Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-26T10:26:50.331Z Has data issue: false hasContentIssue false

Shifting functional roles and the evolution of conifer pollen-producing and seed-producing cones

Published online by Cambridge University Press:  08 April 2016

Andrew B. Leslie*
Affiliation:
Department of the Geophysical Sciences, University of Chicago, 5734 South Ellis Avenue, Chicago, Illinois 60637

Abstract

Exploring patterns in the evolution of seed plant reproductive morphology within a functional context offers a framework in which to identify and evaluate factors that potentially drive reproductive evolution. Conifers are a particularly useful group for studies of this kind because they have a long geologic history and their reproductive organs are borne on separate structures with discrete functions. Multivariate analysis of morphological data collected from pollen-producing and seed-producing cones of Paleozoic, Mesozoic, and extant conifer species shows that seed cones underwent a significant expansion of morphological diversity that began during the Early-Middle Jurassic and has continued into the present day. In contrast, pollen cones show significantly lower levels of morphological diversity and exhibit similar basic morphologies throughout conifer evolutionary history. The increase in seed cone diversity through time is primarily the result of two novel structural and organizational features that evolved independently in different conifer families during the Mesozoic: robust, tightly packed cones in members of Araucariaceae, Cupressaceae sensu lato, and Pinaceae, and highly reduced, fleshy cones or solitary seeds in Podocarpaceae, Taxaceae, and some members of Cupressaceae sensu stricto. In extant conifers, these cone morphologies are associated with species that have strong interactions with vertebrate seed predators, seed dispersers, or a combination of both. This suggests that increases in the strength and complexity of biotic interactions in the Jurassic and Cretaceous were a primary driver of conifer reproductive evolution, and that patterns of character evolution relate to the increasing importance of cone tissue in seed protection and seed dispersal through time.

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

Alvin, K. L. 1960. Further conifers of the Pinaceae from the Wealden Formation of Belgium. Mémoires de l'Institut Royal des Sciences Naturelles de Belgique 146:139.Google Scholar
Anderson, J. M., and Anderson, H. M. 2003. Heyday of the gymnosperms: systematics and biodiversity of the Late Triassic Molteno fructifications. Strelitzia 15. National Botanical Institute, Pretoria.Google Scholar
Archangelsky, S. 1966. New gymnosperms from the Ticó flora, Santa Cruz Province, Argentina. Bulletin of the British Museum (Natural History) 13:261295.Google Scholar
Archangelsky, S., and Cúneo, R. 1987. Ferugliocladaceae, a new conifer family from the Permian of Gondwana. Review of Palaeobotany and Palynology 51:330.Google Scholar
Axsmith, B. J., and Ash, S. R. 2006. Two rare fossil cones from the Upper Triassic Chinle Formation in Petrified Forest National Park, Arizona, and New Mexico. Museum of Northern Arizona Bulletin 62:8294.Google Scholar
Benkman, C. W. 1995. The impact of tree squirrels (Tamiasciurus) on Limber Pine seed dispersal adaptations. Evolution 49:585592.Google Scholar
Benkman, C. W., Parchman, T. L., Favis, A., Siepielski, A. M. 2003. Reciprocal selection causes a coevolutionary arms race between crossbills and lodgepole pine. American Naturalist 162:182194.Google Scholar
Bowe, L. M., Coat, G., and dePamphilis, C. W. 2000. Phylogeny of seed plants based on all three genomic compartments: extant gymnosperms are monophyletic and Gnetales' closest relatives are conifers. Proceedings of the National Academy of Sciences USA 97:40924097.Google Scholar
Boyce, C. K. 2005. Patterns of segregation and convergence in the evolution of fern and seed plant leaf morphologies. Paleobiology 31:117140.2.0.CO;2>CrossRefGoogle Scholar
Boyce, C. K., and Knoll, A. H. 2002. Evolution of developmental potential and the multiple independent origins of leaves in Paleozoic vascular plants. Paleobiology 28:70100.Google Scholar
Cantrill, D. J., and Falcon-Lang, H. J. 2001. Cretaceous (Late Albian) Coniferales of Alexander Island, Antarctica. 2. Leaves, reproductive structures and roots. Review of Palaeobotany and Palynology 115:119145.Google Scholar
Chandler, L. M., and Owens, J. N. 2004. The pollination mechanism of Abies amabilis. Canadian Journal of Forest Research 34:10711080.Google Scholar
Chaw, S.-M., Parkinson, C. L., Cheng, Y., Vincent, T. M., Palmer, J. D. 2000. Seed plant phylogeny inferred from all three plant genomes: monophyly of extant gymnosperms and origin of Gnetales from conifers. Proceedings of the National Academy of Sciences USA 97:40864091.Google Scholar
Chiappe, L. M., and Dyke, G. J. 2002. The Mesozoic radiation of birds. Annual Review of Ecology, Evolution, and Systematics 33:91124.Google Scholar
Clement-Westerhof, J. A. 1987. Aspects of Permian palaeobotany and palynology. VII. The Majonicaceae, a new family of Late Permian conifers. Review of Palaeobotany and Palynology 52:375402.Google Scholar
Clement-Westerhof, J. A. 1988. Morphology and phylogeny of Paleozoic conifers. Pp. 298337in Beck, C. B., ed. Origin and evolution of gymnosperms. Columbia University Press, New York.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
Cresswell, J. E., Henning, K., Pennel, C., Lahoubi, M., Patrick, M. A., Young, P. G., and Tabor, G. R. 2007. Conifer ovulate cones accumulate pollen principally by simple impaction. Proceedings of the National Academy of Sciences USA 104:1814118144.Google Scholar
Doyle, J. A. 2006. Seed ferns and the origin of angiosperms. Journal of the Torrey Botanical Society 133:169209.Google Scholar
Doyle, J. A. 2008. Integrating molecular phylogenetic and paleobotanical evidence on origin of the flower. International Journal of Plant Sciences 169:816843.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.Google Scholar
Eriksson, O. 2008. Evolution of seed size and biotic seed dispersal in angiosperms: paleoecological and neoecological evidence. International Journal of Plant Sciences 169:863870.CrossRefGoogle Scholar
Farjon, A., and Garcia, S. Ortiz 2005. The early development of ovuliferous cones in Cupressaceae s. lat—a survey of the genera. Pp. 2746in Farjon, A., ed. A monograph of Cupressaceae and Sciadopitys. Royal Botanic Gardens, Kew, Richmond, Surrey, U.K.Google Scholar
Florin, R. 1938–1945. Die Koniferen des Oberkarbons und des unteren Perms. I–VII. Palaeontographica, Abteilung B 85:1729.Google Scholar
Florin, R. 1951. Evolution in cordaites and conifers. Acta Horti Bergiani 15:286388.Google Scholar
Florin, R. 1958. On Jurassic taxads and conifers from northwestern Europe and eastern Greenland. Acta Horti Bergiani 17:257402.Google Scholar
Foote, M. 1994. Morphological diversity in Ordovician-Devonian crinoids and the early saturation of morphological space. Paleobiology 20:320344.Google Scholar
Foote, M. 1995. Morphological diversification of Paleozoic crinoids. Paleobiology 21:273299.Google Scholar
Gadek, P. A., Alpers, D. L., Heslewood, M. M., and Quinn, C. J. 2000. Relationships within Cupressaceae sensu lato: a combined morphological and molecular approach. American Journal of Botany 87:10441057.Google Scholar
Geldenhuys, C. J. 1993. Reproductive biology and population structures of Podocarpus falcatus and P. latifolius in southern Cape forests. Botanical Journal of the Linnean Society 112:5974.Google Scholar
Gower, J. C. 1966. Some distance properties of latent root and vector methods used in multivariate analysis. Biometrika 53:325338.Google Scholar
Grauvogel-Stamm, L. 1978. La flore du Grès a Voltzia (Buntsandstein Superior) des Vosges du Nord (France), morphologie, anatomie, interpretations phylogénique et paléogéographique. Sciences Géologiques, Université Louis Pasteur de Strasbourg, Institute de Géologie, Memoir 50:1225.Google Scholar
Harris, T. M. 1935. The fossil flora of Scoresby Sound East Greenland, Part 4. Ginkgoales, Coniferales, Lycopodiales, and isolated fructifications. Meddelelser om Gr⊘nland 112:1176.Google Scholar
Harris, T. M. 1979. The Yorkshire Jurassic flora. V. Coniferales. British Museum of Natural History, London.Google Scholar
Hernandez-Castillo, G. R., Rothwell, G. W., Mapes, G. 2001. Thucydiaceae fam. nov., with a review and reevaluation of Paleozoic walchian conifers. International Journal of Plant Sciences 162:11551185.Google Scholar
Hernandez-Castillo, G. R., Rothwell, G. W., Stockey, R. A., and Mapes, G. 2009a. Reconstructing Emporia lockardii (Emporiaceae) Voltziales, and initial thoughts on Paleozoic conifer ecology. International Journal of Plant Sciences 170:10561074.Google Scholar
Hernandez-Castillo, G. R., Rothwell, G. W., Stockey, R. A., and Mapes, G. 2009b. A new voltzialean conifer Emporia royalii sp. nov. (Emporiaceae) from the Hamilton Quarry, Kansas. International Journal of Plant Sciences 170:12011227.CrossRefGoogle Scholar
Hilton, J., and Bateman, R. M. 2006. Pteridosperms are the backbone of seed-plant phylogeny. Journal of the Torrey Botanical Society 133:119168.Google Scholar
Kerp, J. H. F., Poort, R. J., Swinkels, H. A. J. M., Verwer, R. 1990. Aspects of Permian palaeobotany and palynology. IX. Conifer-dominated Rotliegend Floras from the Saar-Nahe Basin (?Late Carboniferous-Early Permian; SW-Germany) with special reference to the reproductive biology of early conifers. Review of Palaeobotany and Palynology 62:205248.Google Scholar
Kvacek, J. 2000. Frenelopsis alata and its microsporangiate and ovuliferous reproductive structures from the Cenomanian of Bohemia (Czech Republic, Central Europe. Review of Palaeobotany and Palynology 112:5178.Google Scholar
Labandeira, C. C., Kvacek, J., and Mostovski, M. B. 2007. Pollination drops, pollen, and insect pollination of Mesozoic gymnosperms. Taxon 56:663695.Google Scholar
Leslie, A. B. 2008. Interpreting the function of saccate pollen in ancient conifers and other seed plants. International Journal of Plant Sciences 169:10381045.Google Scholar
Leslie, A. B. 2010. Flotation preferentially selects saccate pollen during conifer pollination. New Phytologist 188:273279.Google Scholar
Linhart, Y. B. 1978. Maintenance of variation in cone morphology in California closed-cone pines: the roles of fire, squirrels and seed output. The Southwestern Naturalist 23:2940.Google Scholar
Luo, Z.-X. 2007. Transformation and diversification in early mammal evolution. Nature 450:10111019.Google Scholar
Lupia, R. 1999. Discordant morphological disparity and taxonomic diversity during the Cretaceous angiosperm radiation: North American pollen record. Paleobiology 25:128.Google Scholar
Mabberley, D. J. 2008. Mabberley's plant book: a portable dictionary of plants, their classifications, and uses. Cambridge University Press, New York.Google Scholar
Mapes, G., and Rothwell, G. W. 1998. Pollen cone structure of the Late Pennsylvanian (Stephanian) conifer Emporia. Journal of Paleontology 72:571576.Google Scholar
Mathews, S., Clements, M. D., and Beilstein, M. A. 2010. A duplicate gene rooting of seed plants and the phylogenetic position of flowering plants. Philosophical Transactions of the Royal Society of London B 365:383395.Google Scholar
Mezquida, E. T., and Benkman, C. W. 2005. The geographic selection mosaic for squirrels, crossbills, and Aleppo pine. Journal of Evolutionary Biology 18:348357.Google Scholar
Miller, C. N. Jr. 1977. Mesozoic conifers. Botanical Review 43:217279.Google Scholar
Miller, C. N. Jr. 1999. Implications of fossil conifers for the phylogenetic relationships of living families. Botanical Review 65:239277.Google Scholar
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.Google Scholar
Niklas, K. J. 1982. Simulated and empiric wind pollination patterns of conifer ovulate cones. Proceedings of the National Academy of Sciences USA 79:510514.CrossRefGoogle ScholarPubMed
Niklas, K. J. 1984. The motion of windborne pollen grains around conifer ovulate cones: implications on wind pollination. American Journal of Botany 71:356374.CrossRefGoogle Scholar
Niklas, K. J. 1994. Morphological evolution through complex domains of fitness. Proceedings of the National Academy of Sciences USA 91:67726779.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
Owens, J. N., Catalano, G. L., and Aitken-Christie, J. 1997. The reproductive biology of Kauri (Agathis australis). IV. Late embryology, histochemistry, cone and seed development. International Journal of Plant Sciences 158:395407.CrossRefGoogle Scholar
Owens, J. N., Takaso, T., and Runions, C. J. 1998. Pollination in conifers. Trends in Plant Science 3:479485.CrossRefGoogle Scholar
Quinn, C. J., Price, R. A., and Gadek, P. A. 2002. Familial concepts and relationships in the conifers based on rbcL and matK sequence comparisons. Kew Bulletin 57:513531.Google Scholar
Parchman, T. L., and Benkman, C. W. 2002. Diversifying coevolution between crossbills and black spruce on Newfoundland. Evolution 56:16631672.Google Scholar
Plotnik, R. E., Kenig, F., Scott, A. C., Glasspool, I. J., Eble, C. F., and Lang, W. J. 2009. Pennsylvanian paleokarst and cave fills from Northern Illinois, USA: a window into Late Carboniferous environments and landscapes. Palaios 24:627637.Google Scholar
Rai, H. S., Reeves, P. A., Peakall, R., Olmstead, R. G., and Graham, S. W. 2008. Inference of higher-order conifer relationships from a multi-locus plastid data set. Botany 86:658669.Google Scholar
Rothwell, G. W., and Serbet, R. 1994. Lignophyte phylogeny and the evolution of spermatophytes: a numerical cladistic analysis. Systematic Botany 19:443482.Google Scholar
Rothwell, G. W., Mapes, G., and Hernandez-Castillo, G. R. 2005. Hanskerpia gen. nov. and phylogenetic relationships among the most ancient conifers (Voltziales). Taxon 54:733750.Google Scholar
Runions, C. J., and Owens, J. N. 1996. Pollen scavenging and rain involvement in the pollination mechanism of interior spruce. Canadian Journal of Botany 74:115124.Google Scholar
Scott, A. C., and 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
Serbet, R., Escapa, I., Taylor, T. N., Taylor, E. L., and Cúneo, N. R. 2010. Additional observations on the enigmatic Permian plant Buriadia and implications on early coniferophyte evolution. Review of Palaeobotany and Palynology 161:168178.Google Scholar
Siepielski, A. M., and Benkman, C. W. 2007a. Convergent patterns in the selection mosaic for two North American bird-dispersed pines. Ecological Monographs 77:203220.Google Scholar
Siepielski, A. M., and Benkman, C. W. 2007b. Selection by a predispersal seed predator constrains the evolution of avian seed dispersal in pines. Functional Ecology 21:611618.CrossRefGoogle Scholar
Siepielski, A. M., and 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
Smith, C. C. 1970. The coevolution of Pine Squirrels (Tamiasciurus) and conifers. Ecological Monographs 40:349371.Google Scholar
Stockey, R. A. 1975. Seeds and embryos of Araucaria mirabilis. American Journal of Botany 62:856868.CrossRefGoogle Scholar
Stockey, R. A. 1980a. Anatomy and morphology of Araucaria sphaerocarpa Carruthers from the Jurassic Inferior Oolite of Bruton, Somerset. Botanical Gazette 141:116124.Google Scholar
Stockey, R. A. 1980b. Jurassic araucarian cone from southern England. Paleontology 23:657666.Google Scholar
Takaso, T., and Tomlinson, P. B. 1991. Cone and ovule development in Sciadopitys (Taxodiaceae-Coniferales). American Journal of Botany 78:417428.Google Scholar
Takaso, T., and Tomlinson, P. B. 1992. Seed cone and ovule ontogeny in Metasequoia, Sequoia, and Sequoiadendron (Taxodiaceae-Coniferales). Botanical Journal of the Linnean Society 109:1537.Google Scholar
Takaso, T., and Owens, J. N. 1995. Ovulate cone morphology and pollination in Pseudotsuga and Cedrus. International Journal of Plant Sciences 156:630635.Google Scholar
Taylor, T. N., Taylor, E. L., and Krings, M. 2009. Paleobotany: the biology and evolution of fossil plants. Academic Press, Burlington, Mass.Google Scholar
Thorsen, M. J., Dickinson, K. J. M., and Seddon, P. J. 2009. Seed dispersal systems in the New Zealand flora. Perspectives in Plant Ecology, Evolution, and Systematics 11:285309.CrossRefGoogle Scholar
Tiffney, B. H. 2004. Vertebrate dispersal of seed plants through time. Annual Review of Ecology, Evolution, and Systematics 35:129.Google Scholar
Tomback, D. F., and Linhart, Y. B. 1990. The evolution of bird-dispersed pines. Evolutionary Ecology 4:185219.Google Scholar
Tomlinson, P. B. 1994. Functional morphology of saccate pollen in conifers with special reference to Podocarpaceae. International Journal of Plant Sciences 155:699715.Google Scholar
Tomlinson, P. B., and Takaso, T. 2002. Seed cone structure in conifers in relation to development and pollination: a biological approach. Canadian Journal of Botany 80:12501273.Google Scholar
Tomlinson, P. B., Braggins, J. E., and Ratenbury, J. A. 1991. Pollination drop in relation to cone morphology in Podocarpaceae: a novel reproductive mechanism. American Journal of Botany 78:12891303.Google Scholar
Townrow, J. A. 1967. On Rissikia and Mataia podocarpaceous conifers from the lower Mesozoic of southern lands. Proceedings of the Royal Society of Tasmania 101:103136.CrossRefGoogle Scholar
Vishnu-Mittre, . 1959. Studies on the fossil flora of Nipania (Rajmahal Series). Bihar. Coniferales. Palaeobotanist 6:82112.Google Scholar
Willson, M. F., Sabag, C., Figueroa, J., and Armesto, J. J. 1996. Frugivory and seed dispersal of Podocarpus nubigena in Chiloe, Chile. Revista Chilena de Historia Natural 69:343349.Google Scholar
Wing, S. L., and Sues, H.-D. 1992. Mesozoic and early Cenozoic terrestrial ecosystems. Pp. 327416in Behrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H.-D., and Wing, S. L., eds. Terrestrial ecosystems through time. 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 Palaeobotany and Palynology 50:179210.Google Scholar
Yao, X., Taylor, T. N., and Taylor, E. L. 1997. A taxodiaceous seed cone from the Triassic of Antarctica. American Journal of Botany 84:343354.Google Scholar
Zhou, Z. 1983. Stalagma samara, a new podocarpaceous conifer with monocolpate pollen from the Upper Triassic of Hunan, China. Palaeontographica, Abteilung B 185:5678.Google Scholar
Zhou, Z., and Zhang, F. 2002. A long tailed, seed-eating bird from the Early Cretaceous of China. Nature 418:405409.CrossRefGoogle ScholarPubMed