Hostname: page-component-848d4c4894-75dct Total loading time: 0 Render date: 2024-05-11T04:43:43.727Z Has data issue: false hasContentIssue false

Comparative Paleoecology of Fossils and Fossil Assemblages

Published online by Cambridge University Press:  21 July 2017

Andrew M. Bush
Affiliation:
Department of Ecology and Evolutionary Biology and Center for Integrative Geosciences, University of Connecticut 75 North Eagleville Road, Storrs, CT 06269-3043
Gwen M. Daley
Affiliation:
Department of Chemistry, Physics, and Geology Winthrop University, Rock Hill, SC 29733
Get access

Abstract

Generating and testing hypotheses is an integral part of any science, and some of the most stimulating paleobiological hypotheses of the past few decades relate to the ecological properties of fossils or fossil assemblages. Here, we outline recent methods for framing paleoecological questions that should facilitate the further quantitative evaluation of paleoecological hypotheses. First, we describe theoretical ecospaces, which are frameworks for classifying the ecologic properties of individuals or species based on multiple characters. We discuss the utility of theoretical ecospace in understanding evolutionary constraints and biodiversification, among other topics. Second, we discuss the reconstruction of high-resolution paleoecological gradients using ecological ordination techniques. Ordination can help uncover the paleoenvironmental factors that controlled fossil assemblage composition, track these factors through time, and evaluate the environmental and ecological context of major biotic changes. As an example, we present a new gradient analysis of the Yorktown Formation (Pliocene) of Virginia in which substrate and disturbance controlled molluscan assemblage composition. As a further example, we ordinate samples of mid-Paleozoic and late Cenozoic marine fossil assemblages based on their ecological content (as determined using a theoretical ecospace) to test whether the same environmental and ecological factors controlled the distribution of ecological lifestyles in both time intervals, despite the many differences between them. Although depth-related variation is evident in both data sets, the Cenozoic samples show stronger evidence of environmental control on ecologic content within depth zones. In contrast, Paleozoic gradients are consistent with a more random component in assemblage content. These analyses are quite preliminary, however, and should be verified with more extensive data.

Type
Research Article
Copyright
Copyright © 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

Alroy, J., Marshall, C. R., Bambach, R. K., Bezusko, K., Foote, M., Fürsich, F. T., Hansen, T. A., Holland, S. M., Ivany, L. C., Jablonski, D., Jacobs, D. K., Jones, D. C., Kosnik, M. A., Lidgard, S., Low, S., Miller, A. I., Novack-Gottshall, P. M., Olszewski, T. D., Patzkowsky, M. E., Raup, D. M., Roy, K., Sepkoski, J. J. Jr., Sommers, M. G., Wagner, P. J., and Webber, A. 2001. Effects of sampling standardization on estimates of Phanerozoic marine diversification. Proceedings of the National Academy of Sciences, 98:62616266.Google Scholar
Alroy, J., Aberhan, M., Bottjer, D. J., Foote, M., Fürsich, F. T., Harries, P. J., Hendy, A. J. W., Holland, S. M., Ivany, L. C., Kiessling, W., Kosnik, M. A., Marshall, C. R., McGowan, A. J., Miller, A. I., Olszewski, T. D., Patzkowsky, M. E., Peters, S. E., Villier, L., Wagner, P. J., Bonuso, N., Borrow, P. S., Brenneis, B., Clapham, M. E., Fall, L. M., Ferguson, C. A., Hanson, V. L., Krug, A. Z., Layou, K. M., Leckey, E. H., Nürnberg, S., Powers, C. M., Sessa, J. A., Simpson, C., TomasnOvyach, A., and Visaggi, C. C. 2008. Phanerozoic trends in the global diversity of marine invertebrates. Science, 321:97100.CrossRefGoogle ScholarPubMed
Ausich, W. I., and Bottjer, D. J. 1982. Tiering in suspension-feeding communities on soft substrata throughout the Phanerozoic. Science, 216:173174.Google Scholar
Ausich, W. I., and Bottjer, D. J. 1985. Phanerozoic tiering in suspension-feeding communities on soft substrata: implications for diversity, p. 255274. In Valentine, J. W. (ed.), Phanerozoic Diversity Patterns. Princeton University Press, Princeton, New Jersey.Google Scholar
Bambach, R. K. 1977. Species richness in marine benthic habitats through the Phanerozoic. Paleobiology, 3:152167.CrossRefGoogle Scholar
Bambach, R. K. 1983. Ecospace utilization and guilds in marine communities through the Phanerozoic, p. 719746. In Tevesz, M. J. S. and McCall, P. L. (eds.), Biotic Interactions in Recent and Fossil Benthic Communities. Plenum, New York.Google Scholar
Bambach, R. K. 1985. Classes and adaptive variety: the ecology of diversification in marine faunas through the Phanerozoic, p. 191253. In Valentine, J. W. (ed.), Phanerozoic Diversity Patterns. Princeton University Press, Princeton, New Jersey.Google Scholar
Bambach, R. K., Knoll, A. H., and Sepkoski, J. J. Jr. 2002. Anatomical and ecological constraints on Phanerozoic animal diversity in the marine realm. Proceedings of the National Academy of Sciences, 99:68546959.CrossRefGoogle ScholarPubMed
Bambach, R. K., Bush, A. M., and Erwin, D. H. 2007. Autecology and the filling of ecospace: key metazoan radiations. Palaeontology, 50:122.Google Scholar
Baumiller, T. K., and Gahn, F. J. 2004. Testing predator-driven evolution with Paleozoic crinoid arm regeneration. Science, 305:14531455.Google Scholar
Behrensmeyer, A. K., Fürsich, F. T., Gastaldo, R. A., Kidwell, S. M., Kosnik, M. A., Kowalewski, M., Plotnick, R. E., Rogers, R. R., and Alroy, J. 2005. Are the most durable shelly taxa also the most common in the marine fossil record? Paleobiology, 31:607623.Google Scholar
Bradfield, G. E., and Kenkel, N. C. 1987. Nonlinear ordination using flexible shortest path adjustment of ecological distances. Ecology, 68:750753.Google Scholar
Brett, C. E., and Baird, G. C. 1995. Coordinated stasis and evolutionary ecology of Silurian to Middle Devonian faunas in the Appalachian Basin, p. 285315. In Erwin, D.H. and Anstey, R.L. (eds.), New Approaches to Speciation in the Fossil Record. Columbia University Press, New York.Google Scholar
Bush, A. M., and Bambach, R. K. 2004. Did alpha diversity increase during the Phanerozoic? Lifting the veils of taphonomic, latitudinal, and environmental biases in the study of paleocommunities. Journal of Geology, 112:625642.Google Scholar
Bush, A. M., and Brame, R. I. 2007. Multivariate ordination of an opportunist-dominated fauna from the Frasnian of Virginia: finding the paleoecologic gradients. Geological Society of America Abstracts with Programs, 39:88.Google Scholar
Bush, A. M., Bambach, R. K., and Daley, G. M. 2007a. Changes in theoretical ecospace utilization in marine fossil assemblages between the mid-Paleozoic and late Cenozoic. Paleobiology, 33:7697.Google Scholar
Bush, A. M., Bambach, R. K., and Daley, G. M. 2007b. Paleoecology of the Phanerozoic increase in alpha diversity. Geological Society of America Abstracts with Programs, 39:587.Google Scholar
Bush, A. M., Kowalewski, M., Hoffmeister, A., Bambach, R. K., and Daley, G. M. 2007c. Potential paleoecologic biases from size-filtering of fossils. Palaios, 22:612622.CrossRefGoogle Scholar
Bush, A. M., Bambach, R. K., and Daley, G. M. 2008. Were local ecological interactions linked to secular trends in alpha diversity in level-bottom marine communities? Geological Society of America Abstracts with Programs, 40:in press.Google Scholar
Cherns, L., and Wright, V. P. 2000. Missing molluscs as evidence of large-scale, early skeletal aragonite dissolution in a Silurian sea. Geology, 28:791794.Google Scholar
Clementz, M. T., Hoppe, K. A., and Koch, P. L. 2003. A paleoecological paradox: the habitat and dietary preferences of the extinct tethythere Desmostylus, inferred from stable isotope analysis. Paleobiology, 29:506519.Google Scholar
Colwell, R. K., Rahbek, C., and Gotelli, N. 2004. The mid-domain effect and species richness patterns: what have we learned so far? American Naturalist, 163:E1E23.Google Scholar
Daley, G. M. 1999. Paleocommunities of the Yorktown Formation (Pliocene) of Virginia. Unpublished Ph.D. dissertation, Virginia Polytechnic Institute and State University, 299 p.Google Scholar
De'Ath, G. 1999. Extended dissimilarity: a method of robust estimation of ecological distances from high beta diversity data. Plant Ecology, 144:191199.CrossRefGoogle Scholar
Dietl, G. P., Kelley, P. H., Barrick, R., and Showers, W. 2002. Escalation and extinction selectivity: morphology versus isotopic reconstruction of bivalve metabolism. Evolution, 56:284291.Google Scholar
Duplessis, M. R., Dufour, S. C., Blankenship, L. E., Felbeck, H. and Yayanos, A. A. 2004. Anatomical and experimental evidence for particulate feeding in Lucinoma aequizonata and Parvilucina tenuisculpta (Bivalvia: Lucinidae) from the Santa Barbara Basin. Marine Biology, 145:551561.CrossRefGoogle Scholar
Foote, M. 1993. Contributions of individual taxa to overall morphological disparity. Paleobiology, 19:403419.Google Scholar
Fortey, R. A., and Owens, R. M. 1999. Feeding habits in trilobites. Palaeontology, 42:429465.CrossRefGoogle Scholar
Gould, S. J. 1980. The promise of paleobiology as a nomothetic, evolutionary discipline. Paleobiology, 6:96118.CrossRefGoogle Scholar
Graham, M. H., Kinlan, B. P., Druehl, L. D., Garske, L. E., and Banks, S. 2007. Deep-water kelp refugia as potential hotspots of tropical marine diversity and productivity. Proceedings of the National Academy of Sciences, 104:1657616580.Google Scholar
Hill, M. O., and Gauch, H. G. Jr. 1980. Detrended correspondence analysis: an improved ordination technique. Vegetatio, 42:4758.CrossRefGoogle Scholar
Holland, S. M. 2005. The signatures of patches and gradients in ecological ordinations. Palaios, 20:573580.Google Scholar
Holland, S. M., and Patzkowsky, M. E. 2007. Gradient ecology of a biotic invasion: biofacies of the type Cincinnatian Series (Upper Ordovician), Cincinnati, Ohio region, USA. Palaios, 22:392407.Google Scholar
Holland, S. M., Miller, A. I., Meyer, D. L., and Dattilo, B. F. 2001. The detection and importance of subtle biofacies within a single lithofacies: the Upper Ordovician Kope Formation of the Cincinnati, Ohio region. Palaios, 16:205217.2.0.CO;2>CrossRefGoogle Scholar
Huntley, J. W., and Kowalewski, M. 2007. Strong coupling of predation intensity and diversity in the Phanerozoic fossil record. Proceedings of the National Academy of Sciences, 104:1500615010.Google Scholar
Hurlbert, S. H. 1971. The nonconcept of species diversity: a critique and alternative parameters. Ecology, 52:577586.Google Scholar
Jackson, D. A., and Somers, K. M. 1991. Putting things in order: the ups and downs of defended correspondence analysis. American Naturalist, 137:704712.Google Scholar
Kamermans, P. 1994. Similarity in food source and timing of feeding in deposit- and suspension-feeding bivalves. Marine Ecology Progress Series, 104:6375.CrossRefGoogle Scholar
Kenkel, N. C., and Orlóci, L. 1986. Applying metric and nonmetric multidimensional scaling to ecological studies: some new results. Ecology, 67:919928.Google Scholar
Kidwell, S. M. 2001. Preservation of species abundance in marine death assemblages. Science, 294:10911094.Google Scholar
Kidwell, S. M. 2002. Time-averaged molluscan death assemblages: palimpsests of richness, snapshots of abundance. Geology, 30:803806.Google Scholar
Kowalewski, M., Dulai, A., and Fürsich, F. T. 1998. A fossil record full of holes: the Phanerozoic history of drilling predation. Geology, 26:10911094.Google Scholar
Kowalewski, M., Gürs, K., Nebelsick, J. H., Oschmann, W., Piller, W. E., and Hoffmeister, A. P. 2002. Multivariate hierarchical analyses of Miocene mollusk assemblages of Europe: paleogeographic, paleoecological, and biostratigraphic implications. Geological Society of America Bulletin, 114:239256.Google Scholar
Kowalewski, M., Kiessling, W., Aberhan, M., Fürsich, F. T., Scarponi, D., Barbour Wood, S. L., and Hoffmeister, A. P. 2006. Ecological, taxonomic, and taphonomic components of the post-Paleozoic increase in sample-level species diversity of marine benthos. Paleobiology, 32:533561.Google Scholar
Krueger, D. M., Gallager, S. M., Cavanaugh, C. M. 1992. Suspension feeding on phytoplankton by Solemya velum, a symbiont-containing clam. Marine Ecology Progress Series, 86:145151.Google Scholar
Kruskal, J. B. 1964a. Multidimensional scaling by optimizing goodness of fit to a nonmetric hypothesis. Psychometrika, 29:127.Google Scholar
Kruskal, J. B. 1964b. Nonmetric multidimensional scaling: a numerical method. Psychometrika, 29:115129.CrossRefGoogle Scholar
Labarbera, M. 1981. The ecology of Mesozoic Gryphaea, Exogyra, and Ilymatogyra (Bivalvia: Mollusca) in a modern ocean. Paleobiology, 7:510526.Google Scholar
Legendre, P., and Legendre, L. 1998. Numerical Ecology, 2nd ed. Elsevier, Amserdam, 853 p.Google Scholar
Levinton, J. S. 1970. The paleoecological significance of opportunistic species. Lethaia, 3: 6978.Google Scholar
Levinton, J. S. 1991. Variable feeding behavior in three species of Macoma (Bivalvia: Tellinacea) as a response to water flow and sediment transport. Marine Biology, 110:375383.CrossRefGoogle Scholar
Levinton, J. S., and Bambach, R. K. 1975. A comparative study of Silurian and Recent deposit-feeding bivalve communities. Paleobiology, 1:97124.CrossRefGoogle Scholar
Lockwood, R. 2004. The K/T event and infaunality: morphological and ecological patterns of extinction and recovery in veneroid bivalves. Paleobiology, 30:507521.2.0.CO;2>CrossRefGoogle Scholar
Madin, J. S., Alroy, J., Aberhan, M., Fürsich, F. T., Kiessling, W., Kosnik, M. A., and Wagner, P. J. 2006. Statistical independence of escalatory ecological trends in Phanerozoic marine invertebrates. Science, 312:897900.Google Scholar
Marshall, C. R. 2006. Explaining the Cambrian “Explosion” of animals. Annual Review of Earth and Planetary Science, 34:355384.Google Scholar
McCune, B., and Grace, J. B. 2002. Analysis of Ecological Communities. MjM Software Design, Gleneden Beach, Oregon, 304 p.Google Scholar
McGhee, G. R. Jr. 1998. Theoretical Morphology. Columbia University Press, New York, 316 p.Google Scholar
McGhee, G. R. Jr., Sheehan, P. M., Bottjer, D. J., Droser, M. L. 2004. Ecological ranking of Phanerozoic biodiversity crises: ecological and taxonomic severities are decoupled. Palaeogeography, Palaeoclimatology, Palaeoecology, 211:289297.Google Scholar
Minchin, P. R. 1987. An evaluation of the relative robustness of techniques for ecological ordination. Vegetatio, 69:89107.Google Scholar
Niklas, K. J. 1994. Morphological evolution through complex domains of fitness. Proceedings of the National Academy of Sciences, 91:67726779.CrossRefGoogle ScholarPubMed
Niklas, K. J. 2004. Computer models of early land plant evolution. Annual Review of Earth and Planetary Science, 32:4766.Google Scholar
Novack-Gottshall, P. M. 2007. Using a theoretical ecospace to quantify the ecological diversity of Paleozoic and modern marine biotas. Paleobiology, 33:273294.Google Scholar
Novack-Gottshall, P. M. 2008. The origin of adaptive zones: comparative ecological diversity (richness and disparity) of higher taxonomic categories. Geological Society of America Abstracts with Programs, 39:91.Google Scholar
Peters, S. E. 2004. Evenness of Cambrian-Ordovician benthic marine communities in North America. Paleobiology, 30:325–246.Google Scholar
Peters, S. E. 2006. Genus richness in Cambrian-Ordovician benthic marine communities in North America. Palaios, 21:580587.Google Scholar
Peters, S. E. 2007. The problem with the Paleozoic. Paleobiology, 33:165181.Google Scholar
Powell, M. G., and Kowalewski, M. 2002. Increase in evenness and sampled alpha diversity through the Phanerozoic: comparison of early Paleozoic and Cenozoic marine fossil assemblages. Geology, 30:331334.Google Scholar
Pruss, S. B., and Bottjer, D. J. 2004. Early Triassic trace fossils of the western United States and their implications for prolonged environmental stress from the end-Permian mass extinction. Palaios, 19:551564.Google Scholar
Project, R. 2007. The R Project for Statistical Computing, v. 2.3. www.rproject.org.Google Scholar
Raup, D. M. 1966. Geometric analysis of shell coiling: general problems. Journal of Paleontology, 40:11781190.Google Scholar
Raup, D. M. 1975. Taxonomic diversity estimation using rarefaction. Paleobiology, 1:333342.Google Scholar
Rayfield, E. J. 2007. Finite element analysis and understanding the biomechanics and evolution of living and fossil organisms. Annual Review of Earth and Planetary Sciences, 35:541576.Google Scholar
Rayfield, E. J., Norman, D. B., Horner, C. C., Horner, J. R., May Smith, P., Thomason, J. J., and Upchurch, P. 2001. Cranial design and function in a large theropod dinosaur. Nature, 409:10331037.Google Scholar
Savarese, M. 1992. Functional analysis of Archaeocyathan skeletal morphology and its paleobiological implications. Paleobiology, 18:464480.Google Scholar
Savarese, M. 1995. Functional significance of regular Archaeocyathan central cavity diameter: a biomechanical and paleoecological test. Paleobiology, 21:356378.Google Scholar
Scarponi, D., and Kowalewski, M. 2004. Stratigraphic paleoecology: bathymetric signatures and sequence overprint of mollusk associations from upper Quaternary sequences of the Po Plain, Italy. Geology, 32:989992.Google Scholar
Scarponi, D., and Kowalewski, M. 2007. Sequence stratigraphic anatomy of diversity patterns: late Quaternary benthic mollusks of the Po Plain, Italy. Palaios, 22:296305.Google Scholar
Sepkoski, J. J. Jr. 1984. A kinetic model of Phanerozoic taxonomic diversity. III. Post-Paleozoic families and mass extinctions. Paleobiology, 10:246267.Google Scholar
Sepkoski, J. J. Jr., and Miller, A. I. 1985. Evolutionary faunas and the distribution of Paleozoic benthic communities in space and time, p. 153190. In Valentine, J. W. (ed.), Phanerozoic Diversity Patterns. Princeton University Press, Princeton, New Jersey.Google Scholar
Sessa, J., Baugh, H. L., Ivany, L., Patzkowsky, M., and Bralower, T. 2007. The ecological dynamics of the first thirty million years of the Cenozoic in the Gulf Coastal Plain. Geological Society of America Abstracts with Programs, 39:589.Google Scholar
Shumay, S. E., Scott, T. M., and Shick, J. M. 1983. The effects of anoxia and hydrogen sulphide on survival, activity and metabolic rate in the coot clam, Mulinia lateralis (Say). Journal of Experimental Marine Biology and Ecology, 71:135146.Google Scholar
Signor, P. W. III, and Brett, C. E. 1984. The mid-Paleozoic precursor to the Mesozoic marine revolution. Paleobiology, 10:229245.Google Scholar
Smith, A. B., and McGowan, A. J. 2008. Temporal patterns of barren intervals in the Phanerozoic. Paleobiology, 34:155161.CrossRefGoogle Scholar
Spencer-Davies, P. 1984. The role of zooxanthellae in the nutritional energy requirements of Pocillopora eydouxi. Coral Reefs, 2:181186.CrossRefGoogle Scholar
Stanley, S. M. 1970. Relation of shell form to life habits of the Bivalvia (Mollusca). Geological Society of America Memoir, 125, 296 p.Google Scholar
Stanley, S. M. 1977. Trends, rates, and patterns of evolution in the Bivalvia, p. 209250. In Hallam, A. (ed.), Patterns of Evolution, as Illustrated by the Fossil Record. Elsevier, Amsterdam.Google Scholar
Stanley, S. M. 2007. An analysis of the history of marine animal diversity. Paleobiology, 33(4s):155.Google Scholar
Stanley, S. M. 2008. Predation defeats competition on the seafloor. Paleobiology, 34:121.Google Scholar
Sutton, R. G., and McGhee, G. R. Jr. 1985. The evolution of Frasnian marine “community-types” of south-central New York, p. 211224. In Woodrow, D. L. and Sevon, W. D. (eds.), The Catskill Delta. Geological Society of America Special Paper, 201.Google Scholar
Ter Braak, C. J. F. 1986. Canonical correspondence analysis: a new eigenvector technique for multivariate direct gradient analysis. Ecology, 67:11671179.Google Scholar
Thayer, C. W. 1979. Biological bulldozers and the evolution of marine benthic communities. Science, 203:458461.Google Scholar
Thayer, C. W. 1983. Sediment-mediated biological disturbance and the evolution of the marine benthos, p. 479625. In Tevesz, M. J. S. and McCall, P. L. (eds.), Biotic Interactions in Recent and Fossil Benthic Communities. Plenum, New York.Google Scholar
Twitchett, R. J., Krystyn, L., Baud, A., Wheeley, J. R., and Richoz, S. 2004. Rapid marine recovery after the end-Permian mass-extinction event in the absence of marine anoxia. Geology, 32:805808.Google Scholar
Vermeij, G. J. 1977. The Mesozoic marine revolution: evidence from snails, predators, and grazers. Paleobiology, 3:245258.Google Scholar
Vermeij, G. J. 1987. Evolution and Escalation: an Ecological History of Life. Princeton University Press, Princeton, New Jersey, 544 p.Google Scholar
Ward, L. W. and Blackwelder, B. W. 1980. Stratigraphic revision of Upper Miocene and Lower Pliocene beds of the Chesapeake Group, Middle Atlantic Coastal Plain. U. S. Geological Survey Bulletin, 1482-D, 61 p.Google Scholar
Wartenberg, D., Ferson, S., Rohlf, F. J. 1987. Putting things in order: a critique of detrended correspondence analysis. American Naturalist, 129:434448.CrossRefGoogle Scholar
Williamson, M. H. 1978. The ordination of incidence data. Journal of Ecology, 66:911920.Google Scholar
Williamson, M. H. 1983. The land-bird community of Skokholm: ordination and turnover. Oikos, 41:378384.Google Scholar
Winn, R. N., and Knott, D. M. 1992. An evaluation of the survival of experimental populations exposed to hypoxia in the Savannah River estuary. Marine Ecology Progress Series, 88:161179.Google Scholar
Zuschin, M., Harzhauser, M., and Mandic, O. 2005. Influence of size-sorting on diversity estimates from tempestitic shell beds in the middle Miocene of Austria. Palaios, 20:142158.Google Scholar