National Academies Press: OpenBook

Effects of Past Global Change on Life (1995)

Chapter: ESTABLISHING A CHRONOLOGY FOR ENVIRONMENTAL DECLINE AND MASS EXTINCTION ACROSS THE C-T BOUNDARY

« Previous: A CASE HISTORY: THE PUEBLO, COLORADO, C-T BOUNDARY SECTION
Suggested Citation:"ESTABLISHING A CHRONOLOGY FOR ENVIRONMENTAL DECLINE AND MASS EXTINCTION ACROSS THE C-T BOUNDARY." National Research Council. 1995. Effects of Past Global Change on Life. Washington, DC: The National Academies Press. doi: 10.17226/4762.
×
Page 59
Suggested Citation:"ESTABLISHING A CHRONOLOGY FOR ENVIRONMENTAL DECLINE AND MASS EXTINCTION ACROSS THE C-T BOUNDARY." National Research Council. 1995. Effects of Past Global Change on Life. Washington, DC: The National Academies Press. doi: 10.17226/4762.
×
Page 60

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

GLOBAL CHANGE LEADING TO BIODIVERSITY CRISIS IN A GREENHOUSE WORLD: THE CENOMANIAN-TURONIAN 59 (CRETACEOUS) MASS EXTINCTION possibly warmer climate phases, more normal marine surface conditions, and a dominance of biogenic, hemipelagic sedimentation with great reduction in clay delivery to the basin. The Milankovitch climate cycles represent a predictable aspect of Cretaceous global change, but are especially enhanced in lithological, geochemical, and biological expression during the middle Cretaceous (Barron et al., 1985; Kauffman, 1988a). Both planktic and benthic biotas associated with the Cenomanian-Turonian boundary interval at Pueblo represent warm, normal to near-normal marine environments (excepting intervals of benthic oxygen restriction). These biotas paleobiogeographically belong to the warm-temperate to subtropical Southern Interior Subprovince and its northern biogeographic ecotone with the Central Interior Subprovince—the principal endemic center for molluscan evolution in the Western Interior Basin (Kauffman, 1984b). Against this dynamic environmental backdrop of global sea-level fluctuations, frequent changes in stratification, temperature and water chemistry, active regional tectonism, basin subsidence and volcanism, and more predictable changes in climate related to Milankovitch cyclicity, the global Cenomanian biota became progressively more stressed and underwent accelerated extinction as a series of ecologically graded (tropical to temperate) steps or events during the last million years of the Cenomanian (Kauffman, 1984a, 1988b; Elder, 1985, 1987a,b; Johnson and Kauffman, 1990; Figures 3.4 and 3.6 herein). The final 520,000 yr of this extinction interval are especially well preserved in temperate facies of the Western Interior Cretaceous Basin. Each extinction event was tied to a major short-term shift in ocean temperature and/or chemistry, indicated by the geochemical profiles at Pueblo (Figures 3.3 and 3.6). Each event represents a breached biological threshold for a particular suite of ecosystems, starting with tropical Caribbean Province reef-associated communities (Johnson and Kauffman, 1990). The high-resolution data base on global environmental change, as exemplified by the Pueblo, Colorado section, provides important insights into the rates, patterns, causes, and consequences of the global C-T biotic crisis at scales comparable to those used for long-term sensing of Quaternary global change. This, in turn, allows critical comparisons between Quaternary icehouse and Cretaceous greenhouse worlds. ESTABLISHING A CHRONOLOGY FOR ENVIRONMENTAL DECLINE AND MASS EXTINCTION ACROSS THE C-T BOUNDARY To compare the rates, patterns, and magnitude of environmental changes and biological stress leading to global late Cenomanian-early Turonian mass extinction, with the Quaternary record of global change, a highly detailed real-time scale must be established for the C-T interval. Integration of this time scale with physical, chemical, and biological event surfaces/ beds, and with short-term cyclostratigraphic units (e.g., Milankovitch climate cycle deposits) of regional extent, has resulted in the construction of a high-resolution chronostratigraphic system (Kauffman, 1988a; Kauffman et al., 1991), which allows precise regional correlation and interpretation, individually or collectively, of C-T boundary mass extinction events. In the C- T boundary chronology developed for the Western Interior Basin of North America, biostratigraphic zonation has been resolved to 0.18-0.33 m.y. per biozone; Milankovitch cyclostratigraphy to 41,000 yr per cycle, and event chronostratigraphic resolution to 17,000 yr per event-bounded interval. There are few systems of chronology and correlation that match this level of resolution, which has similarly been developed in western Europe and northern South America. Figures 3.3 and 3.4 show the physical, chemical, biological, and cyclostratigraphic details of the late Cenomanian-early Turonian section at the Pueblo, Colorado standard reference section and proposed boundary stratotype (Kennedy and Cobban, 1991). The Bridge Creek Limestone spans the C-T boundary (Figure 3.5) and contains 24 bentonite beds and 13 additional persistent limonite beds, probably representing altered bentonites. These bentonites record the late phases of one of the most active intervals of volcanism in Cretaceous history (Kauffman, 1985, Figure 4). Thirteen of these bentonites (PBC 4, 5, 11, 17, 19, 20, 30, 32, 44-46, 48, and 50, in Kauffman et al., 1985, p. FRS-11) are much thicker and more highly persistent than the others, and form regional chronostratigraphic marker beds (Hattin, 1971, 1985; Elder, 1985; Kauffman, 1988a). These have yielded materials for radiometric dating. Utilizing single sanidine crystal 40Ar-39Ar dating techniques, Obradovich (1993) has established high-resolution geochronologic tie points to ammonite biozones of Cobban (1985, 1993; Figure 3.4 herein) that allow a calculated time scale to be constructed for the entire interval, if equal time durations are assumed for all intervening ammonite biozones between dated Cenomanian-Turonian bentonite levels (Figure 3.5, second column). This commonly used calibration technique assumes constant evolutionary rates within ammonite lineages, however, which is improbable. Obradovich (1993) has dated the lowest ammonite zone in the Bridge Creek Limestone Member (Vascoceras diartanum), bed 63, as having a mean value of 93.9 Ma, from a bentonite that does not occur in the condensed section at Pueblo (Figure 3.5). A second age of 93.56 Ma has been obtained from the latest Cenomanian Neocardioceras juddi zone in a bed equivalent to the ''Neocardioceras bentonite" of Elder and Kirkland (1985) and Elder (1985, 1987a), unit 80, PBC-11 of the Pueblo

GLOBAL CHANGE LEADING TO BIODIVERSITY CRISIS IN A GREENHOUSE WORLD: THE CENOMANIAN-TURONIAN 60 (CRETACEOUS) MASS EXTINCTION reference section (Kauffman et al., 1985, p. FRS-11). A third mean age of 93.25 Ma was obtained from the early Turonian Pseudaspidoceras flexuosum zone in a bed equivalent to bentonite marker bed 96 (PBC 20; Kauffman et al., 1985, p. FRS-11). The top of the Bridge Creek Limestone Member (lowest part of the Collignoniceras woollgari biozone) has been given an interpolated age of 92.05 Ma based on assigning equal age ranges to ammonite biozones that lie between well-dated levels below and above this biozone (see Figure 3.5). Based on this set of new 40Ar-39Ar ages, and the calculated time scale that can be constructed between them, if equal age ranges are assumed for intervening ammonite biozones, a duration of 1.95 m.y. is calculated for the Bridge Creek Limestone Member at Pueblo. Here, the member spans the entire set of C-T mass extinction steps (Kauffman, 1988b; Figures 3.4 and 3.6 herein), the complete survival interval, and the recovery interval (sensu Harries and Kauffman, 1990) up to basic recovery of the marine ecosystem in the early to early-middle Turonian. The average durations for ammonite subzones and zones, based on the exclusive ranges of biostratigraphically tested ammonite species, equals 190,000 yr for the C-T interval. This is calculated by dividing the number of ammonite biozones into the time duration between the mean values of bounding radiometric ages. Because of the dangers in the assumption that ammonite species evolve at relatively constant rates, even through this thin interval, an independent, more highly refined dating system was constructed to test this time scale by using shale (or marl)-limestone (or chalk, calcarenite) bedding rhythms reflecting Milankovitch climate cycles. The scale of this independent chronology is adequate to evaluate the relative importance and interaction of individual physical, chemical, and biological events associated with the C-T mass extinction, survival, and recovery intervals. At Pueblo, various authors have measured closely spaced sections of the Bridge Creek Limestone Member in great detail, and have reported between 41 and 50 shale (or marl)-limestone (or chalk, calcarenite) bedding couplets. Most sections yield between 44 and 48 bedding couplets that probably reflect Milankovitch cyclicity (Barron et al., 1985; Fischer et al., 1985; Kauffman, 1988a) in the form of climate-regulated dilution cycles (Pratt, 1985) and/or productivity cycles (Eicher and Diner, 1985, 1989). At the proposed C-T boundary stratotype at Pueblo, Elder and Kirkland (1985) and subsequent work by the author have confirmed 46 Milankovitch-type bedding cycles of all scales in the Bridge Creek Limestone Member, 19 of which are capped by relatively thicker, more resistant limestone or calcarenite beds with extensive basinal dispersion (Figure 3.5; Hattin, 1971, 1985; Elder and Kirkland, 1985; Kauffman, 1988a). These thicker and more pervasive limestone units probably reflect the 100,000-yr Milankovitch orbital eccentricity cycle (Barron et al., 1985; Fischer et al., 1985; Kauffman et al., 1987; Kauffman, 1988a) (Figure 3.5). There are an average of 2.5 smaller bedding cycles between these more prominent limestone beds, suggesting that they may represent the 41,000-yr Milankovitch axial eccentricity cycle (Figure 3.5). These interpretations are subsequently tested and confirmed. Given the 1.95-m.y. calculated duration for the Bridge Creek Limestone, C-T boundary section at Pueblo, the 19 thicker, more regionally persistent limestone beds have an average duration of 100,000 yr, and logically represent the Milankovitch orbital eccentricity cycle (Figure 3.5). Further, the average duration for the 46 bedding cycles of all scales is 41,300 yr; these are considered representative of the 41,000-yr Milankovitch axial eccentricity cycles (Figure 3.5). This is the optimal level of age resolution for interpreting regional environmental change and the patterns of C-T mass extinction. These Milankovitch cycle determinations provide an independent means of testing, and recalibrating, the interpolated time scale, based on assigning equal durations to ammonite biozones between radiometrically dated levels. Two methods of calibrating a new C-T boundary time scale, based on integrating radiometric ages with 100,000-yr Milankovitch climate cycles, are presented in Figure 3.5, columns A and B. The first, and most commonly practiced method, starts at a calibrated or radiometric age at the top (or base) of the long Milankovitch cycle sequence, and assigns progressively greater (or lesser) 100,000-yr intervals to the top of each bedding rhythm representing the orbital eccentricity cycle (e.g., Figure 3.5, column B). When this method is applied to the Bridge Creek Limestone Member at Pueblo, the calculated values for the 100,000-yr Milankovitch bedding rhythms deviate from the mean values of 40Ar-39Ar ages within the member by 100,000 to 160,000 yr, and vary from the time scale calculated on the assignment of equal durations to successive ammonite biozones by 150,000 to 360,000 yr. A second, and preferred, calculated time scale was constructed by utilizing a mean 40Ar-39Ar age from the middle of the Bridge Creek Limestone Member as a starting point (93.25 Ma for the bentonite in the middle of the early Turonian Pseudaspidocerasflexuosum biozone; Figure 3.5, column A), and assigning progressively younger 100,000-yr durations for orbital eccentricity cycles up-section, and progressively older 100,000-yr durations for the same cycles down-section. In this method, the ages calculated from the Milankovitch cycle scale deviate by only 10,000 to 50,000 yr from other 40Ar-39Ar ages within the sequence, and only 10,000 to 50,000 yr from ages calculated by assigning equal range durations to late Cenomanian ammonite biozones, which suggests a relatively constant evo

Next: INTERPRETATIONS AND CONCLUSIONS »
Effects of Past Global Change on Life Get This Book
×
Buy Hardback | $65.00 Buy Ebook | $49.99
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

What can we expect as global change progresses? Will there be thresholds that trigger sudden shifts in environmental conditions—or that cause catastrophic destruction of life?

Effects of Past Global Change on Life explores what earth scientists are learning about the impact of large-scale environmental changes on ancient life—and how these findings may help us resolve today's environmental controversies.

Leading authorities discuss historical climate trends and what can be learned from the mass extinctions and other critical periods about the rise and fall of plant and animal species in response to global change. The volume develops a picture of how environmental change has closed some evolutionary doors while opening others—including profound effects on the early members of the human family.

An expert panel offers specific recommendations on expanding research and improving investigative tools—and targets historical periods and geological and biological patterns with the most promise of shedding light on future developments.

This readable and informative book will be of special interest to professionals in the earth sciences and the environmental community as well as concerned policymakers.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!