Hostname: page-component-76fb5796d-dfsvx Total loading time: 0 Render date: 2024-04-26T16:58:57.249Z Has data issue: false hasContentIssue false

Modeling fossil plant form-function relationships: A critique

Published online by Cambridge University Press:  26 February 2019

Karl J. Niklas*
Affiliation:
Department of Plant Biology, Cornell University, Ithaca, New York 14853. E-mail: KJN2@cornell.edu

Abstract

Attempts to model form-function relationships for fossil plants rely on the facts that the physiological and structural requirements for plant growth, survival, and reproductive success are remarkably similar for the majority of extant and extinct species regardless of phyletic affiliation and that most of these requirements can be quantified by means of comparatively simple mathematical expressions drawn directly from the physical and engineering sciences. Owing in part to the advent and rapid expansion of computer technologies, the number of fossil plant form-function models has burgeoned in the last two decades and encompasses every level of biological organization ranging from molecular self-assembly to ecological and evolutionary dynamics. This recent and expansive interest in modeling fossil plant form-function relationships is discussed in the context of the general philosophy of modeling past biological systems and how the reliability of models can be examined (i.e., direct experimental manipulation or observation of the system being modeled). This philosophy is illustrated and methods of validating models are critiqued in terms of four models drawn from the author's work (the quantification of wind-induced stem bending stresses, wind pollination efficiency of early Paleozoic ovulate reproductive structures, population dynamics and species extinction in monotypic and “mixed” communities, and the adaptive radiation of early vascular land plants). The assumptions and logical (mathematical) consequences (predictions) of each model are broadly outlined, and, in each case, the model is shown to be overly simplistic despite its ability to predict the general or particular behavior or operation of the system modeled. Nonetheless, these four models, which illustrate some of pros and cons of modeling fossil form-function relationships, are argued to be pedagogically useful because, like all models, they expose the internal logical consistency of our basic assumptions about how organic form and function interrelate.

Type
Research Article
Copyright
Copyright © 2000 by 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

Beerling, D. J., Chaloner, W. G., Huntley, B., Pearson, J. A., Tooley, M. J., and Woodward, F. I. 1992. Variations in the stomatal density of Salix herbacea L. under the changing atmospheric carbon dioxide concentrations of late and post-glacial time. Philosophical Transactions of the Royal Society of London B 336:215224.Google Scholar
Beerling, D. J., Chaloner, W. G., Huntley, B., Pearson, J. A., and Tooley, M. J. 1993a. Stomatal density responds to glacial cycle of environmental change. Proceedings of the Royal Society of London B 251:133138.Google Scholar
Beerling, D. J., Mattey, D. P., and Chaloner, W. G. 1993b. Shifts in the delta carbon-13 composition of Salix herbacea L. leaves in response to spatial and temporal gradients of atmospheric carbon dioxide concentration. Proceedings of the Royal Society of London B 253:5360.Google Scholar
Bower, F. O. 1930. Size and form in plants, with special reference to conducting tracts. Macmillan, London.Google Scholar
Clarkson, E. N. K. 1966. Schizochroal eyes and vision in some Silurian acastid trilobites. Palaeontology 9:129.Google Scholar
DiMichele, W. A. 1981. Arborescent lycopods of Pennsylvanian age coals: Lepidodendron, with a description of a new species. Palaeontographica 175B:85125.Google Scholar
Ellison, A. M., and Niklas, K. J. 1988. Branching patterns of Salicornia europaea (Chenopodiaceae) at different successional stages: a comparison of theoretical and real plants. American Journal of Botany 75:501512.Google Scholar
Empacher, N., Mosbrugger, V., Roth, A., Wolf, M., and Wunderlin, A. 1995. Qualitative mathematical discussion of different evolutionary states in water transport systems of plants. Journal of Biological Physics 21:241264.Google Scholar
Fairon-Demaret, M., and Scheckler, S. E. 1987. Typification and redescription of Moresnetia zalesskyi Stockmans, 1948, an early seed plant from the Upper Famennian of Belgium. Bulletin de L'Institut Royal des Sciences Naturelles de Belgique, Sciences de la Terre 57:183199.Google Scholar
Gates, D. M. 1965. Energy, plants, and ecology. Ecology 46:116.Google Scholar
Habgood, K. S., Hemsley, A. R., and Thomas, B. A. 1998. Modelling of the dispersal of Lepidocarpon based on experiments using reconstructions. Review of Palaeobotany and Palynology 102:101114.Google Scholar
Hemsley, A. R. 1998. Nonlinear variation in simulated complex pattern development. Journal of Theoretical Biology 192:7379.Google Scholar
Hemsley, A. R., Jenkins, P. D., Collinson, M. E., and Vincent, B. 1996. Experimental modelling of exine self-assembly. Botanical Journal of the Linnean Society 121:177187.Google Scholar
Hemsley, A. R., Vincent, B., Collinson, M. E., and Griffiths, P. C. 1998. Simulated self-assembly of spore exines. Annals of Botany 82:105109.Google Scholar
Hotton, C. L., and Stein, W. E. 1994. An ontogenetic model for the Mississippian seed plant family Calamopityaceae. International Journal of Plant Science 155:119142.Google Scholar
Meyer-Berthaud, B., and Stein, W. E. 1995. A reinvestigation of Stenomyelon from the Late Tournaisan of Scotland. International Journal of Plant Science 156:863895.Google Scholar
Mosbrugger, V., and Roth, A. 1996. Biomechanics in fossil plant biology. Review of Palaeobotany and Palynology 90:195207.Google Scholar
Niklas, K. J. 1976. Morphological and ontogenetic reconstruction of Parka decipiens Fleming and Pachytheca Hooker from the Lower Old Red Sandstone, Scotland. Transactions of the Royal Society of Edinburgh 69:483499.Google Scholar
Niklas, K. J. 1977. Ontogenetic constructions of some fossil plants. Review of Palaeobotany and Palynology 23:337357.Google Scholar
Niklas, K. J. 1978. Branching patterns and mechanical design in Paleozoic plants: a theoretic assessment. Annals of Botany 42:3339.Google Scholar
Niklas, K. J. 1979. Simulations of apical developmental sequences in bryophytes. Annals of Botany 44:339352.Google Scholar
Niklas, K. J. 1981a. Simulated wind pollination and airflow around ovules of some early seed plants. Science 211:275277.Google Scholar
Niklas, K. J. 1981b. 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. Computer simulations of early land plant branching morphologies: canalization of patterns during evolution? Paleobiology 8:196210.Google Scholar
Niklas, K. J. 1983. The influence of Paleozoic ovule and cupule morphologies on wind pollination. Evolution 37:968986.Google Scholar
Niklas, K. J. 1984. Size-related changes in the primary xylem anatomy of some early tracheophytes. Paleobiology 10:487506.Google Scholar
Niklas, K. J. 1985a. The evolution of tracheid diameter in early vascular plants and its implications on the hydraulic conductance of the primary xylem strand. Evolution 39:11101122.Google Scholar
Niklas, K. J. 1985b. The aerodynamics of wind pollination. Botanical Review 51:328386.Google Scholar
Niklas, K. J. 1986a. Computer-simulated plant evolution. Scientific American 254:7886.Google Scholar
Niklas, K. J. 1986b. Evolution of plant shape: design constraints. Trends in Ecology and Evolution 1:6772.Google Scholar
Niklas, K. J. 1986c. Computer simulations of branching-patterns and their implications on the evolution of plants. Lectures on Mathematics in the Life Sciences 18:150.Google Scholar
Niklas, K. J. 1988a. Biophysical limitations on plant form and evolution. Pp. 185220 in Gottlieb, L. D. and Jain, S. K., eds. Plant evolutionary biology. Chapman and Hall, London.Google Scholar
Niklas, K. J. 1988b. The role of phyllotactic pattern as a “Developmental Constraint” on the interception of light by leaf surfaces. Evolution 42:116.Google Scholar
Niklas, K. J. 1990. Biomechanics of Psilotum nudum and some early Paleozoic vascular sporophytes. American Journal of Botany 72:590606.Google Scholar
Niklas, K. J. 1992. Plant biomechanics: an engineering approach to plant form and function. University of Chicago Press, Chicago.Google Scholar
Niklas, K. J. 1993. Ontogenetic-response models and the evolution of plant size. Evolutionary Trends in Plants 7:4349.Google Scholar
Niklas, K. J. 1994a. Morphological evolution through complex domains of fitness. Proceedings of the National Academy of Sciences USA 91:67726779.Google Scholar
Niklas, K. J. 1994b. Plant allometry: the scaling of form and process. University of Chicago Press, Chicago.Google Scholar
Niklas, K. J. 1994c. The scaling of plant and animal body mass, length, and diameter. Evolution 48:4454.Google Scholar
Niklas, K. J. 1994d. Predicting the height of fossil plant remains: an allometric approach to an old problem. American Journal of Botany 81:12351242.Google Scholar
Niklas, K. J. 1997a. Adaptive walks through fitness landscapes for early vascular land plants. American Journal of Botany 84:1625.Google Scholar
Niklas, K. J. 1997b. Effects of hypothetical developmental barriers and abrupt environmental changes on adaptive walks in a computer-generated domain for early vascular land plants. Paleobiology 23:6376.Google Scholar
Niklas, K. J. 1997c. The evolutionary biology of plants. University of Chicago Press, Chicago.Google Scholar
Niklas, K. J. 1998. The influence of gravity and wind on land plant evolution. Review of Palaeobotany and Palynology 102:114.Google Scholar
Niklas, K. J. 1999. Evolutionary walks through a land plant morphospace. Journal of Experimental Botany 50:3952.Google Scholar
Niklas, K. J., and Banks, H. P. 1985. Evidence for xylem constrictions in the primary vasculature of Psilophyton dawsonii, an Emsian trimerophyte. American Journal of Botany 72:674685.Google Scholar
Niklas, K. J., and Buchmann, S. L. 1987. The aerodynamics of pollen capture in two sympatric Ephedra species. Evolution 41:104123.Google Scholar
Niklas, K. J., and Chaloner, W. G. 1976. Simulations of the ontogeny of Spongiophyton, a Devonian plant. Annals of Botany 40:111.Google Scholar
Niklas, K. J., and Kerchner, V. 1984. Mechanical photosynthetic constraints on the evolution of plant shape. Paleobiology 10:79101.Google Scholar
Niklas, K. J., and Phillips, T. L. 1976. Morphology of Protosalvinia from the Upper Devonian of Ohio and Kentucky. American Journal of Botany 63:929.Google Scholar
Nobel, P. S. 1983. Biophysical plant physiology and ecology. W. H. Freeman, New York.Google Scholar
Okubo, A., and Levin, S. A. 1989. A theoretical framework for data analysis of wind dispersal of seeds and pollen. Ecology 70:329338.Google Scholar
Ostrom, J. H. 1964. A functional analysis of the jaw mechanics in the dinosaur Triceratops. Postilla (Peabody Museum of Natural History, Yale University) 88:135.Google Scholar
Phillips, T. L. 1979. Reproduction of heterosporous arborescent lycopods in the Mississippian-Pennsylvanian of Euramerica. Review of Palaeobotany and Palynology 27:239289.Google Scholar
Raup, D. M., and Michelson, A. 1965. Theoretical morphology of the coiled shell. Science 147:12941295.Google Scholar
Retallack, G. 1975. The life and times of a Triassic lycopod. Alcheringa 1:329.Google Scholar
Rex, G. M., and Chaloner, W. G. 1983. The experimental formation of plant compression fossils. Palaeontology 26:231252.Google Scholar
Richter, R. 1928. Psychische Reaktionen fossiler Tiere. Palaeobiologica 1:226244.Google Scholar
Roth, A., and Mosbrugger, V. 1996. Numerical studies of water conduction in land plants: evolution of early stele types. Paleobiology 22:411421.Google Scholar
Roth, A., Mosbrugger, V., and Neugebauer, H. J. 1994. Efficiency and evolution of water transport systems in higher plants: a modelling approach: II. Stelar evolution. Philosophical Transactions of the Royal Society of London B 345:153162.Google Scholar
Roth, A., Mosbrugger, V., Belz, G., and Neugebauer, H. J. 1995. Hydrodynamic modelling study of angiosperm leaf venation types. Botanica Acta 108:121126.Google Scholar
Roth, A., Mosbrugger, V., and Wunderlin, A. 1998. Computer simulations as a tool for understanding the evolution of water transport systems in land plants: a review and new data. Review of Palaeobotany and Palynology 102:7999.Google Scholar
Rothwell, G. W., and Taylor, T. N. 1982. Early seed plant wind pollination studies: a commentary. Taxon 31:308309.Google Scholar
Rudwick, M. J. S. 1964. The inference of function from structure in fossils. British Journal of Philosophy and Science 15:2740.Google Scholar
Salisbury, F. B., and Ross, C. W. 1992. Plant physiology. Wadsworth, Belmont, Calif.Google Scholar
Speck, T., and Vogellehner, D. 1988. Biophysical examinations concerning the mechanical stability of various stele types and the kind of stabilizing system in early vascular land plants. Palaeontographica, Abteilung B 210:91126.Google Scholar
Speck, T., Spatz, H.-C., and Vogellehner, D. 1990a. Contributions to the biomechanics of plants I. Stabilities of plant stems with strengthening elements of different cross-sections against weight and wind forces. Botanica Acta 103:111112.Google Scholar
Speck, T., Spatz, H.-C., and Vogellehner, D. 1990b. Contributions to the biomechanics of plants II. Stability against local buckling in hollow plant stems. Botanica Acta 103:123130.Google Scholar
Stanley, S. M. 1970. Relation of shell form to life habits in the Bivalvia (Mollusca). Geological Society of America Memoir 125.Google Scholar
Stein, W. E. 1993. Modeling the evolution of stelar architecture in vascular plants. International Journal of Plant Sciences 154:229263.Google Scholar
Stein, W. E. 1998. Developmental logic: establishing a relationship between developmental process and phylogenetic pattern in primitive vascular plants. Review of Palaeobotany and Palynology 102:1542.Google Scholar
Stewart, W. N., and Rothwell, G. W. 1993. Paleobotany and the evolution of plants. Cambridge University Press, Cambridge.Google Scholar
Taiz, L., and Zeiger, E. 1991. Plant physiology. Benjamin-Cummings, Redwood City, Calif.Google Scholar
Taylor, T. N., and Taylor, E. L. 1993. The biology and evolution of fossil plants. Prentice Hall, Englewood Cliffs, N. J. Google Scholar
Thompson, D'A. W. 1942. On growth and form. Cambridge University Press, Cambridge.Google Scholar
Vogel, S. 1981. Life in moving fluids. Willard Grant, Boston.Google Scholar
Wagner, F., Bohncke, S. J. P., Dilcher, D. L., Kürschner, W. M., Geel, B. van, and Visscher, H. 1999. Century-scale shifts in Early Holocene atmospheric CO2 concentration. Science 284:19711973.Google Scholar
West, G. B., Brown, J. H., and Enquist, B. J. 1999. The fourth dimension of life: fractal geometry and allometric scaling of organisms. Science 284:16771679.Google Scholar
Zimmermann, M. H. 1983. Xylem structure and the ascent of sap. Springer, Berlin.Google Scholar