Skip to main content
Log in

Biased, Spasmodic, and Ridiculously Incomplete: Sequence Stratigraphy and the Emergence of a New Approach to Stratigraphic Complexity in Paleobiology, 1973–1995

  • Original Research
  • Published:
Journal of the History of Biology Aims and scope Submit manuscript

Abstract

This paper examines the emergence of a new approach to stratigraphic complexity, first in geology and then, following its creative appropriation, in paleobiology. The approach was associated with a set of models that together transformed stratigraphic geology in the decades following 1970. These included the influential models of depositional sequences developed by Peter Vail and others at Exxon. Transposed into paleobiology, they gave researchers new resources for studying the incompleteness of the fossil record and for removing biases imposed by the processes of sedimentary accumulation. In addition, they helped reconfigure the cultural landscape of paleobiology, consolidating a growing emphasis on fieldwork and eroding the barrier that had been erected in the 1970s between “paleontology” and “paleobiology.” This paper traces these developments, paying special attention to the simulation models of stratigraphic paleobiologist Steven Holland. It also considers how the integration of sequence and event stratigraphy and paleobiology has begun to influence long-running discussions of incompleteness and bias in the fossil record.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

Notes

  1. The paleobiological revolution is the name given to a series of developments that saw the emergence of paleobiology as a distinct area of study “centered around the quantitative analysis and interpretation of the history of life” (Sepkoski 2013, p. 402). It is usually dated from about 1970 to 1985 and associated with American invertebrate paleontologists like Gould (1941–2002), David Raup (1933–2015), J. John (Jack) Sepkoski, Jr. (1948–1999), Steven Stanley (1944–), and Thomas Schopf (1939–1984).

  2. Sequence stratigraphy has been almost entirely ignored by historians. The only historian who seems to have noticed it is David Oldroyd (2006). However, its influence on stratigraphic practice has been immense. One commenter has gone so far as to compare this influence to that of plate tectonics on structural geology (Mitchum 2003). Doubtless this is an exaggeration, but even more sober commenters agree that its influence has been pervasive: “Modern stratigraphy is dominated by the study of ‘sequence stratigraphy’” (Miall and Miall 2002, p. 307). A secondary goal of this paper is therefore to draw attention to this crucial development in the recent history of the earth sciences.

  3. Similar criticisms were voiced by German-speaking paleontologists prior to World War Two (Rieppel 2013). A leading figure, the Austrian paleontologist and National Socialist Othenio Abel (1875–1946), even spoke of “the battle to free paleontology from the shackles of geology” (Abel 1929, p. 153), although there is little evidence to suggest that the paleobiology of the 1970s was directly influenced by Abel’s Paläobiologie.

  4. Along with exhibiting this temporal pattern, many stratigraphers expected the rock record to resemble a layer cake in its physical structure as well. This is a more literal, and nowadays universally condemned, application of the layer cake metaphor.

  5. Facies is a Latin word meaning “face” or “external appearance” (Teichert 1958). In geological usage, it means a sedimentary deposit characterized by a set of features that formed in a particular depositional environment, like a coastal plain or reef front.

  6. As the paleontologist and stratigrapher Carlton Brett (1952–) summarized this view, “if a rock unit looks the same in two different places it must be of different ages” (Brett 2000, p. 496).

  7. The middle of the twentieth century was an exceedingly complicated time in the history of stratigraphy, which saw the consolidation of “pure” lithostratigraphy (the project of delineating and correlating rock units based entirely on lithological characteristics) as well as the expansion of petrography and process sedimentology (Seibold and Seibold 2002; Steel and Milliken 2013). (Petrography refers to the descriptive study of rocks, especially under the microscope, whereas process sedimentology refers to the actualistic study of sedimentary bodies and structures.) This section outlines developments in lithostratigraphy to the neglect of these other critical areas (but see Dott 1978 for a complementary account).

  8. This is not to say that biostratigraphic (rock) units were ever explicitly defined as chronostratigraphic (time) units, although in practice they were frequently treated as such (Hedberg 1965). It is just to say that, in the mid-twentieth century, fossils provided the key line of evidence for chronostratigraphic dating.

  9. For example, here is the “grandfather” of American paleobiology, Norman Newell (1919–2004), writing in 1962: “As stratigraphic work in the United States has been increasingly directed to local and minor stratigraphic units there has been a growing emphasis on physical criteria and less attention to fossils… This decline in stratigraphic paleontology has resulted in widespread lack of appreciation of fossils as indices of time and environment, and many stratigraphers relegate fossils to a minor role in classifying and correlating strata” (Newell 1962, p. 592). Ironically, Newell had earlier been keen to articulate worries about the subordination of (invertebrate) paleontology to stratigraphy (Sepkoski 2012, pp. 57–59).

  10. Again, the chronostratigraphic layer cake should be distinguished from the view that the rock record physically resembles a layer cake. The latter analogy, Ager thought, “just will not do” (Ager 1973, p. 75).

  11. These event beds have been likened to frosting layers in a well-marbled cake (Brett 2000). The implication is that even if the cake layers (local facies) are diachronous, stratigraphers can still use the frosting layers to divide the cake into roughly time-parallel units.

  12. A craton is a large and ancient block of crust that comprises the nucleus of a continent (Kay 1974). Cratonic regions are the regions overlying cratons, which contain piles of younger rocks. What Sloss and colleagues showed was that the cratonic region of North America could be subdivided into four “unconformity-bounded successions”: thick packages of strata inferred to have a shared origin in tectonic movements (Sloss et al. 1949). This number was later increased to six in a paper that many regard as the earliest example of the modern “sequence” concept in action (Sloss 1963; see also Sloss 1988).

  13. Seismic reflection data is often presented in the form of seismic [reflection] profiles: visualizations of reflected acoustic energy that picture subsurface structures to depths of tens of kilometers. The innovation that stimulated Vail’s interest was the ability of computer-aided reflection seismology to image subsurface stratification patterns at high levels of resolution (Sloss 1988; see also Oldroyd 2006, pp. 146–147).

  14. This work took place in the 1961.

  15. The extent to which seismic reflectors correspond to chronostratigraphic horizons became an object of controversy in subsequent decades. The most acute point of controversy concerned whether depositional sequences are indeed “geochronologic units” with potentially global validity (Vail et al. 1977b, p. 96). Vail and colleagues argued that they are, and used the postulate of globally synchronous sequence boundaries to construct sea-level curves as a template for dating and correlation (Vail 1977b, c). Critics raised a host of objections: for example, that this application of the sequence model gives too little weight to factors other than global sea-level change in generating sedimentary cycles (Miall 1992; Poulsen et al. 1998). Significantly for the present account, these critics were largely agreed about the utility of the sequence model itself. What they objected to was the notion that sequences are produced by glacially-controlled changes in sea-level, and that this makes them globally correlable (Dewey and Pitman 1998; Dickinson 2003).

  16. Both these applications were ultimately geared toward providing a global framework for petroleum exploration, which would reduce costs from exploratory drilling and increase production profits.

  17. AAPG stands for the American Association of Petroleum Geologists. AAPG Memoir 26 is a collection of twenty-four papers given at a 1975 APPG research symposium, including a series of eleven papers authored by Vail and colleagues on the stratigraphic interpretation of reflection records. For an analysis of its reception and rapid uptake by corporate and academic geologists, see Miall and Miall (2002).

  18. These sequences are considerably smaller than the continent-scale sequences described by Sloss.

  19. To say that a surface is chronostratigraphically significant is to say that all the rocks overlying it are everywhere younger than all the rocks underlying it (which is different from saying that the surface is isochronous, or that it represents a time line). Because of this, chronostratigraphically significant surfaces are often used in local correlation.

  20. More precisely, accommodation [space] is defined as the vertical envelope between the sea surface and the basement of rocks beneath the sedimentary pile, which is available for sedimentation (Jervey 1988). Changes in accommodation reflect the sum of changes in eustatic sea-level change and tectonism, with rising seas and tectonic subsidence increasing accommodation, and falling seas and tectonic elevation decreasing accommodation.

  21. A typical parasequence is between one and ten meters thick and represents tens to hundreds of thousands of years of elapsed time. By contrast, depositional sequences tend to be thicker (comprising multiple stacked parasequences) and represent millions of years of elapsed time (but see Christie-Blick and Driscoll 1995 for complications).

  22. Sequence boundaries typically represent significant periods of time in which no sediment accumulates. But they are not the only chronostratigraphically significant surfaces in a sequence, and other surfaces, like the maximum flooding surface, are also associated with periods of highly reduced deposition.

  23. This is no place to review the development of sequence models since the 1970s. Suffice it to say that the basic model of Vail and others was repeatedly amended as higher-resolution seismic data became available, and as the importance of factors beyond eustatic sea-level change became more widely accepted (Miall and Miall 2001; Embry et al. 2007). A major development was the ability to apply sequence stratigraphy directly to outcrops and well-logs in the absence of seismic reflection data (Van Wagoner et al. 1990). Also important was the development of numerical models of sedimentary accumulation, which helped to clarify the internal structure of sequences and the meaning of key surfaces (Jervey 1988; see also Posamentier and Vail 1988; Van Wagoner et al. 1988). All this was crucial in fashioning sequence stratigraphy into “an entirely new way of practicing [stratigraphic geology]” (Miall and Miall 2001, p. 322).

  24. This process received a great impetus from the controversial Alvarez hypothesis, which sparked renewed interest in the phenomenon of mass extinction (Alvarez et al. 1980; see also Sepkoski 2021). Although the most famous early studies of mass extinction were computer-based (e.g., Raup and Sepkoski 1982), other self-identifying paleobiologists studied mass extinction by taking to the field (e.g., Ward 1983).

  25. Taphonomy refers broadly to the study of “how organic remains are incorporated into the rock record and the fate of these materials after burial” (Behrensmeyer and Kidwell 1985, p. 105). It has been understudied by historians, but cursory historical treatments can be found in Olsen (1980), Behrensmeyer and Kidwell (1985) and Cadée (1991).

  26. Steven Holland (1962–) is an American paleontologist and stratigrapher, who completed his Ph.D. at the University of Chicago under Susan Kidwell in 1990. I have elected to focus on Holland’s 1995 modeling study because it exemplifies the way paleobiologists appropriated resources from the new stratigraphy for distinctively paleobiological ends. But any number of studies from this period might equally have supplied a focus for this section (for example, Kauffman 1984; Kidwell 1986, 1989; MacLeod 1991; Brett and Baird 1992).

  27. To understand this section, it is not necessary to understand in detail how sequence stratigraphy narrates the process of sedimentary basin filling. Nevertheless, here is a quick synopsis. At the base of each sequence is a surface known as the sequence boundary. This forms when relative sea-level is falling. During this interval, no new sedimentation occurs, so sequence boundaries correspond to gaps in the rock record. When relative sea-level begins to rise, deposition is renewed and parasequences stack seaward in a net shallowing pattern. These parasequences form the lowstand systems tract (LST), so named because it sits at a topographically lower position than the rest of the sequence. As relative sea-level rise accelerates such that the rate of sea-level rise exceeds the rate of sedimentation, the pattern of stacking is reversed and successive parasequences exhibit a net deepening trend. This set of parasequences comprises the transgressive systems tract (TST). Separating the LST and the TST is a flooding surface called the transgressive surface, which marks the point at which seaward stacking is replaced by landward stacking, and is often associated with reduced sedimentation (a phenomenon known as condensation). Other flooding surfaces within the TST may also exhibit condensation. Finally, as the rate of sea-level rise begins to slow, parasequences again begin to stack seaward in a net shallowing trend. At this juncture there is another flooding surface, the maximum flooding surface, which records the greatest water depth in the sequence and is often highly condensed. The parasequences deposited atop the maximum flooding surface comprise the highstand systems tract (HST). These are bounded at their top by the sequence boundary, which, again, forms when relative sea-level is falling.

  28. Witness the gibe in a 1969 issue of Nature that “Scientists in general might be excused for thinking that… most paleontologists have staked out a square mile for their life’s work” (Anonymous 1969, p. 903).

  29. Rudwick made a similar observation in his (2018), especially pp. 504–507.

  30. The acronym SEPM refers to the original name of the Society for Sedimentary Geology: the Society of Economic Paleontologists and Mineralogists.

  31. Here it is worth noting that all major figures in 1970s-era paleobiology were trained in traditional paleontological methods. Even Jack Sepkoski, the archetype of the “new model paleontologist,” wrote a dissertation titled, “Stratigraphy and Paleoecology of Dresbachian (Upper Cambrian) Formations in Montana, Wyoming, and South Dakota”—hardly the quantitative work he later became known for (Sepkoski 2005).

  32. This traded on the perception that stratigraphy had become bogged down in minutiae (see “Stratigraphy After 1970”), or as the outspoken stratigrapher P.D. Krynine (1902–1964) is reported to have said, that stratigraphy had seen a “complete triumph of terminology over facts and common sense” (Folk and Ferm 1966, p. 853).

  33. By “trading zone,” I mean a site of substantive and reciprocal communication characterized by an exchange of materials and ideas. For a more explicit treatment, see Collins et al. (2007).

References

  • Abel, Othenio. 1929. Otto Jaekel (21. Februar 1863–6. März 1929). Paläobiologica 2: 143–186.

    Google Scholar 

  • Ager, Derek A. 1973. The nature of the stratigraphic record. New York: Wiley.

    Google Scholar 

  • Ager, Derek A. 1993. The new catastrophism: The importance of the rare event in geological history. Cambridge: Cambridge University Press.

    Google Scholar 

  • Aigner, Thomas. 1985. Storm depositional systems: Dynamic stratigraphy in modern and ancient shallow-marine sequences. Berlin: Springer.

    Google Scholar 

  • Alvarez, Louis W., Walter Alvarez, Frank Asaro, and Helen V. Michel. 1980. Extraterrestrial cause for the end-cretaceous extinction: Experimental results and theoretical interpretation. Science 208: 1095–1108.

    Google Scholar 

  • Anonymous. 1969. What will happen to geology? Nature 221: 903.

    Google Scholar 

  • Bambach, Richard K. 1993. Seafood through time: Changes in biomass, energetics, and productivity in the marine ecosystem. Paleobiology 11: 372–397.

    Google Scholar 

  • Bambach, Richard K. 2009. From empirical paleoecology to evolutionary paleobiology: A personal journey. In The paleobiological revolution: Essays on the growth of modern paleobiology, ed. David Sepkoski and Michael Ruse, 398–415. Chicago: University of Chicago Press.

    Google Scholar 

  • Baron, Christian. 2011. A web of controversies: Complexity in the Burgess Shale debate. Journal of the History of Biology 44: 745–780.

    Google Scholar 

  • Behrensmeyer, Anna K. 1975. The taphonomy and paleoecology of Plio-Pleistocene vertebrate assemblages of Lake Rudolf, Kenya. Bulletin Museum of Comparative Zoology 146: 473–578.

    Google Scholar 

  • Behrensmeyer, Anna K., and Susan M. Kidwell. 1985. Taphonomy’s contributions to paleobiology. Paleobiology 11: 105–119.

    Google Scholar 

  • Benton, Michael J. 1985. Mass extinction among non-marine tetrapods. Nature 316: 811–814.

    Google Scholar 

  • Benton, Michael J. 2013. Origins of biodiversity. Paleontology 56: 1–7.

    Google Scholar 

  • Bhattacharya, Janok P., and Vitor Abreu. 2016. Wheeler’s confusion and the seismic revolution: How geophysics saved stratigraphy. The Sedimentary Record 14: 4–11.

    Google Scholar 

  • Bokulich, Alisa. 2018. Using models to correct data: Paleodiversity and the fossil record. Synthese. https://doi.org/10.1007/s11229-018-1820-x.

    Article  Google Scholar 

  • Bottjer, David J., and David Jablonski. 1988. Paleoenvironmental patterns in the evolution of post-paleozoic benthic marine invertebrates. Palaios 3: 540–560.

    Google Scholar 

  • Brett, Carlton E. 1995. Sequence stratigraphy, biostratigraphy, and taphonomy in shallow marine environments. PALAIOS 10: 597–616.

    Google Scholar 

  • Brett, Carlton E. 1998. Sequence stratigraphy, paleoecology, and evolution: Biotic clues and responses to sea-level fluctuations. PALAIOS 13: 241–262.

    Google Scholar 

  • Brett, Carlton E. 2000. A slice of the “layer cake”: The paradox of “frosting continuity”. PALAIOS 15: 495–498.

    Google Scholar 

  • Brett, Carlton E., and Gordon C. Baird. 1992. Coordinated stasis and evolutionary ecology of silurian to Middle Devonian faunas in the Appalachian Basin. In New approaches to speciation in the fossil record, eds. Douglas H. Erwin, and R.L. Anstey, 285–315. New York: Columbia University Press.

    Google Scholar 

  • Brett, Carlton E., and Gordon C. Baird. 1996. Middle Devonian sedimentary cycles and sequences in the northern Appalachian Basin. In Paleozoic sequence stratigraphy: Views from the North American Craton, ed. Brian Witzke, Greg A. Ludvigson, and Jed Day, 213–241. Geological Society for America, Special Paper 306.

  • Brett, Carlton E., C. Linda, Ivany, and Kenneth M. Schopf. 1996. Coordinated stasis: An overview. Paleogeography Paleoclimatology and Paleoecology 127: 1–20.

    Google Scholar 

  • Brett, Carlton E., I. Patrick, and McLaughlin. 2007. Eo-Ulrichian to neo-ulrichian views: The renaissance of “layer-cake stratigraphy. Stratigraphy 4: 210–215.

    Google Scholar 

  • Brown, Larry D. 2013. From the layer cake to complexity: 50 years of geophysical investigations of the earth. In The web of geological sciences: Advances, impacts, interactions, ed. Marion E. Bickford, 233–258. The Geological Society of America, Special Paper 500.

  • Cadée, Gerhard C. 1991. The history of taphonomy. In The processes of fossilization, ed. S.K. Donovan, 3–21. New York: Columbia University Press.

    Google Scholar 

  • Christie-Blick, Nicholas, and Neal W. Driscoll. 1995. Sequence stratigraphy. Annual Review of Earth and Planetary Sciences 23: 451–471.

    Google Scholar 

  • Collins, Harry, Robert Evans, and Mike Gorman. 2007. Trading zones and interactional expertise. Studies in History and Philosophy of Science 38: 657–666.

    Google Scholar 

  • Cramer, Bradley D., Thijs R.A. Vandenbroucke, and Gregory A. Ludvigson. 2015. High-resolution event stratigraphy (HiRES) and the quantification of stratigraphic uncertainty: Silurian examples of the quest for precision in stratigraphy. Earth-Science Reviews 141: 136–153.

    Google Scholar 

  • Cuffey, Roger J. 1998. An introduction to the type-cincinnatian. In Sampling the layer-cake that isn’t; the stratigraphy and paleontology of the type-cincinnatian, ed. Richard A. Davis, and Roger J. Cuffey, 2–9. Columbus: Department of Natural Resources, Division of Geological Survey.

  • Darwin, Charles. 1859. On the origin of species by means of natural selection, or the preservation of favoured races in the struggle for life. London: John Murray.

  • De la Beche, Henry T. 1839. Report on the geology of Cornwall, Devon, and West Somerset. London: Longman, Orme, Brown, Green and Longmans.

    Google Scholar 

  • Dewey, John F., and Walter C. Pitman. 1998. Sea-level changes: Mechanisms, magnitudes and rates. In Paleogeographic evolution and non-glacial Eustasy, northern South America, vol. 58, ed. James L. Pindell and Charles L. Drake, 1–16. Society for Sedimentary Geology Special Publication.

  • Dickinson, William R. 2003. The place and power of myth in geoscience: An associate editor’s perspective. American Journal of Science 303: 856–864.

    Google Scholar 

  • Dott, Robert H., Jr. 1978. Tectonics and sedimentation a century later. Earth-Science Reviews 14: 1–34.

    Google Scholar 

  • Doyle, Peter, and Matthew R. Bennett, eds. 1998. Unlocking the stratigraphical record. Hoboken: Wiley.

    Google Scholar 

  • Dresow, Max. 2017. Before hierarchy: The rise and fall of Stephen Jay Gould’s first macroevolutionary synthesis. History and Philosophy of the Life Sciences 39: 6. https://doi.org/10.1007/s40656-017-0133-6.

    Article  Google Scholar 

  • Dresow, Max. 2019. Macroevolution evolving: Punctuated equilibria and the roots of Stephen Jay Gould’s second macroevolutionary synthesis. Studies in History and Philosophy of Biological and Biomedical Sciences 75: 15–23.

    Google Scholar 

  • Dresow, Max. 2021. Measuring time with fossils: A start-up problem in scientific practice. Philosophy of Science. https://doi.org/10.1086/714855.

    Article  Google Scholar 

  • Droser, Mary. 1995. Paleobiology goes into the field. PALIOS 10: 507–516.

    Google Scholar 

  • Edwards, Paul N. 2010. A vast machine: Computer models, climate data, and the politics of global warming. Cambridge: MIT Press.

    Google Scholar 

  • Eldredge, Niles, and Stephen Jay Gould. 1972. Punctuated equilibria: An alternative to phyletic gradualism. In Models in paleobiology, ed. Thomas J.M. Schopf, 82–115. San Francisco: Freeman, Cooper and Company.

    Google Scholar 

  • Embry, Ashton, Erik Johannessen, Donald Owen, Benoit Beauchamp, and Piero Gianolla. 2007. Sequence stratigraphy as a “concrete” stratigraphic discipline. Report of the ISSC Task Group on Sequence Stratigraphy.

  • Folk, Robert L., and John C. Ferm. 1966. A portrait of Paul D. Krynine. Journal of Sedimentary Petrology 36: 853–863.

    Google Scholar 

  • Foote, Mike. 1999. Evolutionary patterns in the fossil record. Evolution 50: 1–11.

    Google Scholar 

  • Fortey, Richard A. 1997. Life: An unauthorized biography. New York: Harper Collins Publishers LLC.

    Google Scholar 

  • Gould, Stephen Jay. 1969. An evolutionary microcosm: Pleistocene and recent history of the land snail P (poecilozonites) in Bermuda. Bulletin of the Museum of Comparative Zoology 138: 407–532.

    Google Scholar 

  • Gould, Stephen Jay. 1980. The promise of paleobiology as a nomothetic, evolutionary discipline. Paleobiology 6: 96–118.

    Google Scholar 

  • Gould, Stephen Jay. 1995. A task for paleobiology at the threshold of majority. Paleobiology 21: 1–14.

    Google Scholar 

  • Gould, Stephen Jay, and Niles Eldredge. 1977. Punctuated equilibria: The tempo and mode of evolution reconsidered. Paleobiology 3: 115–151.

    Google Scholar 

  • Grantham, Todd A. 2009. Taxic paleobiology and the pursuit of a unified evolutionary theory. In The paleobiological revolution: Essays on the growth of modern paleobiology, ed. David Sepkoski and Michael Ruse, 215–234. Chicago: University of Chicago Press.

    Google Scholar 

  • Greene, Mott. 2009. Geology. In The Cambridge history of science. Volume 6. The modern biological and earth sciences, ed. Peter J. Bowler and J.V. Pickstone, 167–184. Cambridge: Cambridge University Press.

    Google Scholar 

  • Hallam, Anthony. 1978. Eustatic cycles in the jurassic. Palaeogeography Palaeoclimatology and Palaeoecology 23: 1–32.

    Google Scholar 

  • Hallam, Anthony. 1988. A re-evaluation of Jurassic Eustasy in the light of new data and the revised Exxon curve. In Sea-level changes: An integrated approach, vol. 42, ed. Cheryl K. Wilgus, Bruce S. Hastings, Henry W. Posamentier, John C. Van Wagoner, Charles A. Ross, and Christopher G. St Kendall, 261–273. Society of Economic Paleontologists and Mineralogists Special Publication.

  • Hedberg, Hollis D. 1965. Chronostratigraphy and biostratigraphy. Geological Magazine 102: 451–461.

    Google Scholar 

  • Holland, Steven M. 1995. The stratigraphic distribution of fossils. Paleobiology 21: 92–109.

    Google Scholar 

  • Holland, Steven M. 1999. The new stratigraphy and its promise for paleobiology. Paleobiology 25: 409–416.

    Google Scholar 

  • Holland, Steven M. 2000. The quality of the fossil record: A sequence stratigraphic perspective. Paleobiology 26: 148–168.

    Google Scholar 

  • Holland, Steven M. 2017. Structure, not bias. Paleontology 91: 1315–1317.

    Google Scholar 

  • Jablonski, David. 1980. Apparent versus real biotic effects of transgressions and regressions. Paleobiology 6: 397–407.

    Google Scholar 

  • Jervey, Macomb T. 1988. Quantitative geological modeling of Siliciclastic rock sequences and their seismic expression. In Sea-level changes—An integrated approach, vol. 42, ed. C.K. Wilgus, B.S. Hastings, Henry W. Posamentier, John C. Van Wagoner, C.A. Ross, and Christopher G. St Kendall, 47–69. Society of Economic Paleontologists and Mineralogists (SEPM) Special Publications.

  • Johnson, Kristin. 2007. Natural history as stamp-collecting: A brief history. Archives of Natural History 34: 244–254.

    Google Scholar 

  • Kauffman, Erle G. 1984. Paleobiogeography and evolutionary response dynamic in the cretaceous western Interior Seaway of North America. In Jurassic-cretaceous biochronology and paleogeography of North America, vol. 27, ed. G.E.G. Westermann, 273–306. St. John’s: Geological Association of Canada Special Publication.

  • Kauffman, Erle G. 1988. Concepts and methods of high-resolution event stratigraphy. Annual Review of Earth and Planetary Sciences 16: 605–654.

    Google Scholar 

  • Kauffman, Erle G., and Bradley B. Sageman. 1992. Biological patterns in sequence stratigraphy; cretaceous of the western Interior Basin, North America. In Fifth North America Paleontological Convention, abstracts and program, ed. Scott Lidgard and Peter R. Crane. Knoxville: The Paleontological Society.

    Google Scholar 

  • Kay, Marshall. 1974. Geosynclines, flysch and melanges. In Modern and ancient geosynclinal sedimentation, vol. 19, eds. R.H. Dott Jr., and R.H. Shaver, 377–380. Tulsa: Society of Economic Paleontologists and Mineralogists, Special Publication.

  • Kelley, Patricia H., David E. Fastovsky, Mark A. Wilson, Richard A. Laws, and Anne Raymond. 2013. From paleontology to paleobiology: A half-century of progress in understanding life. In The web of geological sciences: advances, impacts, interactions, vol. 500, ed. Marion E. Bickford, 191–232. The Geological Society of America, Special Paper.

  • Kerr, Richard A. 1980. Changing global sea levels as a geologic index. Science 209: 483–486.

    Google Scholar 

  • Kidwell, Susan M. 1986. Models for fossil concentration: Paleobiological implications. Paleobiology 12: 6–24.

    Google Scholar 

  • Kidwell, Susan M. 1989. Stratigraphic condensation of marine transgressive records: Origin of major shell deposits in the Miocene of Maryland. Journal of Geology 97: 1–24.

    Google Scholar 

  • Kidwell, Susan M., and Steven M. Holland. 2002. The quality of the fossil record: Implications for evolutionary analysis. Annual Review of Ecology and Systematics 33: 561–588.

    Google Scholar 

  • Kidwell, Susan M., and David Jablonski. 1983. Taphonomic feedback: Ecological consequences of shell accumulation. In Biotic interactions in recent and fossil benthic communities, ed. Michael J.S. Tevesz, and Peter L. McCall, 195–248. New York: Springer.

    Google Scholar 

  • Knight, J., and Brookes. 1947. Paleontologist or geologist. Bulletin of the Geological Society of America 58: 281–286.

    Google Scholar 

  • Kohler, Robert E. 2002. Landscapes and labscapes: Exploring the lab-field border in biology. Chicago: University of Chicago Press.

    Google Scholar 

  • MacLeod, Norman. 1991. Punctuated anagenesis and the importance of stratigraphy to paleobiology. Paleobiology 17: 167–188.

    Google Scholar 

  • Miall, Andrew D. 1992. Exxon global cycle chart: An event for every occasion? Geology 20: 787–790.

    Google Scholar 

  • Miall, Andrew D. 2010. The geology of stratigraphic sequences. New York: Springer.

    Google Scholar 

  • Miall, Andrew D. 2015. Making stratigraphy respectable: From stamp collecting to astronomical calibration. Geoscience Canada 42: 271–302.

    Google Scholar 

  • Miall, Andrew D., and Charlene E. Miall. 2001. Sequence stratigraphy as a scientific enterprise: The evolution and persistence of conflicting paradigms. Earth-Science Reviews 54: 321–348.

    Google Scholar 

  • Miall, Charlene E., and Andrew D. Miall. 2002. The Exxon factor: The roles of corporate and academic science in the emergence and legitimation of a new global model of sequence stratigraphy. The Sociological Quarterly 43: 307–334.

    Google Scholar 

  • Mitchum, Robert M. 2003. Penrose Medal Citation. Presented to Peter R. Vail. GSA Medals & Awards, Geological Society of America. https://www.geosociety.org/awards/03speeches/penrose.htm.

  • Mitchum, Robert M., Jr., and John C. Van Wagoner. 1991. High-frequency sequences and their stacking patterns: Sequence-stratigraphic evidence of high-frequency Eustatic cycles. Sedimentary Geology 70: 131–160.

    Google Scholar 

  • Mitchum, Robert M., Peter R. Vail, and Samuel Thompson III. 1977. Seismic stratigraphy and global changes of sea level, part 2: The depositional sequence as a basic unit for stratigraphic analysis. In Seismic stratigraphy—Applications to hydrocarbon exploration, ed. C.E. Payton, 53–62. Tulsa: American Association of Petroleum Geologists Memoir 26.

    Google Scholar 

  • Newell, Norman D. 1962. Paleontological gaps and geochronology. Journal of Paleontology 36: 592–610.

    Google Scholar 

  • Oldroyd, David. 2006. Earth cycles: A historical perspective. Westport, CT: Greenwood Press.

    Google Scholar 

  • Olsen, Everett C. 1980. Taphonomy: Its history and role in community evolution. In Fossils in the making, ed. Anna K. Behrensmeyer and Andrew P. Hill, 6–19. Chicago: University of Chicago Press.

    Google Scholar 

  • Patzkowsky, Mark E., and Steven M. Holland. 2012. Stratigraphic paleobiology: Understanding the distribution of fossil taxa in space and time. Chicago: University of Chicago Press.

    Google Scholar 

  • Posamentier, Henry W., and Peter R. Vail. 1988. Eustatic controls on clastic deposition II—Sequence and systems tract models. In Sea-level changes—An integrated approach, vol. 42, ed. Cheryl K. Wilgus, Bruce S. Hastings, Henry W. Posamentier, John C. Van Wagoner, Charles A. Ross, and Christopher G. St Kendall, 125–154. Society of Economic Paleontologists and Mineralogists (SEPM) Special Publications.

  • Poulsen, Christopher J., Peter B. Flemings, R.A.J. Robinson, and John M. Metzger. 1998. Three-dimensional stratigraphic evolution of the Miocene Baltimore Canyon region: Implications for Eustatic interpretations and the systems tract model. Geological Society of America Bulletin 110: 1105–1122.

    Google Scholar 

  • Raup, David M., and John J. Sepkoski Jr. 1982. Mass extinctions in the marine fossil record. Science 215: 1501–1503.

    Google Scholar 

  • Rieppel, Olivier. 2013. Othenio Abel (1875–1946) and the rise and decline of paleobiology in German paleontology. Historical Biology 25: 1–13.

    Google Scholar 

  • Rudwick, Martin J.S. 1968. Some analytical methods in the study of ontogeny in fossils with accretionary skeletons. Paleontological Society Memoir 2: 35–49.

    Google Scholar 

  • Rudwick, Martin J.S. 1972. The meaning of fossils: Episodes in the history of paleontology. Chicago: University of Chicago Press.

    Google Scholar 

  • Rudwick, Martin J.S. 1985. The great devonian controversy: The shaping of scientific knowledge among gentlemanly specialists. Chicago: University of Chicago Press.

    Google Scholar 

  • Rudwick, Martin J.S. 2008. Worlds before Adam: The reconstruction of geohistory in the age of reform. Chicago: University of Chicago Press.

    Google Scholar 

  • Rudwick, Martin J.S. 2017. Functional morphology in paleobiology: Origins of the method of ‘paradigms.’ Journal of the History of Biology 50: 1–44.

    Google Scholar 

  • Rudwick, Martin J.S. 2018. The fate of the method of ‘paradigms’ in paleobiology. Journal of the History of Biology 51: 479–533.

    Google Scholar 

  • Ruse, Michael. 2009. Punctuations and paradigms: Has paleobiology been through a paradigm shift? In The paleobiological revolution: Essays on the growth of modern paleobiology, ed. David Sepkoski and Michael Ruse, 518–527. Chicago: University of Chicago Press.

    Google Scholar 

  • Sadler, Peter M. 1981. Sediment accumulation rates and the completeness of stratigraphic sections. The Journal of Geology 89: 569–584.

    Google Scholar 

  • Seibold, Eugen, and Ilse Seibold. 2002. Sedimentology: From single grain to recent and past environments: Some trends in sedimentology in the twentieth century. In The earth inside and out: Some major contributions to geology in the twentieth century, ed. D. Oldroyd, 241–250. The Geological Society of London.

  • Sepkoski, John J., Jr. 1978. Taphonomic factors influencing the lithologic occurrence of fossils in Dresbachian (Upper Cambrian) shaly facies. Geological Society of America Abstracts with Programs 10: 490.

    Google Scholar 

  • Sepkoski, John J., Jr. 1993. Ten years in the library: New data confirm paleontological patterns. Paleobiology 19: 43–51.

    Google Scholar 

  • Sepkoski, David. 2005. Stephen Jay Gould, Jack Sepkoski, and the “quantitative revolution” in American paleobiology. Journal of the History of Biology 38: 209–237.

    Google Scholar 

  • Sepkoski, David. 2012. Rereading the fossil record: The growth of paleobiology as an evolutionary discipline. Chicago: University of Chicago Press.

    Google Scholar 

  • Sepkoski, David. 2013. Towards ‘a natural history of data’: Evolving practices and epistemologies of data in paleontology, 1800–2000. Journal of the History of Biology 46: 401–444.

    Google Scholar 

  • Sepkoski, David. 2017. The earth as archive: Contingency, narrative and the history of life. In Science in the archives, ed. Lorraine Daston, 53–84. Chicago: University of Chicago Press.

    Google Scholar 

  • Sepkoski, David. 2019. The unfinished synthesis? Paleontology and evolutionary biology in the 21st century. Journal of the History of Biology 52: 687–703.

    Google Scholar 

  • Sepkoski, David. 2021. Catastrophic thinking: Extinction and the value of diversity from Darwin to the Anthropocene. Chicago: University of Chicago Press.

    Google Scholar 

  • Sepkoski, David, and Marco Tamborini. 2018. An image of science: Cameralism, statistics and the visual language of natural history in the nineteenth century. Historical Studies of the Natural Sciences 48: 56–109.

    Google Scholar 

  • Sepkoski, John J., Jr. 1984. A factor analytic description of the Phanerozoic marine fossil record. Paleobiology 7: 36–53.

    Google Scholar 

  • Shaw, Alan B. 1964. Time in stratigraphy. New York: McGraw-Hill.

    Google Scholar 

  • Sloss, Laurence L. 1963. Sequences in the cratonic interior of North America. Geological Society of America Bulletin 74: 93–114.

    Google Scholar 

  • Sloss, Laurence L. 1988. Forty years of sequence stratigraphy. Geological Society of America Bulletin 100: 1661–1665.

    Google Scholar 

  • Sloss, Laurence L., William C. Krumbein, and Edward C. Dapples. 1949. Integrated facies analysis. In Sedimentary facies in geologic history, ed. Charles R. Longwell, 91–124. Geological Society of America Memoir 39.

  • Steel, Ronald J., and Kitty L. Milliken. 2013. Major advances in siliciclastic sedimentary geology, 1960–2012. In The web of the geological sciences: Advances, impacts, interactions, vol. 500, ed. Marion E. Bickford, 121–168. Geological Society of America Special Paper.

  • Tamborini, Marco. 2019. Technoscientific approaches to deep time. Studies in History and Philosophy of Science 79: 57–67.

    Google Scholar 

  • Tamborini, Marco. 2022. A plea for a new synthesis: From twentieth-century paleobiology to twenty-first-century paleontology and back again. Biology 11: 1120. https://doi.org/10.3390/biology11081120.

    Article  Google Scholar 

  • Teichert, Curt. 1958. The concepts of facies. AAPG Bulletin 42: 2718–2744.

    Google Scholar 

  • Turner, Derek. 2009. Beyond detective work: Empirical testing in paleobiology. In The paleobiological revolution: Essays on the growth of modern paleobiology, ed. David Sepkoski and Michael Ruse, 201–214. Chicago: University of Chicago Press.

    Google Scholar 

  • Turner, Derek. 2011. Paleontology: A philosophical introduction. Cambridge: Cambridge University Press.

    Google Scholar 

  • Vail, Peter R. 1992. The evolution of seismic stratigraphy and the global sea-level curve. In Eustasy: The historical ups and downs of a major geological concept, ed. Robert H. Dott Jr., 83–92. Boulder: The Geological Society of America.

    Google Scholar 

  • Vail, Peter R., R.G. Todd, and J.B. Sangree. 1977a. Seismic stratigraphy and global changes of sea-level, part 5: Chronostratigraphic significance of seismic reflections. In Seismic stratigraphy—Applications to hydrocarbon exploration, vol. 26, ed. C.E. Payton, 99–116. Tulsa: American Association of Petroleum Geologists Memoir.

    Google Scholar 

  • Vail, Peter R., Robert M. Mitchum, and Samuel Thompson III. 1977b. Seismic stratigraphy and global changes of sea-level, part 3: Relative changes in sea level from coastal onlap. In Seismic stratigraphy—Applications to hydrocarbon exploration, vol. 26, ed. C.E. Payton, 63–81. Tulsa: American Association of Petroleum Geologists Memoir.

    Google Scholar 

  • Vail, Peter R., Robert M. Mitchum, and Samuel Thompson III. 1977c. Seismic stratigraphy and global changes of sea-level, part 4: Global cycles of relative change of sea level. In Seismic stratigraphy—Applications to hydrocarbon exploration, vol. 26, ed. C.E. Payton, 83–97. Tulsa: American Association of Petroleum Geologists Memoir.

    Google Scholar 

  • Valentine, James W. 2009. The infusion of biology into paleontological research. In The paleobiological revolution: Essays on the growth of modern paleobiology, ed. David Sepkoski and Michael Ruse, 385–397. Chicago: University of Chicago Press.

    Google Scholar 

  • Van Wagoner, John C., Henry W. Posamentier, Robert M. Mitchum, Peter R. Vail, J. Fredrick Sarg, T.S. Loutit, and Jan Hardenbol. 1988. An Overview of the fundamentals of sequence stratigraphy and key definitions. In Sea-level changes—An integrated approach, vol. 42, ed. Cheryl K. Wilgus, Bruce S. Hastings, Henry W. Posamentier, John C. Van Wagoner, Charles A. Ross, and Christopher G. St Kendall, 39–45. Society of Economic Paleontologists and Mineralogists (SEPM) Special Publications.

  • Van Wagoner, John C., Robert M. Mitchum, Kirt M. Campion, and Victor D. Rahmanian. 1990. Siliciclastic sequence stratigraphy in well logs, cores, and outcrops, vol. 7. Tulsa: American Association of Petroleum Geologists.

  • Voorhies, Michael. 1969. Taphonomy and population dynamics of an early pliocene vertebrate fauna, Knox County, Nebraska. Contributed Geological Society of America Special Paper No. 1. Laramie: University of Wyoming.

  • Ward, Peter. 1983. The extinction of the ammonites. Scientific American 249: 136–147.

    Google Scholar 

  • Ward, Peter. 1994. The end of evolution: On mass extinctions and the preservation of biodiversity. New York: Bantam Books.

    Google Scholar 

Download references

Acknowledgements

I would like to thank David Sepkoski, Steven Holland, Scott Lidgard, Douglas Erwin, Marco Tamborini, Alan Love, Bennett McNulty, and Jos Uffink for comments on earlier drafts of this paper. I would also like to thank Jean-Baptiste Grodwohl and an anonymous reviewer for comments and criticisms that greatly improved the manuscript, and Betty Smocovitis for her editorial insights.

Funding

No funding was received to assist with the preparation of this manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Max Dresow.

Ethics declarations

Competing Interests

The author is aware of no potential conflicts of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dresow, M. Biased, Spasmodic, and Ridiculously Incomplete: Sequence Stratigraphy and the Emergence of a New Approach to Stratigraphic Complexity in Paleobiology, 1973–1995. J Hist Biol 56, 419–454 (2023). https://doi.org/10.1007/s10739-023-09720-0

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10739-023-09720-0

Keywords

Navigation