Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-25T08:37:14.234Z Has data issue: false hasContentIssue false

Macrostratigraphy and Its Promise for Paleobiology

Published online by Cambridge University Press:  21 July 2017

Shanan E. Peters*
Affiliation:
Department of Geology & Geophysics University of Wisconsin-Madison 1215 W. Dayton St. Madison, WI 53706
Get access

Abstract

Macrostratigraphy is the study and statistical analysis of sediment packages that formed continuously at a specified scale of temporal resolution and that are bound by gaps recognizable at that same scale. The temporal ranges of gap-bound packages, compiled separately for different geographic locations, permit area-weighted, survivorship-based measures of rock quantity and spatio-temporal environmental continuity to be measured. Analytical basin fill models suggest that the parameters controlling sedimentation and sequence stratigraphic architecture, such as base level and sediment supply, can be detected quantitatively by macrostratigraphy.

Macrostratigraphic analysis of the marine sedimentary rock record in the United States at a temporal resolution of ~106 years reproduces most of the well-known Sloss sequences, but it also identifies two prominent megasequences, the Paleozoic and Modern megasequences, which are separated by a Permian-Triassic discontinuity and Phanerozoic minimum in rock quantity. Many short- and long-term features of the macroevolutionary history of marine animals are reproduced by macrostratigraphy, including 1) many patterns in genus richness, 2) patterns in rates of genus extinction and, to a lesser degree, rates of origination, and 3) patterns of extinction selectivity and the shifting relative richness of Sepkoski's Paleozoic and Modern evolutionary faunas. The extent to which macrostratigraphy reproduces the macroevolutionary history of marine animals transcends what is expected by geologically-controlled sampling biases. Instead, the processes which control the spatio-temporal dynamics of shelf sedimentation, including expansions and contractions of shallow epicontinental seas, have probably exerted a consistent influence on the macroevolutionary history of marine animals. Exploring the common cause hypothesis by putting fossils back into rocks and rocks into a new quantitative framework for physical environmental change holds considerable promise for paleobiology.

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

Algeo, T. J., Lyons, T. W., Blakey, R. C., and Over, D. J. 2007. Hydographic conditions of the Devono-Carboniferous North American seaway inferred from sedimentary Mo-TOC relationships: Palaeogeography, Palaeoclimatology, Palaeoecology, 256: 204230.Google Scholar
Algeo, T. J., Heckel, P. H., Maynard, J. B., Blakey, R. C., and Rowe, H. In press. Modern and ancient epicontinental seas and the superesturaine circulation model of marine anoxia. In Holmden, C. and Pratt, B., eds. Geology of Epeiric Seas: Geological Association of Canada.Google Scholar
Allen, J. R. L. 1986. Earthquake magnitude-frequency, epicentral distance, and soft-sediment deformation in sedimentary basins. Sedimentary Geology, 46:6775.Google Scholar
Allen, P. A., and Allen, J. R. 2005. Basin analysis: principles and applications. Wiley-Blackwell, Oxford, 560 p.Google Scholar
Aller, R. C. 1982. The effects of macrobenthos on chemical properties of marine sediment and overlying water, p. 53102. In McCall, P. and Tevesz, M., (eds.), Animal-sediment relations. Plenum Press, New York, New York.Google Scholar
Aller, J. Y., Woodin, A., and Aller, R. C., (eds.). 2002. Organism-sediment interactions. University of South Carolina Press, 403 p.Google Scholar
Allison, P. A., and Briggs, D. E. G. 1993. Paleolatitudinal sampling bias, Phanerozoic speciesdiversity, and the end-Permian extinction. Geology, 21:6568.Google Scholar
Alroy, J., Marshall, C. R., Bambach, R. K., Bezusko, K., Foote, M., Fursich, 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., 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 of the United States of America, 98:62616266.Google Scholar
Allison, P. A., and Wright, V. P. 2005. Switching off the carbonate factory: a-tidality, stratification and brackish wedges in epeiric seas. Sedimentary Geology, 179:175184.Google Scholar
Alvarez, L. W., Alvarez, W., Asaro, F., and Michel, Hv. 1980. Extraterrestrial cause for the Creataceous-Tertiary extinction. Science, 208:10951108.Google Scholar
Amundson, R., Richter, D. D., Humphreys, G. S., Jobbagy, E. G., and Gaillardet, J. 2007. Coupling between biota and Earth materials in the critical zone. Elements, 3:327332.Google Scholar
Anders, M. H., Krueger, S. W., and Sadler, P. M. 1987. A new look at sedimentation rates and the completeness of the stratigraphic record. Journal of Geology, 95:114.Google Scholar
Bambach, R. K. 1977. Species richness in marine habitats through the Phanerozoic. Paleobiology, 3:152167.Google Scholar
Bambach, R. K. 2006. Phanerozoic biodiversity mass extinctions. Annual Review of Earth and Planetary Sciences, 34:127155.Google Scholar
Bambach, R. K., Knoll, A. H., and Wang, S. 2004. Origination, extinction, and mass depletions of marine diversity. Paleobiology, 30:522542.Google Scholar
Behrenfeld, M., O'Malley, R., Siegel, D., McClain, C., Sarmiento, J., Feldman, G., Milligan, A., Falkowski, P., Letelier, R., and Boss, E. 2006. Climate-driven trends in contemporary ocean productivity. Nature, 444:752755.Google Scholar
Behrensmeyer, A. K., Kidwell, S. M., and Gastaldo, R. A. 2000. Taphonomy and paleobiology. Paleobiology, 26:103147.Google Scholar
Berner, R. A. 2003. The long-term carbon cycle, fossil fuels, and atmospheric composition. Nature, 426:323326.Google Scholar
Berner, R. A., Beerling, D. J., Dudley, R., Robinson, J. M., and Wildman, R. A. Jr. 2003. Phanerozoic atmospheric oxygen. Annual Review of Earth and Planetary Sciences, 31:105134.Google Scholar
Berner, R. A., and Canfield, D. E. 1989. A new model of atmospheric oxygen over Phanerozoic time. American Journal of Science, 289:333–61.Google Scholar
Berry, J. P., and Wilkinson, B. H. 1994. Paleoclimatic and tectonic control on the accumulation of North American cratonic sediment. Geological Society of America Bulletin, 106:855865.Google Scholar
Blatt, H., and Jones, R. L. 1975. Proportions of exposed igneous, metamorphic, and sedimentary rocks. Geological Society of America Bulletin, 86:10851088.Google Scholar
Bond, D. P. G., and Wignall, P. B. 2008. The role of sea-level change and marine anoxia in the Frasnian-Famennian (Late Devonian) mass extinction. Palaeogeogrpahy, Palaeoclimatology, Palaeoecology, 263:107118.Google Scholar
Bonelli, J. R., Brett, C. E., Miller, A. I., and Bennington, J. B. 2006. Testing for faunal stability across a regional biotic transition: quantifying stasis and variation among recurring coral-rich biofacies in the Middle Devonian Appalachian Basin.Google Scholar
Bonnet, C., Malavieille, J., and Mosar, J. 2007. Interactions between tectonics, erosion, and sedimentation during the recent evolution of the Alpine orogen: analogue modeling insights. Tectonics, 26.Google Scholar
Brenchley, P. J., Marshall, J. D., Carden, G. A. F., Robertson, D. B. R., Long, D. G. F., Meidla, T., Hints, L., and Anderson, T. F. 1994. Bathymetric and isotopic evidence for a short-lived Late Ordovician glaciation in a greenhouse period. Geology, 4:295298.Google Scholar
Brett, C. E. 1998. Sequence stratigraphy, paleoecology, and evolution: biotic clues and responses to sea level fluctuations. Palaios, 3:241262.Google Scholar
Brett, C. E., Allison, P. A., Tsujita, C. J., Soldani, D., and Moffat, H. A. 2006. Sedimentology, taphonomy, and paleoecology of meter-scale cycles from the Upper Ordovician of Ontario. Palaios, 21:530547.CrossRefGoogle Scholar
Brett, C. E., Hendy, A. J. W., Bartholomew, A. J., Bonelli, J. R., and McLaughlin, P. I. 2007. Response of shallow marine biotas to sea-level fluctuations: a review of faunal replacement and the process of habitat tracking. Palaios, 22:228244.Google Scholar
Brockwell, R. A., and Davis, R. A. 1991. Time series: theory and methods. Springer, New York, 577 p.Google Scholar
Burne, R. V., and Moore, L. S. 1987. Microbialites: organosedimentary deposits of benthic microbial communities. Palaios, 2:241254.Google Scholar
Bush, A. M. 2006. Extinction: how life on earth nearly ended 250 million years ago (review). Science, 311:18681869.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. Journal of Geology, 112:625642.Google Scholar
Bush, A. M., Markey, M. J., and Marshall, C. R. 2004. Removing bias from diversity curves: the effects of spatially organized biodiversity on sampling standardization. Paleobiology, 30:666686.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
Childs, O. E. 1985. Correlation of stratigraphic units of North America: COSUNA. AAPG Bulletin, 69:173180.Google Scholar
Chorover, J., Kretzschmar, R., Garciapichel, F., and Sparks, D. L. 2007. Soil biogeochemical processes within the critical zone. Elements, 3:321326.Google Scholar
Coblentz, D. D., and Riiters, K. H. 2004. Topographic controls on the regional-scale biodiversity of the south-western USA. Journal of Biogeography, 31:11251138.Google Scholar
Cobbold, P. R., Davy, P., Gapais, D., Rossello, A., Sadybakasov, E., Thomas, J. C., Tondji Biyo, J. J., and De Urreiztieta, M. 1993. Sedimentary basins and crustal thickening. Sedimentary Geology, 86:7789.Google Scholar
Coe, A. L. (ed.). 2003. The sedimentary record of sea-level change. Cambridge University Press, Cambridge, 288 p.Google Scholar
Cook, T. D., and Bally, A. W., eds. 1975. Stratigraphic Atlas of North and Central America. Princeton University Press, Princeton, New Jersey.Google Scholar
Crampton, S. C., Beu, A. G., Cooper, R. A., Jones, C. M., Marshall, B., and Maxwell, P. A. 2003. Estimating the rock volume bias in paleobiodiversity studies. Science, 301:358360.Google Scholar
Dashtgard, S. E., Gingras, M. K., and Pemberton, S. G. 2007. Grain-size controls on the occurrence of bioturbation. Palaeogeography, Palaeoclimatology, Palaeoecology, 257:224243.Google Scholar
Dietrich, W. E., and Perron, J. T. 2006. The search for a topographic signature of life. Nature, 439:411418.Google Scholar
Droser, M. L., and Finnegan, S. 2003. The Ordovician radiation: a follow-up to the Cambrian explosion? Integrative and Comparative Biology, 43:178184.Google Scholar
Einsele, G. 2000. Sedimentary basins: evolution, facies and sediment budget. Springer, Berlin, Germany, 792 p.Google Scholar
Erwin, D. H. 2006. Extinction: how life on earth nearly ended 250 million years ago. Princeton University Press, Princeton, 296 p.Google Scholar
Flessa, K. W., and Jablonski, D. 1985. Declining Phanerozoic background extinction rates—effects of taxonomic structure. Nature, 313:216218.Google Scholar
Foote, M. 1988. Survivorship analysis of Cambrian and Ordovician trilobites. Paleobiology, 14:258271.Google Scholar
Foote, M. 2000. Origination and extinction components of taxonomic diversity: general problems. Paleobiology, 26 (Suppl. to no. 4):74102.Google Scholar
Foote, M. 2003. Origination and extinction through the Phanerozoic: a new approach. Journal of Geology, 111:125148.Google Scholar
Gabet, E. J., Reichman, O. J., and Seabloom, E. W. 2003. The effects of bioturbation on soil processes and sediment transport. Annual Review of Earth and Planetary Sciences, 31:249273.Google Scholar
Gaines, R. R., and Droser, M. L. 2003. Paleoecology of the familiar trilobite Elrathia kingii: an early exaerobic zone inhabitant. Geology 31,941944.Google Scholar
Gerbersdorf, S. U., Jancke, T., Westrich, B., and Paterson, D. M. 2008. Microbial stabilization of riverine sediments by extracellular polymeric substances. Geobiology, 6:5769.Google Scholar
Gilinsky, N. L. and Bambach, R. K. 1987. Asymmetrical patterns of origination and extinction in higher taxa. Paleobiology, 13:427445.Google Scholar
Gregor, C. B. 1968. The rate of denudation in post-Algonkian time. Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen. Series B: Palaeontology, Geology, Physics, Chemistry, Anthropology, 71:2230.Google Scholar
Hallam, A. 1989. The case for sea-level change as a dominant casual factor in mass extinction. Proceedings of the Royal Society of London B, 325:437455.Google Scholar
Hallam, A. 1992. Phanerozoic sea-level changes. Columbia University Press, New York, New York, 266 p.Google Scholar
Hallam, A., and Wignall, P. B. 1997. Mass extinctions and their aftermath. Oxford University Press, Oxford, 320 p.Google Scholar
Hallam, A., and Wignall, P. B. 1999. Mass extinction and sea-level changes. Earth-Science Reviews, 48:217250.Google Scholar
Hannisdal, B. 2006. Phenotypic evolution in the fossil record: numerical experiments. Journal of Geology, 114:133153.Google Scholar
Heller, P. L., and Angevine, C. L. 1985. Sea-level cycles during the growth of Atlantic-type oceans. Earth and Planetary Science Letters, 75:417426.Google Scholar
Holland, S. M. 1999. The New Stratigraphy and its promise for paleobiology. Paleobiology, 25:409416.Google Scholar
Holland, S. M. 2000. The quality of the fossil record: a sequence stratigraphic perspective. Paleobiology, 26:148168.Google Scholar
Holland, S. M. 2005. The signature of patches and gradients in ecological ordinations. Palaios, 20:573580.Google Scholar
Holland, S. M., and Patzkowsky, M. E. 1999. Models for simulating the fossil record. Geology, 27:491494.Google Scholar
Holland, S. M., and Patzkowsky, M. E. 2002. Stratigraphic variation in the timing of first and last occurrences. Palaios, 17:134146.Google Scholar
Holland, S. M., and Patzkowsky, M. E. 2004. Ecosystem structure and stability: Middle Ordovician of central Kentucky, USA. Palaios, 19:316331.Google Scholar
Holland, S. M., Meyer, D. L., and Miller, A. I. 2000. High-resolution correlation in apparently monotonous rocks: Upper Ordovician Kope Formation, Cincinnati Arch. Palaios, 15:7380.2.0.CO;2>CrossRefGoogle ScholarPubMed
Holland, S. M., Miller, A. I., Miller, 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.Google Scholar
House, M. R., and Gale, A. S. 1995. Orbital forcing timescales and cyclostratigraphy. Geological Society Special Publication 85, 226 p.Google Scholar
Hunda, B. R., Hughes, N. C., and Flessa, K. W. 2006. Trilobite taphonomy and temporal resolution in the Mt. Orab shale bed (Upper Ordovician, Ohio, U.S.A.). Palaios, 21:2645.Google Scholar
Jablonski, D., Roy, K., Valentine, J. W., Price, R. M., and Anderson, P. S. 2003. The impact of the pull of the recent on the history of marine diversity. Science, 300:11331135.Google Scholar
Johnson, J. G. 1974. Extinction of perched faunas. Geology, 2:479482.Google Scholar
Johnson, J. G., Klapper, G. C. A. and Sandberg, . 1985. Devonian eustatic fluctuations in Euramerica. Geological Society of America Bulletin, 96:567587.Google Scholar
Kidwell, S. M. 1985. Palaeobiological and sedimentological implications of fossil concentrations. Nature, 318:457460.Google Scholar
Kidwell, S. M. 1986. Taphonomic feedback in Miocene assemblages: testing the role of dead hardparts in benthic communities. Palaios, 1:239255.Google Scholar
Kidwell, S. M., and Jablonski, D. 1983. Taphonomic feedback: ecological consequences of shell accumulation, p. 195248. In Tevesz, M. J. S. and McCall, P. L. (eds.), Biotic interactions in Recent and fossil benthic communities. Plenum Press, New York, New York.Google Scholar
Kidwell, S. M., and Holland, S. M. 2002. The quality of the fossil record: implications for evolutionary analyses. Annual Review of Ecology and Systematics, 33:561588.Google Scholar
Kidwell, S. M., Best, M. M. R., Kaufman, D. S. 2005. Taphonomic trade-offs in tropical marine death assemblages: differential time averaging, shell loss, and probable bias in siliciclastic vs. carbonate facies. Geology, 33:729732.Google Scholar
Kiessling, W., Flugel, E., and Golonka, J. 2003. Patterns of Phanerozoic carbonate platform sedimentation. Lethaia, 36:195225.Google Scholar
Kosnik, M. A., Hua, Q., Jacobsen, G. E., Kaufman, D. S., and Wuest, R. A. 2007. Sediment mixing and stratigraphic disorder revealed by the age-structure of Tellina shells in Great Barrier Reef sediment. Geology, 35:811814.Google Scholar
Kowalewski, M., Goodfriend, G. A., and Flessa, K. W. 1998. High-resolution estimates of temporal mixing within shell beds: the evils and virtues of time-averaging. Paleobiology, 24:287304.Google Scholar
Kowalewski, M., Kiessling, W., Aberhan, M., Fürsich, F. T., Scarponi, D., Wood, S. L. B., 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
Knoll, A. H., Bambach, R. K., Payne, J. L., Pruss, S., and Fischer, W. W. 2007. Paleophysiology and the end-Permian mass extinction. Earth and Planetary Science letters, 256:295313.Google Scholar
Knoll, M. A., James, W. C. 1987. Effect of the advent and diversification of vascular land plants on mineral weathering through geologic time. Geology, 15:10991102.Google Scholar
Lavier, L. L., and Steckler, M.S. 1997. The effects of sedimentary cover on the flexural strength of continental lithosphere. Nature, 389:476479.Google Scholar
Leeder, M. 1999. Sedimentology and sedimentary basins: from turbulence to tectonics. Wiley-Blackwell, Oxford, 608 p.Google Scholar
Maruoka, T., Koeberl, C., Hancox, P. J., and Reimold, W. U. 2002. Sulfur geochemistry across a terrestrial Permian-Triassic boundary section in the Karoo Basin, South Africa. Earth and Planetary Science Letters, 206:101117.Google Scholar
Maslin, M., Owen, M., Day, S., and Long, D. 2004. Linking continental-slope failures and climate change: testing the clathrate gun hypothesis. Geology, 32:5356.Google Scholar
Meysman, F., and Middelburg, J. 2006. Bioturbation: a fresh look at Darwin's last idea. Trends in Ecology and Evolution, 21:688695.Google Scholar
McGowan, A. J., and Smith, A. B. 2008. Are global Phanerozoic marine diversity curves truly global? A study of the relationship between regional rock records and global Phanerozoic marine diversity. Paleobiology, 34:80103.Google Scholar
McLaughlin, P. I., and Brett, C. E. 2004. Eustatic and tectonic control on the distribution of marine seismites: examples from the Upper Ordovician of Kentucky, USA. Sedimentary Geology, 168:165192.Google Scholar
Miall, A. D. 1999. Principles of sedimentary basin analysis. Springer, Berlin, Germany, 616 p.Google Scholar
Miller, A. I. 2000. Conversations about Phanerozoic global diversity. Paleobiology 26 (Suppl. to no. 4):5373.Google Scholar
Miller, A. I., and Mao, S. G. 1995. Association of orogenic activity with the Ordovician radiation of marine life. Geology, 23:305308.Google Scholar
Miller, A. I., Holland, S. M., Meyer, D. L., Dattilo, B. F. 2001. The use of faunal gradient analysis for intraregional correlation and assessment of changes in sea-floor topography in the type Cincinnatian. Journal of Geology, 109:603613.Google Scholar
Miller, K. G., Kominz, M. A., Browning, J. V., Wright, J. D., Mountain, G. S., Katz, M. E., Sugarman, P. J., Cramer, B. S., Christie-Blick, N., and Pekar, S. F. 2005. The Phanerozoic record of global sea-level change. Science, 310:12931298.Google Scholar
Müller, R. D., Sdrolias, M., Gaina, C., Steinberger, B., and Heine, C. 2008. Long-term sea-level fluctuations driven by ocean basin dynamics. Science, 319:13571362.Google Scholar
Neave, M., and Abrahams, A. D. 2001. Impact of small mammal disturbances on sediment yield from grassland and shrubland ecoystems in the Chihuahuan Desert. Catena, 4:285303.Google Scholar
Newell, N. D. 1952. Periodicity in invertebrate paleontology. Journal of Paleontology, 26:371385.Google Scholar
Newell, N. D. 1962. Paleontological gaps and geochronology. Journal of Paleontology, 36:592610.Google Scholar
Newell, N. D. 1967. Revolutions in the history of life. Geological Society of America Special Paper, 89:6391.Google Scholar
Paterson, D. M. 1994. Microbiological mediation of sediment structure and behaviour, p. 97109. In Stal, L. J. and Caumette, P. (eds.), Microbial mats. Springer, Berlin, Germany.Google Scholar
Payne, J. L. 2005. Evolutionary dynamics of gastropod size across the end-Permian extinction and through the Triassic recovery interval. Paleobiology, 31:269290.Google Scholar
Payne, J. L., Lehrmann, D. J., Wei, J. Y., Orchard, M. J., Schrag, D. P., Knoll, A. H. 2004. Large perturbations of the carbon cycle during recovery from the end-Permian extinction. Science, 305:506509.Google Scholar
Payne, J. L., Lehrmann, D. J., Follett, D., Seibel, M., Kump, L., Riccardi, A., Altiner, D., Sano, H., and Wei, J. Y. 2007. Erosional truncation of uppermost Permian shallow-marine carbonates and implications for Permian-Triassic boundary events. Geological Society of America Bulletin, 119:771784.Google Scholar
Peters, S. E. 2005. Geological constraints on the macroevolution of marine animals. Proceedings of the National Academy of Science USA, 102:1232612331.Google Scholar
Peters, S. E. 2006a. Macrostratigraphy of North America. Journal of Geology, 114:391412.Google Scholar
Peters, S. E. 2006b. Genus extinction, origination, and the durations of sedimentary hiatuses. Paleobiology, 32:387407.Google Scholar
Peters, S. E. 2008. Environmental determinants of extinction selectivity in the fossil record. Nature.Google Scholar
Peters, S. E., and Foote, M. 2001. Biodiversity in the Phanerozoic: a reinterpretation. Paleobiology, 27:583601.Google Scholar
Peters, S. E., and Foote, M. 2002. Determinants of extinction in the fossil record. Nature, 416:420424.Google Scholar
Pimentel, D., and Kounang, N. 1998. Ecology of soil erosion in ecosystems. Ecosystems, 1:416426.Google Scholar
R DEVELOPMENT CORE TEAM. 2006. R: A language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria (URL http://www.R-project.org).Google Scholar
Raup, D. M. 1972. Taxonomic Diversity During the Phanerozoic. Science, 177:10651071.Google Scholar
Raup, D. M. 1976. Species diversity in the Phanerozoic: an interpretation. Paleobiology, 2:289297.Google Scholar
Raup, D. M. 1978. Cohort analysis of generic survivorship. Paleobiology, 4:115.Google Scholar
Raup, D. M. 1979. Biases in the fossil record of species and genera. Bulletin of the Carnegie Museum of Natural History, 13:8591.Google Scholar
Raup, D. M., and Sepkoski, J. J. Jr. 1982. Mass Extinctions in the Marine Fossil Record. Science, 215:15011503.Google Scholar
Reynolds, D. J., Steckler, M. S., and Coakley, B. J. 1991. The role of the sediment load in sequence stratigraphy: the influence of flexural isostasy and compaction. Journal of Geophysical Research, 96:69316949.Google Scholar
Rhode, R. A., and Muller, R. A. 2005. Cycles in fossil diversity. Nature, 434:208210.Google Scholar
Ronov, A. B. 1978. The Earth's sedimentary shell. International Geology Review, 24:13131363.Google Scholar
Ronov, A. B., Khain, V. E., Balukhovsky, A. N., and Seslavinsky, K. B. 1980. Quantitative analysis of Phanerozoic sedimentation. Sedimentary Geology, 25:311325.Google Scholar
Rothwell, R. G., Thomson, J., and Kahler, G. 1998. Low sea-level emplacement of a very large Late Pleistocene ‘megaturbidite’ in the western Mediterranean Sea. Nature, 392:377380.Google Scholar
Ruddiman, W. F. 2007. Earth's climate: past and future (second edition). Freeman, New York, New York, 465 p.Google Scholar
Sadler, P. M. 1981. Sediment accumulation rates and the completeness of stratigraphic sections. Journal of Geology, 89:569584.Google Scholar
Sadler, P. M. 2004. Quantitative biostratigraphy—achieving finer resolution in global correlation. Annual Review of Earth and Planetary Science, 32:187213.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. 1976. Species diversity in the Phanerozoic: species-area effects. Paleobiology, 2:298303.Google Scholar
Sepkoski, J. J. Jr. 1981. A factor analytic description of the Phanerozoic marine fossil record. Paleobiology, 7:3653.Google Scholar
Sepkoski, J. J. Jr. 2002. A compendium of fossil marine animal genera. Bulletins of American Paleontology, 363, 560 p.Google Scholar
Simberloff, D. S. 1974. Permo-Triassic extinctions: effects of area on biotic equilibrium. Journal of Geology, 82:267274.Google Scholar
Sloss, L. L. 1963. Sequences in the cratonic interior of North America. Geological Society of America Bulletin, 74:93113.Google Scholar
Sloss, L. L. 1976. Areas and volumes of cratonic sediments, western North America and eastern Europe. Geology, 4:272276.Google Scholar
Smith, A. B. 2001. Large-scale heterogeneity of the fossil record: implications for Phanerozoic biodiversity studies. Philosophical Transactions of the Royal Society of London Series B, 356:351367.Google Scholar
Smith, A. B. 2007. Marine diversity through the Phanerozoic: problems and prospects. Journal of the Geological Society of London, 164:731745.Google Scholar
Smith, A. B., Gale, A. S., and Monks, N. E. A. 2001. Sea-level change and rock-record bias in the Cretaceous: a problem for extinction and biodiversity studies. Paleobiology, 27:241253.Google Scholar
Smith, A. B., and McGowan, A. J. 2005. Cyclicity in the fossil record mirrors rock outcrop area. Biological Letters, 2:13.Google Scholar
Smith, A. B., and McGowan, A. J. 2007. The shape of the Phanerozoic marine palaeodiversity curve: how much can be predicted from the sedimentary rock record of western Europe? Palaeontology, 50:765774.Google Scholar
Smith, A. B., and McGowan, A. J. 2008. Temporal patterns of barren intervals in the Phanerozoic. Paleobiology, 34:155162.Google Scholar
Stanley, S. M. 2007. An analysis of the history of marine animal diversity. Paleobiology 33 (Suppl. to no. 4): 155.Google Scholar
Stanley, S. M. 2008. Predation defeats competition on the seafloor. Paleobiology, 34:121.Google Scholar
Steckler, M. S., Mountain, G. S., Miller, K. G., and Christie-Blick, N. 1999. Reconstruction of Tertiary progradation and clinoform development on the New Jersey passive margin by 2-D backstripping. Marine Geology, 154:399420.Google Scholar
Syvitski, J. P. M., and Hutton, E. W. H. 2001. 2D SEDFLUX 1.0C: an advanced process-response numerical model for the fill of marine sedimentary basins. Computers and Geosciences, 27:731753.Google Scholar
Vail, P. R., Mitchum, R. M., and Thompson, S. 1977. Seismic stratigraphy and global changes of sea-level. American Association of Petroleum Geologists Memoir, 26:8397.Google Scholar
Valentine, J. W. 1971. Plate tectonics and shallow marine diversity and endemism, an actualistic model. Systematic Zoology, 20:253264.Google Scholar
Van Schmus, W. R., and Hinze, W. J. 1985. The midcontinent rift system. Annual Review of Earth and Planetary Sciences, 13:345383.Google Scholar
Van Valen, L. H. 1984. A resetting of Phanerozoic community evolution. Nature, 307:5052.Google Scholar
Van Wagoner, J. C., Posamentier, H. W., Mitchum, R. M. Jr., Vail, P. R., Sarg, J. F., Loutit, T. S., and Hardenbol, J. 1988. An overview of the fundamentals of sequence stratigraphy and key definitions. Pp. 3945. In Wilgus, C. K., Hastings, B. S., Ross, C. A., Posamentier, H. W., Van Wagoner, J., and Kendall, C. G. S. C., (eds.), Special Publication. Society for Sedimentary Geology, Tulsa.Google Scholar
Walker, L. J., Wilkinson, B., and Ivany, L. C. 2002. Continental drift and Phanerozoic carbonate accumulation in shallow-shelf and deep-marine settings. Journal of Geology, 110:7587.Google Scholar
Webster, M., Gaines, R. R., and Hughes, N. C. 2008. Microstratigraphy, trilobite biostratinomy, and depositional environment of the “Lower Cambrian” Ruin Wash Lagerstätte, Pioche Formation, Nevada. Palaeogeography, Palaeoclimatology, Palaeoecology, 264:100122.Google Scholar
Wheeler, H. E. 1964. Baselevel, lithosphere surface, and time-stratigraphy. Geological Society of America Bulletin, 75:599610.Google Scholar
Whitford, W. A., and Kay, F. R. 1999. Biopedturbation by mammals in deserts: a review. Journal of Arid Environments, 41:203230.Google Scholar
Wignall, P. B., and Hallam, A. 1992. Anoxia as a cause of the Permo-Triassic mass extinction: facies evidence from northern Italy and the western United States. Palaeogeogrpahy, Palaeoclimatology, Palaeoecology, 93:2146.Google Scholar
Wignall, P. B., Hallam, A., Lai, X. L., and Yang, F. Q. 1995. Palaeoenvironmental changes across the Permian/Triassic boundary at Shangsi (N. Sichuan, China). Historical Biology, 10:175189.Google Scholar
Wignall, P. B., Newton, R., and Brookfield, M. E. 2005. Pyrite framboid evidence for oxygen-poor deposition during the Permian-Triassic crisis in Kashmir. Palaeogeography, Palaeoclimatology, Palaeoecology, 216:183188.Google Scholar
Wilkinson, B. H. 2005. Humans as geologic agents: a deep-time perspective. Geology, 33:161164.Google Scholar
Wilkinson, B. H., and McElroy, B. 2007. The impact of humans on continental erosion and sedimentation. Bulletin of the Geological Society of America, 119:140156.Google Scholar
Wold, C. N., and Hay, W. W. 1990. Estimating ancient sediment fluxes. American Journal of Science, 290:10691089.Google Scholar
Woodcock, N. H. 2004. Life span and fate of basins. Geology, 32:685688.Google Scholar
Wright, P., Cherns, L., and Hodges, P. 2003. Missing molluscs: Field testing taphonomic loss in the Mesozoic through early large-scale aragonite dissolution. Geology, 31:211214.Google Scholar
Wright, P. V., and Burgess, P. M. 2005. The carbonate factory continuum, facies mosaics and microfacies: an appraisal of some of the key concepts underpinning carbonate. Facies, 51:1723.Google Scholar
Zhang, Y. 2005. Global tectonic and climatic control of mean elevation of continents, and Phanerozoic sea level change. Earth and Planetary Science Letters, 237:524531.Google Scholar
Ziegler, A. M., Eshel, G., Rees, P. M., Rothfus, T. A., Rowley, D. B., and Sunderlin, D. 2003. Tracing the tropics across land and sea: Permian to present. Lethaia, 36:227:254.Google Scholar
Ziegler, P. A. 1982. Geological Atlas of Western and Central Europe. Elsevier, Amsterdam.Google Scholar