Hostname: page-component-7479d7b7d-rvbq7 Total loading time: 0 Render date: 2024-07-12T09:23:32.427Z Has data issue: false hasContentIssue false

Natural history of a plant trait: branch-system abscission in Paleozoic conifers and its environmental, autecological, and ecosystem implications in a fire-prone world

Published online by Cambridge University Press:  25 February 2013

Cindy V. Looy*
Affiliation:
Department of Integrative Biology and Museum of Paleontology, University of California, 3060 Valley Life Sciences Building #3140, Berkeley, California 94720, U.S.A. E-mail: looy@berkeley.edu

Abstract

Within conifers, active abscission of complete penultimate branch systems is not common and has been described mainly from juveniles. Here I present evidence for the abscission of penultimate branch systems within early so-called walchian conifers—trees with a plagiotropic branching pattern. The specimens studied originate from a middle Early Permian gymnosperm-dominated flora within the middle Clear Fork Group of north-central Texas. Complete branch systems of three walchian conifer morphotypes are preserved; all have pronounced swellings and smooth separation faces at their bases. The source plants grew in a streamside habitat under seasonally dry climatic conditions. The evolution of active branch abscission appears to correspond to an increase in the size of conifers, and this combination potentially contributed to the restructuring of conifer-rich late Paleozoic landscapes. Moreover, trees shedding branch systems and producing abundant litter have the potential to affect the fire regime, which is a factor of evolutionary importance because wildfires must have been a source of frequent biotic disturbance throughout the hyperoxic Early Permian.

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

Addicott, F. T. 1982. Abscission behavior. Pp. 4496inAddicott, F. T., ed. Abscission. University of California Press, Berkeley.Google Scholar
Addicott, F. T. 1991. Abscission: shedding of parts. Pp. 273300inRaghavendra, A. S., ed. Physiology of trees. Wiley, New York.Google Scholar
Addicott, F. T., and Lyon, J. L. 1973. Physiological ecology of abscission. Pp. 149204inKozlowski, T. T., ed. Shedding of plant parts. Academic Press, New York.Google Scholar
Barnard, C. 1926. Preliminary note on branch fall in the Coniferales. Proceedings of the Linnean Society of New South Wales 51:114128.Google Scholar
Behrensmeyer, A. K., and Hook, R. W. 1992. Paleoenvironmental contexts and taphonomic modes. Pp. 15136inBehrensmeyer, A. K., Damuth, J. D., DiMichele, W. A., Potts, R., Sues, H. D., and Wing, S. L., eds. Evolutionary paleoecology of terrestrial plants and animals. University of Chicago Press, Chicago.Google Scholar
Belcher, C. M., Yearsley, J. M., Hadden, R. M., McElwain, J. C., and Rein, G. 2010. Baseline intrinsic flammability of Earth's ecosystems estimated from paleoatmospheric oxygen over the past 350 million years. Proceedings of the National Academy of Sciences USA 107:2244822453.CrossRefGoogle ScholarPubMed
Bellani, L. M., and Bottacci, A. 1995. Anatomical studies of branchlet abscission related to crown modification in Quercus cerris L. Trees 10:2023.CrossRefGoogle Scholar
Benson, J., and Allen, C. 2007. Vegetation associated with Wollemia nobilis (Araucariaceae). Cunninghamia 10:255262.Google Scholar
Bergman, N. M., Lenton, T. M., and Watson, A. L. 2004. COPSE: a new model of biogeochemical cycling over Phanerozoic time. American Journal of Science 304:397437.CrossRefGoogle Scholar
Berner, R. A. 2009. GEOCARBSULF: Phanerozoic atmospheric oxygen: new results using the GEOCARBSULF model. American Journal of Science 309:603606.CrossRefGoogle Scholar
Biffin, E., Hill, R. S., and Lowe, A. J. 2010. Did Kauri (Agathis: Araucariaceae) really survive the Oligocene drowning of New Zealand? Systematic Biology 59:594602.CrossRefGoogle ScholarPubMed
Boland, D. J., Brooker, M. I. H., Chippendale, G. M., Hall, N., Hyland, B. P. M., Johnston, R. D., Kleinig, D. A., McDonald, M. W., and Turner, J. D. 2006. Forest trees of Australia. CSIRO Publishing, Collingwood, Australia.CrossRefGoogle Scholar
Bond, W. J., and Keeley, J. E. 2005. Fire as a global ‘herbivore': the ecology and evolution of flammable ecosystems. Trends in Ecology and Evolution 20:387394.CrossRefGoogle ScholarPubMed
Büsgen, M. 1897. Bau und Leben unsere Waldbäume. Fischer, Jena, Germany.CrossRefGoogle Scholar
Burrows, G. E., Meagher, P. F., and Heady, R. D. 2007. An anatomical assessment of branch abscission and branch-base hydraulic architecture in the endangered Wollemia nobilis. Annals of Botany 99:609623.CrossRefGoogle ScholarPubMed
Chaloner, W. G. 1989. Fossil charcoal as an indicator of paleoatmospheric oxygen level. Journal of the Geological Society 146:171174.CrossRefGoogle Scholar
Chaney, D. S., and DiMichele, W. A. 2007. Paleobotany of the classic redbeds (Clear Fork Group - Early Permian) of North Central Texas. Pp. 357366inWong, T., ed. Proceedings of the XVth International Congress on Carboniferous and Permian Stratigraphy. Utrecht, The Netherlands.Google Scholar
Chaney, D. S., Mamay, S. H., DiMichele, W. A., and Kerp, H. 2009. Auritifolia gen. nov., probable seed plant foliage with comioid affinities from the Early Permian of Texas, U.S.A. International Journal of Plant Sciences 170:247266.CrossRefGoogle Scholar
Clement-Westerhof, J. A. 1984. Aspects of Permian palaeobotany and palynology. IV. The conifer Ortiseia Florin from the Val Gardena Formation of the Dolomites and the Vicentinian Alps (Italy) with special reference to a revised concept of the Walchiaceae (Göppert) Schimper. Review of Palaeobotany and Palynology 41:51166.CrossRefGoogle Scholar
Cunningham, C. R., Feldman, H. R., Franseen, E. K., Gastaldo, R. A., Mapes, G., Mapes, C. G., and Schultze, H. P. 1993. The upper Carboniferous Hamilton fossil-lagerstätte in Kansas: a valley-fill, tidally influenced deposit. Lethaia 26:225236.CrossRefGoogle Scholar
De Deyn, G. B., and Van der Putten, W. H. 2005. Linking aboveground and belowground diversity. Trends in Ecology and Evolution 20:625633.CrossRefGoogle ScholarPubMed
Dewit, L., and Reid, D. M. 1992. Branch abscission in Balsam Poplar (Populus balsamifera): characterization of the phenomenon and the influence of wind. International Journal of Plant Sciences 153:556564.CrossRefGoogle Scholar
Diessel, C. F. K. 2010. The stratigraphic distribution of inertinite. International Journal of Coal Geology 81:251–168.CrossRefGoogle Scholar
DiMichele, W. A., and Aronson, R. B. 1992. The Pennsylvanian-Permian vegetational transition: a terrestrial analogue to the Onshore-Offshore Hypothesis. Evolution 46:807824.CrossRefGoogle Scholar
DiMichele, W. A., Mamay, S. H., Chaney, D. S., Hook, R. W., and Nelson, J. W. 2001. An Early Permian flora with Late Permian and Mesozoic affinities from north-central Texas. Journal of Paleontology 75:4494602.0.CO;2>CrossRefGoogle Scholar
DiMichele, W. A., Kerp, H., and Chaney, D. S. 2004. Tropical floras of the Late Pennsylvanian-Early Permian transition: Carrizo Arroyo in context. InLucas, S. G. and Ziegler, K. B., eds. Carboniferous-Permian transition. Museum of Natural History and Science Bulletin 25:105109.Google Scholar
DiMichele, W. A., Chaney, D. S., Nelson, W. J., Lucas, S. G., Looy, C. V., Quick, K., and Jun, W. 2007. A low diversity, seasonal tropical landscape dominated by conifers and peltasperms: Early Permian Abo Formation, New Mexico. Review of Palaeobotany and Palynology 145:249273.CrossRefGoogle Scholar
DiMichele, W. A., Kerp, H., Tabor, N. J., and Looy, C. V. 2008. The so-called “Paleophytic-Mesophytic” transition in equatorial Pangea: multiple biomes and vegetational tracking of climate change through geological time. Palaeogeography, Palaeoclimatology, Palaeoecology 268:152163.CrossRefGoogle Scholar
Dunn, K. A., McLean, R. J. C., Upchurch, G. R., and Folk, R. L. 1997. Enhancement of leaf fossilization potential by bacterial biofilms. Geology 25:11191122.2.3.CO;2>CrossRefGoogle Scholar
Enright, N. J., and Hill, R. S. 1995. Ecology of the southern conifers. Melbourne University Press, Melbourne.Google Scholar
Erwin, D. H. 2008. Macroevolution of ecosystem engineering, niche construction and diversity. Trends in Ecology and Evolution 23:304310.CrossRefGoogle ScholarPubMed
Falcon-Lang, H. J. 2006. Vegetation ecology of Early Pennsylvanian alluvial fan and piedmont environments in southern New Brunswick, Canada. Palaeogeography, Palaeoclimatology, Palaeoecology 233:3450.CrossRefGoogle Scholar
Florin, R. 1938 –1945. Die Koniferen des Oberkarbons und des unteren Perms I–VII. Palaeontographica B 85:1729.Google Scholar
Florin, R. 1950. Upper Carboniferous and Lower Permian conifers. Botanical Review 16:258282.CrossRefGoogle Scholar
Gastaldo, R. A. 1991. Plant taphonomy character of the Late Carboniferous Hamilton Quarry, Kansas, USA: preservational modes of walchian conifers and implied relationships for residency time in aquatic environments. Pp. 393399inKovar-Eder, J., ed. Palaeovegetational development in Europe and regions relevant to its palaeofloristic evolution. Pan-European Paleobotanical Conference, Museum of Natural History, Vienna.Google Scholar
Glasspool, I. J., and Scott, A. S. 2010. Phanerozoic concentrations of atmospheric oxygen reconstructed from sedimentary charcoal. Nature Geosciences 3:627630.CrossRefGoogle Scholar
Harris, T. M. 1976. Two neglected aspects of fossil conifers. American Journal of Botany 63:902910.CrossRefGoogle Scholar
He, T., Pausas, J. G., Belcher, C. M., Schwilk, D. W., and Lamont, B. B. 2012. Fire-adapted traits of Pinus arose in the fiery Carboniferous. New Phytologist 94:751759.CrossRefGoogle Scholar
Heady, R. D., and Burrows, G. E. 2008. Features of the secondary xylem that facilitate branch abscission in juvenile Wollemia nobilis. IAWA Journal 29:225236.CrossRefGoogle Scholar
Hernandez-Castillo, G. R., Rothwell, G. W., and Mapes, G. 2001. Thucydiaceae fam. nov., with a review and reevaluation of Paleozoic walchian conifers. International Journal of Plant Sciences 162:11551185CrossRefGoogle Scholar
Hernandez-Castillo, G. R., Rothwell, G. W., Stockey, R. A., and Mapes, G. 2003. Growth architecture of Thucydia mahoningensis, a model for primitive walchian conifer plants. International Journal of Plant Sciences 164:443452.CrossRefGoogle 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.CrossRefGoogle Scholar
Hernandez-Castillo, G. R., Rothwell, G. W., Stockey, R. A., and Mapes, G. 2009b. Reconstruction of the Pennsylvanian-age walchian conifer Emporia cryptica sp. nov. (Emporiaceae: Voltziales). Review of Palaeobotany and Palynology 157:218237.CrossRefGoogle Scholar
Hernandez-Castillo, G. R., Rothwell, G. W., Stockey, R. A., and Mapes, G. 2009c. A new voltzialean conifer Emporia royalii (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.CrossRefGoogle Scholar
Hunt, A. 1983. Plant fossils and lithostratigraphy of the Abo Formation (Lower Permian) in the Socorro area and plant biostratigraphy of Abo red beds in New Mexico. Pp. 157163inNew Mexico Geological Society Guidebook, 34th Field Conference of the New Mexican Geological Society, Socorro, N.M.CrossRefGoogle Scholar
Kerp, H. 2000. The modernization of landscapes during the Late Paleozoic-Early Mesozoic. InGastaldo, R. A. and DiMichele, W. A., eds. Phanerozoic terrestrial ecosystems. Paleontological Society Papers 6:79113.CrossRefGoogle Scholar
Kerp, H., and Fichter, J. 1985. Die Makrofloren des saarpfälzischen Rotliegenden (? Ober-Karbon – Unter Perm; SW Deutschland). Mainzer Geowissenschaftliche Mitteilungen 14:159286.Google Scholar
Kerp, J. H. F., Poort, R. J., Swinkels, H. A. J. M., and 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.CrossRefGoogle Scholar
Keeley, J. E., and Zedler, P. H. 1998. Evolution of life histories in Pinus. Pp. 219249inRichardson, D. M., ed. Ecology and biogeography of Pinus. Cambridge University Press, Cambridge.Google Scholar
Kunzmann, L. 2007. Araucariaceae (Pinopsida): aspects in palaeobiogeography and palaeobiodiversity in the Mesozoic. Zoologischer Anzeiger 246:257277.CrossRefGoogle Scholar
Lawton, J. H. 1994. What do species do in ecosystems? Oikos 71:367374.CrossRefGoogle Scholar
Leisman, G. A., Gillespie, W. H., and Mapes, G. 1988. Plant megafossils form the Hartford Limestone (Virgillian-Upper Pennsylvanian) near Hamilton, Kansas. InMapes, G. and Mapes, R. H., eds. Regional geology and paleontology of Upper Paleozoic Hamilton quarry area in southeastern Kansas. Kansas Geological Survey Guidebook Series 6:203212.Google Scholar
Licitis-Lindbergs, R. 1956. Branch abscission and disintegration of the female cones of Agathis australis Salisb. Phytomorphology 6:151167.Google Scholar
Lucas, S. G., Anderson, O. J., Heckert, A. B., and Hunt, A. P. 1995. Geology of Early Permian tracksites, Robledo Mountains, south-central New Mexico. New Mexico Museum of Natural History and Science Bulletin 6:1332.Google Scholar
Mapes, G., and Rothwell, G. W. 1991. Structure and relationships of primitive conifers. Neues Jahrbuch für Geologie und Paläontologie Monatshefte 183:269287.Google Scholar
Millington, W. F., and Chaney, W. R. 1973. Shedding of shoots and branches. Pp. 149204inKozlowski, T. T., ed. Shedding of plant parts. Academic Press, New York.CrossRefGoogle Scholar
Montañez, I. P., Tabor, N. J., Niemeier, D., DiMichele, W. A., Frank, T. D., Fielding, C. R., Isbell, J. L., Birgenheier, L. P., and Rygel, M. C. 2007. CO2-forced climate and vegetation instability during Late Paleozoic deglaciation. Science 315:791.CrossRefGoogle ScholarPubMed
Moore, R. C. 1933. Historical geology. McGraw-Hill, New York.Google Scholar
Nelson, W. J., Hook, R. W., and Tabor, N. J. 2001. Clear Fork Group (Leonardian, Lower Permian) of north-central Texas. InJohnson, K. S., ed. Pennsylvanian and Permian geology and petroleum in the southern Midcontinent, 1998 symposium. Oklahoma Geological Survey Circular 104:167169.Google Scholar
NSW NPWS. 2006. Fire management strategy Wollemi National Park. Department of Environment and Conservation (NSW NPWS, Central Coast Hunter Range and Blue Mountains Regions). Hurstville, New South Wales.Google Scholar
Ohlson, M., Brown, K. J., Birks, H. J. B., Grytnes, J. A., Hörnberg, G., Niklasson, M., Seppä, H., and Bradshaw, R. H. W. 2011. Invasion of Norway spruce diversifies the fire regime in boreal European forests. Journal of Ecology 99:395403CrossRefGoogle Scholar
Rees, P. M., Ziegler, A. M., Gibbs, M. T., Kutzbach, J. E., Behrling, P. J., and Rowley, D. B. 2002. Permian phytogeographic patterns and climate data/model comparisons. Journal of Geology 110:131.CrossRefGoogle Scholar
Rothwell, G. W., and Mapes, G. 1988. Vegetation of a Paleozoic conifer community. InMapes, G. and Mapes, R. H., eds. Regional geology and paleontology of Upper Paleozoic Hamilton quarry area in southeastern Kansas. Kansas Geological Survey Guidebook Series 6:213223.Google Scholar
Rothwell, G. W., and Mapes, G. 2001. Barthelia furcata gen. et sp. nov., with a review of Paleozoic coniferophytes and a discussion of coniferophyte systematics. International Journal of Plant Sciences 162:637667.CrossRefGoogle Scholar
Rothwell, G. W., Mapes, G., and Mapes, R. H. 1997. Late Paleozoic conifers of North America: structure, diversity and occurrences. Review of Palaeobotany and Palynology 95:95113.CrossRefGoogle 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.CrossRefGoogle Scholar
Schaffner, J. H. 1902. The self-pruning of woody plants. Ohio Naturalist 6:4551.Google Scholar
Schwilk, D. W., and Ackerly, D. D. 2001. Flammability and serotiny as strategies: correlated evolution in pines. Oikos 94:326336CrossRefGoogle Scholar
Scott, A. C., and Glasspool, J. J. 2006. The diversification of Paleozoic fire systems and fluctuations in the atmospheric oxygen concentration. Proceedings of the National Academy of Sciences USA 103:1086110865.CrossRefGoogle ScholarPubMed
Stockey, R. A. 1994. Mesozoic Araucariaceae: morphology and systematic relationships. Journal of Plant Research 107:493502.CrossRefGoogle Scholar
Tabor, N. J., and Montañez, I. P. 2004. Morphology and distribution of fossil soils in the Permo-Pennsylvanian Wichita and Bowie Groups, north-central Texas, USA: implications for western equatorial Pangean palaeoclimate during icehouse-greenhouse transition. Sedimentology 51:851884.CrossRefGoogle Scholar
Tabor, N. J., Montañez, I. P., Scotese, C. R., Poulsen, C. J. and Mack, G. H. 2008. Paleosol archives of environmental and climatic history in paleotropical western Pangea during the latest Pennsylvanian through Early Permian. InFielding, C. R., Frank, T. D., and Isbell, J. L., eds. Resolving the late Paleozoic ice age in time and space. Geological Society of America Special Paper 441:291303.Google Scholar
Taiz, L., and Zeiger, E. 2006. Plant physiology, 4th ed. Sinauer, Sunderland, Mass.Google Scholar
Thomas, B. A., and Cleal, C. J. 1999. Abscission in the fossil record. Pp. 183203inKuhrmann, M. H. and Hemsley, A. R., eds. The evolution of plant architecture. Royal Botanic Gardens, Kew, U.K.Google Scholar
Uhl, D., Lausberg, S., Noll, R., and Stapf, K. R. G. 2004. Wildfires in the Late Palaeozoic of Central Europe—an overview of the Rotliegend (Upper Carboniferous–Lower Permian) of the Saar-Nahe Basin (SW-Germany). Palaeogeography, Palaeoclimatology, Palaeoecology 207:2335.CrossRefGoogle Scholar
Van der Pijl, L. 1952. Absciss-joints in the stems and leaves of tropical plants. Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen C 55:574586.Google Scholar
Visscher, H., Kerp, J. H. F., and Clement-Westerhof, J. A. 1986 Aspects of Permian palaeobotany and palynology. VI. Towards a flexible system of naming Paleozoic conifers. Acta Botanica Neerlandia 35:8799.CrossRefGoogle Scholar
Wardlaw, B. R. 2005. Age assessment of the Pennsylvanian-Early Permian succession of north-central Texas. Permophiles 46:2122.Google Scholar
Wardle, D. A., Bardgett, R. D., Klironomos, J. N., Setälä, H., Van der Putten, W. H., and Wall, D. H. 2004. Ecological linkages between aboveground and belowground biota. Science 304:16291633.CrossRefGoogle ScholarPubMed
Watson, A. J. 1978. Consequence for the biosphere of forest and grassland fires. Ph.D. thesis. University of Reading, Reading, U.K.Google Scholar
Weidman, R. H. 1939. Evidence of racial influence in a 25-year test of Ponderosa Pine. Agricultural Research 59:855887.Google Scholar
Whitmore, T. C., and Page, C. N. 1980. Evolutionary implications of the distribution and ecology of the tropical conifer Agathis. New Phytologist 84:407416.CrossRefGoogle Scholar
Wilson, V. R., Gould, K. S., Lovell, P. H., and Aitken-Christie, J. 1998a. Branch morphology and abscission in Kauri, Agathis australis (Auracariaceae). New Zealand Journal of Botany 36:135140.CrossRefGoogle Scholar
Wilson, V. R., Gould, K. S., Lovell, P. H., and Aitken-Christie, J. 1998b. In vitro abscission of Kauri (Agathis australis) branches. New Zealand Journal of Botany 36:495501.CrossRefGoogle Scholar
Winston, R. B. 1984. The Upper Pennsylvanian conifer Walchia garnettensis: structure and affinities. Palaeontographica B 194:97108.Google Scholar