyle, J. A. and Endress, P. K. 2000. Morphological phylogenetic analysis of basal angiosperms: comparison and combination with molecular data. International Journal of Plant Sciences 161(Suppl.):S121S153.Google Scholar
Doyle, J. A., and Hickey, L. J. 1976. Pollen and leaves from the mid-Cretaceous Potomac Group and their bearing on early angiosperm evolution. Pp. 139206in Beck, C. B., ed. Origin and early evolution of angiosperms. Columbia University Press, New York.Google Scholar
Doyle, J. A., Biens, P., Doerenkamp, A., and Jardiné, S. 1977. Angiosperm pollen from the pre-Albian Cretaceous of Equatorial Africa. Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine 1:451473.Google Scholar
Doyle, J. A., Jardiné, S., and Doerenkamp, A. 1982. Afropollis, a new genus of early angiosperm pollen, with notes on the Cretaceous palynostratigraphy and paleoenvironments of Northern Gondwana. Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine 6:39117.Google Scholar
Doyle, J. A., Donoghue, M. J., and Zimmer, E. A. 1994. Integration of morphological and ribosomal RNA data on the origin of angiosperms. Annals of the Missouri Botanical Garden 81:419450.Google Scholar
Doyle, J. A., Eklund, H., and Herendeen, P. S. 2003. Floral evolution in Chloranthaceae: implications of a phylogenetic analysis of morphological characters. International Journal of Plant Sciences 164(Suppl.):S365S382.CrossRefGoogle Scholar
Edwards, G. E., and Baker, N. R. 1993. Can assimilation in maize leaves be predicted accurately from chlorophyll fluorescence analysis? Photosynthesis Research 37:89102.CrossRefGoogle ScholarPubMed
Eklund, H., Doyle, J. A., and Herendeen, P. S. 2004. Morphological phylogenetic analysis of living and fossil Chloranthaceae. International Journal of Plant Sciences (in press).Google Scholar
Endress, P. K. 1987. The Chloranthaceae: reproductive structures and phylogenetic position. Botanische Jahrbücher für Systematik 109:153226.Google Scholar
Endress, P. K. 2001. The flowers in extant basal angiosperms and inferences on ancestral flowers. International Journal of Plant Sciences 162:11111140.Google Scholar
Eriksson, O., Friis, E. M., and Crane, P. R. 2000. Seed size, fruit size, and dispersal systems in angiosperms from the Early Cretaceous to the Late Tertiary. American Naturalist 156:4758.Google Scholar
Esau, K. 1977. Anatomy of seed plants. Wiley, New York.Google Scholar
Farquhar, G. D., O'Leary, M. H., and Berry, J. A. 1982. On the relationship between carbon isotopic discrimination and the intercellular carbon dioxide concentration in leaves. Australian Journal of Plant Physiology 9:121137.Google Scholar
Feild, T. S., Zwieniecki, M. A., Donoghue, M. J., and Holbrook, N. M. 1998. Stomatal plugs of Drimys winteri (Winteraceae) protect leaves from mist but not drought. Proceedings of the National Academy of Sciences USA 95:1425614259.Google Scholar
Feild, T. S., Brodribb, T., Jaffré, T., and Holbrook, N. M. 2001. Responses of leaf anatomy, photosynthetic light-use, and xylem hydraulics to light in the basal angiosperm Amborella trichopoda (Amborellaceae). International Journal of Plant Sciences 162:9991008.Google Scholar
Feild, T. S., Zwieniecki, M. A., Brodribb, T., Jaffré, T., Donoghue, M. J., and Holbrook, N. M. 2000. Structure and function of tracheary elements in Amborella trichopoda. International Journal of Plant Sciences 161:705712.Google Scholar
Feild, T. S., Franks, P. J., and Sage, T. L. 2003a. The shade-adapted ecophysiology of Austrobaileya scandens. International Journal of Plant Sciences 164:313324.Google Scholar
Feild, T. S., Arens, N. C., and Dawson, T. E. 2003b. The ancestral ecology of flowering plants: emerging perspectives from extant basal lineages. International Journal of Plant Sciences 164(Suppl.):S129S142.Google Scholar
Fontaine, W. M. 1889. The Potomac or Younger Mesozoic flora. U.S. Geological Survey Monograph 15.Google Scholar
Forbis, T. A., Floyd, S. K., and de Queiroz, A. 2002. The evolution of embryo size in angiosperms and other seed plants: implications for the evolution of seed dormancy. Evolution 56:21122125.Google Scholar
Franco, A. C., and Lüttge, U. 2002. Midday depression in savanna trees: coordinated adjustments in photochemical efficiency, photorespiration, CO2 assimilation, and water-use efficiency. Oecologia 131:356365.Google Scholar
Friis, E. M., Crane, P. R., and Pedersen, K. R. 1997. Fossil history of magnoliid angiosperms. Pp. 121156in Iwatsuki, K. and Raven, P. R., eds. Evolution and diversification of land plants. Springer, Tokyo.Google Scholar
Friis, E. M., Pedersen, K. R., and Crane, P. R. 1999. Early angiosperm diversification: the diversity of pollen associated with angiosperm reproductive structures in Early Cretaceous floras from Portugal. Annals of the Missouri Botanical Garden 86:259296.Google Scholar
Friis, E. M., Pedersen, K. R., and Crane, P. R. 2000. Reproductive structure and organization of basal angiosperms from the Early Cretaceous (Barremian or Aptian) of western Portugal. International Journal of Plant Sciences 161(Suppl.):S169S182.Google Scholar
Friis, E. M., Pedersen, K. R., and Crane, P. R. 2001. Fossil evidence of water lilies (Nymphaeales) in the Early Cretaceous. Nature 410:357360.Google Scholar
Friis, E. M., Doyle, J. A., Endress, P. K., and Leng, Q. 2003. Archaefructus: angiosperm precursor or specialized early angiosperm? Trends in Plant Science 8:369373.Google Scholar
Gandolfo, M. A., Nixon, K. C., and Crepet, W. L. 2001. Turonian Pinaceae of the Raritan Formation, New Jersey. Plant Systematics and Evolution 226:187203.Google Scholar
Gavin, D. G., and Peart, D. R. 1999. Vegetative life history of a dominant rain forest canopy tree. Biotropica 31:288294.Google Scholar
Genty, B., Briantais, J.-M., and Baker, N. R. 1989. The relationship between the quantum yield of photosynthetic electron transport rate and quenching of chlorophyll fluorescence. Biochimica Biophysica Acta 990:8792.Google Scholar
Givnish, T. J. 1979. On the adaptive significance of leaf form. Pp. 375407in Solbrig, O. T., Jain, S., Johnson, G. B., and Raven, P. H., eds. Topics in plant population biology. Columbia University Press, New York.Google Scholar
Givnish, T. J. 1988. Adaptation to sun and shade: a whole-plant perspective. Australian Journal of Plant Physiology 15:6392.Google Scholar
Gottsberger, G. 1988. The reproductive biology of primitive angiosperms. Taxon 37:630643.Google Scholar
Graham, S. W., and Olmstead, R. G. 2000. Utility of 17 chloroplast genes for inferring the phylogeny of the basal angiosperms. American Journal of Botany 87:17121730.CrossRefGoogle ScholarPubMed
Greig, N. 1993. Regeneration mode in neotropical Piper: habitat and species comparisons. Ecology 74:21252135.Google Scholar
Grime, J. P. 1979. Plant strategies and vegetation processes. Wiley, Chichester, U.K.Google Scholar
Grubb, P. J. 1998. Seeds and fruits of tropical rainforest plants: interpretation of the range in seed size, degree of defense, and flesh/seed quotients. Pp. 124in Newbery, D. M., Prins, H. H. T., and Brown, N. D., eds. Dynamics of tropical communities. Blackwell Scientific, Oxford.Google Scholar
Hao, G., Saunders, R. M. K., and Chye, M.-L. 2000. A phylogenetic analysis of the Illiciaceae based on sequences of internal transcribed spacers (ITS) of nuclear ribosomal DNA. Plant Systematics and Evolution 223:8190.Google Scholar
Hao, G., Chye, M.-L., and Saunders, R. M. K. 2001. A phylogenetic analysis of the Schisandraceae based on morphology and nuclear ribosomal ITS sequences. Botanical Journal of the Linnean Society 135:401411.Google Scholar
Harper, J. L., Lovell, P. H., and Moore, K. G. 1970. The shapes and sizes of seeds. Annual Review of Ecology and Systematics 1:327356.Google Scholar
Hickey, L. J., and Doyle, J. A. 1977. Early Cretaceous fossil evidence for angiosperm evolution. Botanical Review 43:1104.Google Scholar
Hickman, J. C. 1993. The Jepson manual of higher plants of California. University of California Press, Berkeley.Google Scholar
Hughes, N. F. 1994. The enigma of angiosperm origins. Cambridge University Press, Cambridge.Google Scholar
Kelly, C. K. 1995. Seed size in tropical trees: a comparative study of factors affecting seed size in Peruvian angiosperms. Oecologia 102:377388.Google Scholar
Kong, H.-Z. 2001. Comparative morphology of leaf epidermis in the Chloranthaceae. Botanical Journal of the Linnean Society 136:281296.Google Scholar
Kong, H.-Z., Chen, Z.-D., and Lu, A.-M. 2002. Phylogeny of Chloranthus (Chloranthaceae) based on nuclear ribosomal ITS and plastid trnL-F sequence data. American Journal of Botany 89:940946.Google Scholar
Kott, L. S., and Britton, D. M. 1985. Role of morphological characteristics of leaves and the sporangial region in the taxonomy of Isoetes in northeastern North America. American Fern Journal 75:4455.Google Scholar
Krause, G. H., and Weis, E. 1991. Chlorophyll fluorescence and photosynthesis: the basics. Annual Review of Plant Physiology and Plant Molecular Biology 42:313349.Google Scholar
Kwit, C., Swartz, M. W., Platt, W. J., and Geaghen, J. P. 1998. The distribution of tree species in steepheads of the Apalachicola River Bluffs, Florida. Journal of the Torrey Botanical Society 125:309318.Google Scholar
Leishman, M. R., Wright, I., Moles, A. T., and Westoby, M. 2000. The evolutionary ecology of seed size. Pp. 3157in Fenner, M., ed. Seeds: the ecology of regeneration in plant communities. CABI Publishing, Wallingford, U.K.Google Scholar
Les, D. H., Schneider, E. L., Padgett, D. J., Solis, P. S., Soltis, D. E., and Zanis, M. 1999. Phylogeny, classification, and floral evolution of water lilies (Nymphaeaceae: Nymphaeales): a synthesis of non-molecular, rbcL, matK, and 18S rDNA data. Systematic Botany 24:2846.Google Scholar
Luo, Y.-B., and Li, Z.-Y. 1999. Pollination ecology of Chloranthus serratus (Thunb.) Roem. et Schult. and Ch. fortunei (A. Gray) Solms-Laub. (Chloranthaceae). Annals of Botany 83:489499.Google Scholar
Lupia, R., Lidgard, S., and Crane, P. R. 1999. Comparing palynological abundance and diversity: implications for biotic replacement during the Cretaceous angiosperm radiation. Paleobiology 25:305340.Google Scholar
Lusk, C. K. 1995. Seed size, establishment sites, and species coexistence in a Chilean rainforest. Journal of Vegetation Science 6:249256.Google Scholar
Lusk, C. K., and Kelly, C. K. 2003. Interspecific variation in seed size and safe sites in a temperate rain forest. New Phytologist 158:535541.CrossRefGoogle Scholar
Maddison, W. P. 1991. Squared-change parsimony reconstructions of ancestral states for continuous-valued characters on a phylogenetic tree. Systematic Zoology 40:305314.Google Scholar
Maddison, W. P., and Maddison, D. R. 2001. MacClade 4: analysis of phylogeny and character evolution, Version 4.03. Sinauer, Sunderland, Mass.Google Scholar
Magallón, S., and Sanderson, M. J. 2001. Absolute diversification rates in angiosperm clades. Evolution 55:17621780.Google Scholar
Martin, T. J., and Ogden, J. 2002. The seed ecology of Ascarina lucida: a rare New Zealand tree adapted to disturbance. New Zealand Journal of Botany 40:397404.Google Scholar
Mathews, S., and Donoghue, M. J. 1999. The root of angiosperm phylogeny inferred from duplicate phytochrome genes. Science 286:947950.Google Scholar
McElwain, J. C., and Chaloner, W. G. 1996. The fossil cuticle as a skeletal record of environment change. Palaios 11:376388.Google Scholar
Metcalfe, C. R. 1987. Anatomy of the dicotyledons, 2d ed., Vol. III. Magnoliales, Illiciales, and Laurales (sensu Armen Takhtajan). Clarendon, Oxford.Google Scholar
Mohr, B. A. R., and Friis, E. M. 2000. Early angiosperms from the Lower Cretaceous Crato Formation (Brazil), a preliminary report. International Journal of Plant Sciences 161(Suppl.):S155S167.Google Scholar
Morse, S. R. 1990. Water balance in Hemizonia luzulifolia: the role of extracellular polysaccharides. Plant, Cell and Environment 13:3948.Google Scholar
Nguyen Tu, T. T., Bocherens, H., Mariotti, A., Baudin, F., Pons, D., Broutin, J., Derenne, S., and Largeau, C. 1999. Ecological distribution of Cenomanian terrestrial plants based on 13C/12C ratios. Palaeogeography, Palaeoclimatology, Palaeoecology 145:7993.Google Scholar
Niinemets, U., and Kalevi, K. 1994. Leaf weight per area and leaf size of 85 Estonian woody species in relation to shade tolerance and light availability. Forest Ecology and Management 70:110.Google Scholar
Niklas, K. J. 1997. The evolutionary biology of plants. University of Chicago Press, Chicago.Google Scholar
Nixon, K. C., Crepet, W. L., Stevenson, D., and Friis, E. M. 1994. A reevaluation of seed plant phylogeny. Annals of the Missouri Botanical Garden 81:484533.Google Scholar
Parkinson, C. L., Adams, K. L., and Palmer, J. D. 1999. Multigene analyses identify the three earliest lineages of extant flowering plants. Current Biology 9:14851488.Google Scholar
Parrish, J. T. 1987. Global palaeogeography and palaeoclimate of the Late Cretaceous and Early Tertiary. Pp. 5173in Friis, E. M., Chaloner, W. G., and Crane, P. R., eds. The origins of angiosperms and their biological consequences. Cambridge University Press, Cambridge.Google Scholar
Parrish, J. T., Ziegler, A. M., and Scotese, C. R. 1982. Rainfall patterns and the distribution of coals and evaporites in the Mesozoic and Cenozoic. Palaeogeography Palaeoclimatology Palaeoecology 40:67101.Google Scholar
Pearcy, R. W. 1990. Sunflecks and photosynthesis in plant canopies. Annual Review of Plant Physiology and Plant Molecular Biology 41:421453.Google Scholar
Philipson, W. R. 1986. Trimeniaceae. Flora Malesiana Series I 10:327333.Google Scholar
Playford, G. 1971. Palynology of Lower Cretaceous (Swan River) strata of Saskatchewan and Manitoba. Palaeontology 14:533565.Google Scholar
Pons, D. 1984. Le Mésozoïque de Colombie: macroflores et microflores. Editions du CNRS, Paris.Google Scholar
Qiu, Y.-L., Lee, J., Bernasconi-Quadroni, F., Soltis, D. E., Soltis, P. S., Zanis, M., Zimmer, E. A., Chen, Z., Savolainen, V., and Chase, M. W. 1999. The earliest angiosperms: evidence from mitochondrial, plastid and nuclear genomes. Nature 402:404407.Google Scholar
Rascher, U., Liebig, M., and Lüttge, U. 2000. Evaluation of instant light-response curves of chlorophyll fluorescence parameters obtained with a portable chlorophyll fluorometer on site in the field. Plant, Cell and Environment 23:13971405.Google Scholar
Rees, P. M., Ziegler, A. M., and Valdes, P. J. 2000. Jurassic phytogeography and climates: new data and model comparisons. Pp. 297318in Huber, B. T., MacLeod, K. G., and Wing, S. L., eds. Warm climates in earth history. Cambridge University Press, Cambridge.Google Scholar
Roberts, M. L., and Haynes, R. R. 1983. Ballistic seed dispersal in Illicium (Illiciaceae). Plant Systematics and Evolution 143:227232.Google Scholar
Rodenburg, W. F. 1971. A revision of the genus Trimenia (Trimeniaceae). Blumea 19:315.Google Scholar
Romero, E. J., and Archangelsky, S. 1986. Early Cretaceous angiosperm leaves from southern South America. Science 234:15801582.Google Scholar
Roth-Nebelsick, A., Uhl, D., Mosbrugger, V., and Kerp, H. 2001. Evolution and function of leaf venation architecture: a review. Annals of Botany 87:553566.Google Scholar
Royer, D. L., Hickey, L. J., and Wing, S. L. 2003. Ecological conservatism in the “living fossil” Ginkgo. Paleobiology 29:84104.Google Scholar
Sakai, A., Ohsawa, T., and Ohsawa, M. 1995. Adaptive significance of sprouting of Euptelea polyandra, a deciduous tree growing on steep slopes with shallow soil. Journal of Plant Research 108:377386.Google Scholar
Salisbury, E. J. 1942. The reproductive capacity of plants: studies in quantitative biology. G. Bell, London.Google Scholar
Sanderson, M. J., and Donoghue, M. J. 1994. Shifts in diversification rate with the origin of angiosperms. Science 264:15901593.Google Scholar
Saunders, R. M. K. 1998. Monograph of Kadsura (Schisandraceae). Systematic Botany Monographs 54:1106.Google Scholar
Saunders, R. M. K. 2000. Monograph of Schisandra (Schisandraceae). Systematic Botany Monographs 58:1146.Google Scholar
Schabilion, J. T., and Reihman, M. A. 1985. Anatomy of petrified Neuropteris scheuchzeri pinnules from the Middle Pennsylvanian of Iowa: a paleoecological interpretation. Neuvième Congrès International de Stratigraphie et de Géologie du Carbonifère, Compte Rendu, 5:312.Google Scholar
Schneider, E. L., and Carlquist, S. 1995. Vessels in the roots of Barclaya rotundifolia (Nymphaeaceae). American Journal of Botany 82:13431349.Google Scholar
Schneider, E. L., and Williamson, P. S. 1993. Nymphaeaceae. Pp. 486493in Kubitzki, K., ed. The families and genera of vascular plants. Springer, New York.Google Scholar
Schneider, E. L., Carlquist, S., Beamer, K., and Kohn, A. 1995. Vessels in Nymphaeaceae: Nuphar, Nymphaea, and Ondinea. International Journal of Plant Sciences 156:857862.Google Scholar
Schrank, E. 1983. Scanning electron and light microscopic investigations of angiosperm pollen from the Lower Cretaceous of Egypt. Pollen et Spores 25:213242.Google Scholar
Schrank, E. 1990. Palynology of the clastic Cretaceous sediments between Dongola and Wadi Muqaddam, northern Sudan. Berliner Geowissenschaftliche Abhandlungen, Reihe A 120:149167.Google Scholar
Sculthorpe, C. D. 1967. The biology of aquatic vascular plants. Arnold, London.Google Scholar
Smith, A. C. 1947. The families Illiciaceae and Schisandraceae. Sargentia 7:1224.Google Scholar
Smith, A. C. 1976. Studies of Pacific Island plants, XXXIII. The genus Ascarina (Chloranthaceae) in the southern Pacific. Journal of the Arnold Arboretum 57:405425.Google Scholar
Smith, W. K., Vogelmann, T. C., DeLucia, E. H., Bell, D. T., and Shepherd, K. A. 1997. Leaf form and photosynthesis. Bioscience 47:785793.Google Scholar
Smits, A. J. M., van Ruremonde, R., and van der Velde, G. 1990. Germination requirements and seed banks of some nymphaeid macrophytes: Nymphaea alba L., Nuphar lutea (L.) Sm., and Nymphoides peltata (Gmel.) O. Kuntze. Freshwater Biology 24:315326.Google Scholar
Soltis, D. E., Soltis, P. S., Nickrent, D. L., Johnson, L. A., Hahn, W. J., Hoot, S. B., Sweere, J. A., Kuzoff, R. K., Kron, K. A., Chase, M. W., Swensen, S. M., Zimmer, E. A., Chaw, S.-M., Gillespie, L. J., Kress, W. J., and Sytsma, K. J. 1997. Angiosperm phylogeny inferred from 18S ribosomal DNA sequences. Annals of the Missouri Botanical Garden 84:149Google Scholar
Soltis, D. E., Soltis, P. S., Chase, M. W., Mort, M. E., Albach, D. C., Zanis, M., Savolainen, V., Hahn, W. H., Hoot, S. B., Fay, M. F., Axtell, M., Swensen, S. M., Nixon, K. C., and Farris, J. S. 2000. Angiosperm phylogeny inferred from 18S rDNA, rbcL, and atpB sequences. Botanical Journal of the Linnean Society 133:381461.Google Scholar
Sperry, J. S. 2003. Evolution of water transport and xylem structure. International Journal of Plant Sciences 164(Suppl.):S115S127.Google Scholar
Stebbins, G. L. 1965. The probable growth habit of the earliest flowering plants. Annals of the Missouri Botanical Garden 52:457468.Google Scholar
Stebbins, G. L. 1974. Flowering plants: evolution above the species level. Harvard University Press, Cambridge.Google Scholar
Stone, D. E. 1968. Cytological and morphological notes on the southeastern endemic Schisandra glabra (Schisandraceae). Journal of the Elisha Mitchell Scientific Society 84:351356.Google Scholar
Sun, G., Dilcher, D. L., Zheng, S., and Zhou, Z. 1998. In search of the first flower: a Jurassic angiosperm, Archaefructus, from northeast China. Science 182:16921695.Google Scholar
Sun, G., Ji, Q., Dilcher, D. L., Zheng, S., Nixon, K. C., and Wang, X. 2002. Archaefructaceae, a new basal angiosperm family. Science 296:899904.Google Scholar
Swamy, B. G. L., and Bailey, I. W. 1950. Sarcandra, a vesselless genus of the Chloranthaceae. Journal of the Arnold Arboretum 31:117129.Google Scholar
Takhtajan, A. L. 1969. Flowering plants: origin and dispersal. Smithsonian Institution Press, Washington, D.C.Google Scholar
Taylor, D. W., and Hickey, L. J. 1990. An Aptian plant with attached leaves and flowers: implications for angiosperm origin. Science 247:702704.Google Scholar
Taylor, D. W., and Hickey, L. J. 1992. Phylogenetic evidence for the herbaceous origin of angiosperms. Plant Systematics and Evolution 180:137156.Google Scholar
Taylor, D. W., and Hickey, L. J. 1996. Evidence for and implications of an herbaceous origin for angiosperms. Pp. 232266in Taylor, D. W. and Hickey, L. J., eds. Flowering plant origin, evolution, and phylogeny. Chapman and Hall, New York.Google Scholar
Thien, L. B., White, D. A., and Yatsu, L. Y. 1983. The reproductive biology of a relict: Illicium floridanum Ellis. American Journal of Botany 70:719727.Google Scholar
Thorne, R. F. 1976. A phylogenetic classification of the Angiospermae. Evolutionary Biology 9:35106.Google Scholar
Tiffney, B. H. 1984. Seed size, dispersal syndromes, and the rise of the angiosperms: evidence and hypothesis. Annals of the Missouri Botanical Garden 71:551576.Google Scholar
Todzia, C. A. 1988. Chloranthaceae: Hedyosmum. Flora Neotropica Monograph 48:1139.Google Scholar
Upchurch, G. R. 1984a. Cuticular anatomy of angiosperm leaves from the Lower Cretaceous Potomac Group. I. Zone I leaves. American Journal of Botany 71:192202.Google Scholar
Upchurch, G. R. 1984b. Cuticle evolution in Early Cretaceous angiosperms from the Potomac Group of Virginia and Maryland. Annals of the Missouri Botanical Garden 71:522550.Google Scholar
Upchurch, G. R. 1995. Dispersed angiosperm cuticles: their history, preparation, and application to the rise of angiosperms in Cretaceous and Paleocene coals, southern western interior of North America. International Journal of Coal Geology 28:161227.Google Scholar
Upchurch, G. R., and Dilcher, D. L. 1990. Cenomanian angiosperm leaf megafossils, Dakota Formation, Rose Creek locality, Jefferson County, southeastern Nebraska. U.S. Geological Survey Bulletin 1915:155.Google Scholar
Upchurch, G. R., and Doyle, J. A. 1981. Paleoecology of the conifers Frenelopsis and Pseudofrenelopsis (Cheirolepidiaceae) from the Cretaceous Potomac Group of Maryland and Virginia. Pp. 167202in Romans, R. C., ed. Geobotany II. Plenum, New York.Google Scholar
Upchurch, G. R., and Wolfe, J. A. 1987. Mid-Cretaceous to Early Tertiary vegetation and climate: evidence from fossil leaves and woods. Pp. 75105in Friis, E. M., Chaloner, W. G., and Crane, P. R., eds. The origins of angiosperms and their biological consequences. Cambridge University Press, Cambridge.Google Scholar
Van de Water, P. K., Leavitt, S. W., and Betancourt, J. L. 1994. Trends in stomatal density and 13C/12C ratios of Pinus flexilis needles during last glacial-interglacial cycle. Science 264:239243.Google Scholar
Verdcourt, B. 1986. Chloranthaceae. Flora Malesiana Series I 10:123144.Google Scholar
Vogelmann, T. C., Nishio, J. N., and Smith, W. K. 1996. Leaves and light capture: light propagation and gradients of carbon fixation within leaves. Trends in Plant Science 1:6570.Google Scholar
Volkheimer, W., and Salas, A. 1975. Die älteste Angiospermen-Palynoflora Argentiniens von der Typuslokalität der unterkretazischen Huitrín-Folge des Neuquén-Beckens. Mikrofloristische Assoziation und biostratigraphische Bedeutung. Neues Jahrbuch für Geologie und Paläontologie Monatshefte 1975:424436.Google Scholar
Walker, J. W., and 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.Google Scholar
Wells, P. V. 1969. The relation between mode of reproduction and extent of species in woody genera of the California chaparral. Evolution 23:264267.Google Scholar
Wilf, P., Wing, S. L., Greenwood, D. R., and Greenwood, C. L. 1998. Using fossil leaves as paleoprecipitation indicators: an Eocene example. Geology 26:203206.Google Scholar
Williamson, P. S., and Schneider, E. L. 1993. Cabombaceae. Pp. 157161in Kubitzki, K., ed. The families and genera of vascular plants. Springer, New York.Google Scholar
Williamson, P. S., Schneider, E., and Malins, L. 1989. Tuber and leaf structure of Ondinea purpurea Den Hartog (Nymphaeaceae). Western Australian Naturalist 18:5261.Google Scholar
Wing, S. L., and Boucher, L. D. 1998. Ecological aspects of the Cretaceous flowering plant radiation. Annual Review of Earth and Planetary Sciences 26:379421.Google Scholar
Wing, S. L., and DiMichele, W. A. 1992. Ecological characterization of plants. Pp. 140180in Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H.-D., and Wing, S. L., eds. Terrestrial ecosystems through time: evolutionary paleoecology of terrestrial plants and animals. University of Chicago Press, Chicago.Google Scholar
Wing, S. L., and Tiffney, B. H. 1987. The reciprocal interaction of angiosperm evolution and tetrapod herbivory. Review of Paleobotany and Palynology 50:179210.Google Scholar
Wolfe, J. A. 1993. A method of obtaining climatic parameters from leaf assemblages. U.S. Geological Survey Bulletin 2040:171.Google Scholar
Wolfe, J. A., Doyle, J. A., and Page, V. M. 1975. The bases of angiosperm phylogeny: paleobotany. Annals of the Missouri Botanical Garden 62:801824.Google Scholar
Zanis, M. J., Soltis, D. E., Solits, P. S., Mathews, S., and Donoghue, M. J. 2002. The root of the angiosperms revisited. Proceedings of the National Academy of Sciences USA 99:68486853.Google Scholar
Ziegler, A. M., Raymond, A. L., Gierlowski, T. C., Howell, M. A., Rowley, D. B., and Lottes, A. L. 1987. Coal, climate, and terrestrial productivity: the present and Early Cretaceous compared. Pp. 2549in Scott, A. C., ed. Coal and coal-bearing strata: recent advances. Geological Society, London.Google Scholar