Hostname: page-component-7c8c6479df-fqc5m Total loading time: 0 Render date: 2024-03-27T08:03:40.851Z Has data issue: false hasContentIssue false

The biotic effects of large bolide impacts: size versus time and place

Published online by Cambridge University Press:  01 October 2008

Gordon Walkden
Affiliation:
University of Aberdeen, Geology and Petroleum Geology, Kings College, Aberdeen, AB24 3UE, UK
Julian Parker
Affiliation:
University of Aberdeen, Geology and Petroleum Geology, Kings College, Aberdeen, AB24 3UE, UK

Abstract

In estimating the biotic effects of large terrestrial impacts we are reliant upon apparent crater diameter as a proxy for impact magnitude. This underlies the ‘kill-curve’ approach which graphs crater diameter directly against likely percentage losses of taxa. However, crater diameter is a complex product of syn- and post-impact processes that can be site-dependent. Furthermore, location (global positioning) and timing (moment in geological history) also strongly influence biotic effects. We examine four of our largest and best-documented Phanerozoic impacts to explore this more holistic size–time–place relationship. Only the c. 180 km end-Cretaceous Chicxulub crater (Mexico) links to any substantial immediate extinction and some of the worst effects stem from where it struck the planet (a continental margin carbonate platform site) and when (a time of high regional and global biodiversity). Both the c. 100 km late Triassic Manicouagan crater in NE Canada (arid continental interior, low regional and world biodiversity) and the c. 35 Ma 100 km Popigai crater, Siberia (continental arctic desert) provide much less damaging scenarios. However the c. 90 km Chesapeake Bay crater, Eastern USA (also c. 35 Ma) marks a far more sensitive (Chicxulub-like) site but it also proved relatively benign. Here the rheologically varied shallow marine target site produced an anomalously broad crater, and the scale of the impact has evidently been overestimated. We offer a new approach to the graphical prediction of biotic risk in which both crater diameter and a generalised time/place factor we term ‘vulnerability’ are variables.

Type
Research Article
Copyright
Copyright © 2008 Cambridge University Press

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

Belcher, C.M., Collinson, M.E. & Scott, A. (2005). Constraints on the thermal energy released from the Chicxulub impactor: new evidence from multi-method charcoal analysis. J. Geol. Soc. Lond. 162, 591602.CrossRefGoogle Scholar
Benton, M.J. (2001). Biodiversity through time. In Palaeobiology II, ed. Briggs, D.E.G. & Crowther, P.R., pp. 211220. Blackwell Science, Oxford.CrossRefGoogle Scholar
Bottomley, R.J., Grieve, R., York, D. & Masaitis, V. (1997). The age of the Popigai impact event and its relation to events at the Eocene/Oligocene boundary. Nature 388, 365368.CrossRefGoogle Scholar
Claeys, P., Kiessling, W. & Alvarez, W. (2002). Distribution of Chicxulub Ejecta at the Cretaceous-Tertiary boundary. In Catastrophic events and Mass Extinctions, ed. Koeberl, C. Special Paper 356, p. 55. Geological Society of America.Google Scholar
Collins, G.S. & Melosh, H.J. (2004). Hydrocode Simulations of the Chesapeake Bay Impact. In Lunar and Planetary Science XXXV, Abstract no. 1005. Lunar and Planetary Institute, Houston, TX.Google Scholar
Collins, G.S. & Wunnemann, K. (2005). How big was the Chesapeake Bay impact? Insight from numerical modelling. Geology 33, 925928.CrossRefGoogle Scholar
Collins, G.S., Melosh, H.J., Morgan, J.V. & Warner, M.R. (2002). Hydrocode Simulations of Chicxulub Crater Collapse and Peak-Ring Formation. Icarus 157(8), 2433.CrossRefGoogle Scholar
Deutsch, A. & Koeberl, C. (2006). Establishing the link between the Chesapeake Bay impact structure and the North American tektite strewn field: the Sr-Nd isotopic evidence. Meteorit. Planet. Sci. 41, 689703.CrossRefGoogle Scholar
Durda, D.D. & Kring, D.A. (2004). Ignition threshold for impact-generated fires. J. Geophys. Res. 109(E8), 114.Google Scholar
Golonka, J. (2002). Plate-tectonic maps of the Phanerozoic. In Phanerozoic Reef Patterns, ed. Kiessling, W., Flügel, E. & Golonka, J. SEPM Special Publication Series No.72. Society for Sedimentary Geology, Tulsa.Google Scholar
Grieve, R.A.F. & Head, J.W. (1983). The Manicouagan impact structure – an analysis of its original dimensions. J. Geophys. Res. 88, A807A818.Google Scholar
Gupta, S.C., Ahrens, T.J. & Yang, W. (2001). Shock-induced vaporization of anhydrite and global cooling from the K/T impact. Earth Planet. Sci. Lett. 188, 399412.CrossRefGoogle Scholar
Hallam, A. (2002). How catastrophic was the end-Triassic mass extinction? Lethaia 35, 147157.Google Scholar
Hildebrand, A.R. et al. (1998a). Mapping Chicxulub crater structure with overlapping gravity and seismic surveys. In Lunar and Planetary Science XXIX, Abstract No.1821. Lunar and Planetary Institute, Houston, TX.Google Scholar
Hildebrand, A.R, Pilkington, M., Ortiz-Aleman, C., Chavez, R.E., Urrutia-Fucugauchi, J., Connors, M., Graniel-Castro, E., Camara-Zi, A., Halpenny, J.F. & Niehaus, D. (1998b). Mapping Chicxulub crater structure with gravity and seismic reflection data. J. Geol. Soc. Lond. Special Publications 140, 155176.CrossRefGoogle Scholar
Ivanov, B.A., Artemieva, N.A. & Pierazzo, E. (2004). Popigai impact structure modeling: Morphology and worldwide ejecta. In Lunar and Planetary Science XXXV, Abstract No.1240, pp. 16481649. Lunar and Planetary Institute, Houston, TX.Google Scholar
Jansa, L.F. (1993). Cometary impacts into ocean: Their recognition and the threshold constraints for biological extinction Palaeogeography, Palaeoclimatology, Palaeoecology 104, 271286.CrossRefGoogle Scholar
Keller, G., Adatte, T., Stinnesbeck, W., Rebolledo-Vieyra, M., Fucugauchi, J.U., Kramar, U., Stüben, D. & Morgan, W.J. (2004). Chicxulub impact predates the K-T boundary mass extinction. Proc. of the Nat. Acad. of Sci. of the U.S.A. 101(11), 37533758.CrossRefGoogle ScholarPubMed
Kelley, S.P. & Gurov, E. (2002). Boltysh, Another end-Cretaceous impact. Meteoritics & Planetary Science 37(8), 10311043.CrossRefGoogle Scholar
Kring, D.A. (2002). Re-evaluating the impact cratering kill curve. Meteorit. Planet. Sci. 37, 16481649.CrossRefGoogle Scholar
Kring, D.A. (2003). Environmental consequences of impact cratering events as a function of ambient conditions on Earth. Astrobiology 3, 133152.CrossRefGoogle ScholarPubMed
Kring, D.A. & Durda, D.D. (2002). Trajectories and distribution of material ejected from the Chicxulub impact crater: Implications for post-impact wildfires. J. Geophys. Res. 107(E6), 10.1029/2001JE001532.Google Scholar
Langenhorst, F. (1996). Characteristics of shocked quartz in Late Eocene impact ejecta. Clues to shock conditions and source crater. Geology 24, 487490.2.3.CO;2>CrossRefGoogle Scholar
Masaitis, V.L. (1994). Impactites from Popigai crater. In Large Meteorite Impacts and Planetary Evolution, ed. Dressler, B.O., Grieve, R.A.F. & Sharpton, V.L. Special Paper 293, pp. 153162. Geological Society of America.Google Scholar
Masaitis, V.L. (1998). Popigai crater: origin and distribution of diamond-bearing impactites. Meteorit. Plan. Sci. 33, 349359.CrossRefGoogle Scholar
Masaitis, V.L., Naumov, M.V. & Mashchak, M.S. (1999). Anatomy of the Popigai impact crater, Russia. In Large Meteorite Impacts and Planetary Evolution II, ed. Dressler, B.O. & Sharpton, V.L. Special Paper 339, pp. 117. Geological Society of America.Google Scholar
Melosh, H.J. & Ivanov, B.A. (1999). Impact crater collapse. Ann. Rev. Earth Planet. Sci. 27, 385415.CrossRefGoogle Scholar
Morgan, J.V., Warner, M.R. & the Chicxulub Working Group. (1997). Size and morphology of the Chicxulub impact crater. Nature 390, 472476.CrossRefGoogle Scholar
Morgan, J.V., Warner, M.R., Collins, G.S., Melosh, H.J. & Christeson, G.L. (2000). Peak ring formation in large impact craters. Earth Planet. Sci. Lett. 183(3–4), 347354.CrossRefGoogle Scholar
Murtaugh, J.C. (1976). Manicouagan impact structure Quebec Department of Natural Resources, Open-File Report DPV-432, 180 p.Google Scholar
Orphal, D.L. & Schultz, P.H. (1978). An alternative model for the Manicouagan impact structure. Proc. Lunar and Planetary Science Conference IX, 2, pp. 26952712.Google Scholar
Osinski, G.R. & Spray, J.G. (2001). Impact-generated carbonate melts: Evidence from the Haughton structure, Canada. Earth and Planetary Science Letters 194(1–2), 1729.CrossRefGoogle Scholar
Osinski, G.R. & Spray, J.G. (2003). Evidence for the shock melting of sulfates from the Haughton impact structure, Arctic Canada. Earth and Planetary Science Letters 215(3–4), 357370.CrossRefGoogle Scholar
Pierazzo, E., Hahmann, A.N. & Sloan, L.C. (2003). Chicxulub and climate: radiative perturbations of impact-produced S-bearing gases. Astrobiology 3, 99118.CrossRefGoogle ScholarPubMed
Poag, C.W. (1997). Roadblocks on the Kill Curve: testing the Raup Hypothesis. Palaios 12, 582590.CrossRefGoogle Scholar
Poag, C.W. & Aubry, M.P. (1995). Upper Eocene impactites of the U.S. East coast: depositional origins, biostratigraphic framework, and correlation. Palaios 10, 1643.CrossRefGoogle Scholar
Poag, C.W., Koeberl, C. & Reimold, W.U. (2004). Chesapeake Bay Crater. Geology and Geophysics of a Late Eocene Submarine Impact Structure. Springer Verlag, Heidelberg.Google Scholar
Poag, C.W., Powars, D.S., Poppe, L.J., Mixon, R.B., Edwards, L.E., Folger, D.W. & Bruce, S. (1992). Deep sea drilling project Site 612 bolide event: new evidence of a late Eocene impact-wave deposit and a possible impact site, U.S. east coast. Geology 20, 771774.2.3.CO;2>CrossRefGoogle Scholar
Pope, K.O., Baines, K.H., Ocampo, A.C. & Ivanov, B.A. (1997). Energy, volatile production, and climatic effects of the Chicxulub Cretaceous/Tertiary impact. J. Geophys. Res. 102(E9), 21 64521 664.CrossRefGoogle ScholarPubMed
Raup, D.M. (1992). Large-body impact and extinction in the Phanerozoic. Paleobiology 18, 8088.CrossRefGoogle ScholarPubMed
Robertson, D.S., McKenna, M.C., Toon, O.B., Hope, S. & Lillegraven, J.A. (2004). Survival in the first hours of the Cenozoic. Geol. Soc. Am. Bull. 116, 760768.CrossRefGoogle Scholar
Skelton, P.W. (2003). The Cretaceous World. Cambridge University Press and The Open University.Google Scholar
Smit, J. (1999). The global stratigraphy of the Cretaceous Tertiary boundary impact ejecta. Annu. Rev. Earth Planet. Sci. 27, 7591.CrossRefGoogle Scholar
Spray, J.G., Kelley, S.P. & Rowley, D.B. (1998). Evidence for a late Triassic multiple impact event on Earth. Nature 392, 171173.CrossRefGoogle Scholar
Tanner, L.H., Lucas, S.G. & Chapman, M.G. (2004). Assessing the record and causes of late Triassic extinctions. Earth Sci. Rev. 65, 103139.CrossRefGoogle Scholar
Thackrey, S. et al. (2008). Determining source of ejecta using heavy mineral provenance techniques; a Manicougan distal ejecta case study. In Lunar and Planetary Science XXXIX, Abstract No.1254. Lunar and Planetary Institute, Houston, TX.Google Scholar
Toon, O.B., Zahnle, K., Morrison, D., Turco, R.P. & Covey, C. (1997). Environmental perturbations caused by the impacts of asteroids and comets. Rev. Geophys. 35, 4178.CrossRefGoogle Scholar
Turtle, E.P., Pierazzo, E., Collins, G.S., Osinski, G.R., Melosh, H.J., Morgan, J.V. & Reimold, W.U. (2005). Impact structures: What does crater diameter mean? In Large Meteorite Impacts III, ed. Kenkmann, T., Hörz, F. & Deutsch, A., Special Paper 384, pp. 124. Geological Society of America.Google Scholar
Urrutia-Fucugauchi, J., Marín, L. & García Trejo, A. (1996). UNAM Scientific Drilling Program of Chicxulub impact structure – evidence for a 300 km crater diameter. Geophys. Res. Lett. 23, 15651568.CrossRefGoogle Scholar
Vishnevsky, S. & Montanari, A. (1999). Popigai impact structure (Arctic Siberia, Russia): Geology, petrology, geochemistry, and geochronology of glass-bearing impactites. In Large Meteorite Impacts and Planetary Evolution II, ed. Dressler, B.O. and Sharpton, V.L., Special Paper 339, pp. 1959. Geological Society of America.Google Scholar
Walkden, G., Parker, J. & Kelley, S.P. (2002). A Late Triassic impact ejecta layer in southwestern Britain. Science 298, 21852188.CrossRefGoogle ScholarPubMed
Walkden, G.M. & Parker, J. (2006). Large bolide impacts: is it only size that counts? Proc. 1st Int. Conf. on Impact Cratering in the Solar System, European Space Agency Special Publication No. 612.Google Scholar
Ward, S. & Asphaug, E. (2000). Asteroid impact tsunami: a probabilistic hazard assessment. Icarus 145, 6478.CrossRefGoogle Scholar
Wrobel, K.E. & Schultz, P.H. (2003). The effect of rotation on the deposition of terrestrial impact ejecta. In Lunar and Planetary Science XXIX, Abstract No.1190. Lunar and Planetary Institute, Houston, TX.Google Scholar