Hostname: page-component-848d4c4894-75dct Total loading time: 0 Render date: 2024-06-10T18:11:44.191Z Has data issue: false hasContentIssue false

Carbon-isotope stratigraphy of the uppermost Cambrian in eastern Laurentia: implications for global correlation

Published online by Cambridge University Press:  12 February 2018

KAREM AZMY*
Affiliation:
Department of Earth Sciences, Memorial University of Newfoundland, St. John's, Newfoundland A1B 3X5, Canada
*
*Author for correspondence: kazmy@mun.ca

Abstract

The δ13C profile from an interval of the Martin Point section in western Newfoundland (Canada) spans the upper Furongian (uppermost Cambrian). The interval (~90 m) is a part of the Green Point Formation of the Cow Head Group and consists of the Martin Point (lower) and the Broom Point (upper) members. It is formed of slope marine carbonates alternating with shales (rhythmites) and conglomeratic interbeds. The preservation of the investigated micritic carbonates was meticulously evaluated by multiple petrographic and geochemical screening tools. The δ13C and δ18O values (−0.5 ± 0.8 ‰VPDB and −7.1 ± 0.3 ‰VPDB, respectively) exhibit insignificant correlation (R2 = 0.002) and similarly the correlation of δ13C values with their Sr and Mn counterparts, which supports the preservation of at least near-primary δ13C signatures that can be utilized to construct a reliable high-resolution carbon-isotope profile for global correlations.

The δ13C profile exhibits two main negative excursions, a lower broad excursion (~3 ‰) that reaches its maximum at ~70 m below the Martin Point / Broom Point members boundary and an upper narrow excursion (~2.5 ‰) immediately below the same boundary. The lower excursion can be correlated with the global latest Furongian HERB event (TOCE), which is also recognized in the C-isotope profile of the GSSP boundary section at Green Point whereas the upper excursion matches with that of the Cambrian‒Ordovician boundary in the same section. The peak of the HERB δ13C excursion is correlated with positive shifts on the Th/U and Ni profiles (redox and productivity proxies).

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2018 

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

Arnaboldi, M. & Meyers, P. A. 2007. Trace element indicators of increased primary production and decreased water-column ventilation during deposition of latest Pliocene sapropels at five locations across the Mediterranean Sea. Palaeogeography, Palaeoclimatology, Palaeoecology 249, 425–43.Google Scholar
Azmy, K., Kaufman, A. J., Misi, A. & Oliveira, T. F. 2006. Isotope stratigraphy of the Lapa Formation, São Francisco Basin, Brazil: implications for Late Neoproterozoic glacial events in South America. Precambrian Research 149, 231–48.Google Scholar
Azmy, K., Kendall, K., Brand, U., Stouge, S. & Gordon, G. W. 2015. Redox conditions across the Cambrian-Ordovician boundary: elemental and isotopic signatures retained in the GSSP carbonates. Palaeogeography, Palaeoclimatology, Palaeoecology 440, 440–54.Google Scholar
Azmy, K., Stouge, S., Brand, U., Bagnoli, G. & Riiperdan, R. 2014. High-resolution chemostratigraphy of the Cambrian‒Ordovician GSSP in western Newfoundland, Canada: enhanced global correlation tool. Palaeogeography, Palaeoclimatology, Palaeoecology 409, 135‒44.Google Scholar
Azmy, K., Veizer, J., Misi, R., De Olivia, T., Sanches, A. L. & Dardenne, M. 2001. Isotope stratigraphy of the Neoproterozoic carbonate of Vazante Formation Saõ Francisco Basin, Brazil. Precambrian Research 112, 303–29.Google Scholar
Azomani, E., Azmy, K., Blamey, N., Brand, U. & Al-Aasm, I. 2013. Origin of Lower Ordovician dolomites in eastern Laurentia: controls on porosity and implications from geochemistry. Marine and Petroleum Geology 40, 99114.Google Scholar
Bahamonde, J. R., Merino-Tomé, O. A. & Heredia, N. 2007. A Pennsylvanian microbial boundstone-dominated carbonate shelf in a distal foreland margin (Picos de Europa Province, NW Spain). Sedimentary Geology 198, 167–93.Google Scholar
Banner, J. L. & Hanson, G. N. 1990. Calculation of simultaneous isotopic and trace element variations during water-rock interaction with applications to carbonate diagenesis. Geochimica et Cosmochimica Acta 54 (11), 3123‒37.Google Scholar
Barnes, C. R. 1988. The proposed Cambrian‒Ordovician global boundary stratotype and point (GSSP) in western Newfoundland, Canada. Geological Magazine 125, 381414.Google Scholar
Bartley, J. K., Kah, L. C., Frank, T. D. & Lyons, T. W. 2015. Deep-water microbialites of the Mesoproterozoic Dismal Lakes Group: microbial growth, lithification, and implications for coniform stromatolites. Geobiology 13, 1532.Google Scholar
Brand, U., Logan, A., Bitner, M. A., Griesshaber, E., Azmy, K. & Buhl, D. 2011. What is the ideal proxy of Paleozoic seawater? Memoirs of the Association of Australasian Palaeontologists 41, 924.Google Scholar
Buggisch, W., Keller, M. & Lehnert, O. 2003. Carbon isotope record of Late Cambrian to Early Ordovician carbonates of the Argentine Precordillera. Palaeogeography, Palaeoclimatology, Palaeoecology 195, 357–73.Google Scholar
Cawood, P. A., McCausland, P. J. A. & Dunning, G. R. 2001. Opening Iapetus: constraints from Laurentian margin in Newfoundland. Geological Society of America Bulletin 113, 443–53.Google Scholar
Chen, J., Zhang, J., Nicoll, R. S. & Nowlan, G. S. 1995. Carbon and oxygen isotopes in carbonate rocks within Cambrian–Ordovician boundary interval at Dayangcha, China. Acta Palaeontologica Sinica 34, 393408.Google Scholar
Coniglio, M. & James, N. P. 1990. Origin of fine-grained carbonate and siliciclastic sediments in an Early Paleozoic slope sequence, Cow Head Group, Western Newfoundland. Sedimentology 37, 215‒30.Google Scholar
Cooper, R. A., Nowlan, G. S. & Williams, S. H. 2001. Global Stratotype Section and Point for base of the Ordovician System. Episodes 24, 1928.Google Scholar
Della Porta, G., Kenter, J. A. M., Bahamonde, J. R., Immenhauser, A. & Villa, E. 2003. Microbial boundstone dominated carbonate slopes (Upper Carbonifierous, N. Spain): microfacies, lithofacies distribution and stratal geometry. Facies 49, 175207.Google Scholar
Derry, L. A., Kaufman, A. J. & Jacobsen, S. B. 1992. Sedimentary cycles and environmental change in the Late Proterozoic: evidence from stable and radiogenic isotopes. Geochimica et Cosmochimica Acta 56, 1317–29.Google Scholar
Dickson, J. A. D. 1966. Carbonate identification and genesis as revealed by staining. Journal of Sedimentary Petrology 36, 491505.Google Scholar
George, A. D. 1999. Deep-water stromatolites, Canning Basin, Northwestern Australia. Palaios 14, 493505.Google Scholar
Glumac, B. & Mutti, L. E. 2007. Late Cambrian (Steptoean) sedimentation and responses to sea-level change along the northeastern Laurentian margin: insights from carbon isotope stratigraphy. Geological Society of America Bulletin 119, 623–36.Google Scholar
Halverson, G. P., Hoffman, P. F., Schrag, D. P., Maloof, A. C. & Rice, A. H. N. 2005. Toward a Neoproterozoic composite carbon-isotope record. Geological Society of America Bulletin 117, 1181–207.Google Scholar
Hibbard, J. P., Van Staal, C. R. & Rankin, D. W. 2007. A comparative analysis of pre-Silurian crustal building blocks of the northern and southern Appalachian Orogen. American Journal of Science 307, 2345.Google Scholar
James, N. P. & Stevens, P. K. 1986. Stratigraphy and correlation of the Cambro–Ordovician Cow Head Group, western Newfoundland. Geological Survey of Canada Bulletin 366, 143 pp.Google Scholar
James, N. P., Stevens, R. K., Barnes, C. R. & Knight, I. 1989. Evolution of a Lower Paleozoic continental-margin carbonate platform, northern Canadian Appalachians. In Controls on Carbonate Platform and Basin Development (eds Crevello, P. D., Wilson, J. L., Sarg, J. F. & Read, J. F.), pp. 123–46. Society of Economic Paleontologists and Mineralogists, Special Publication no. 44.Google Scholar
Jing, X-C, Deng, S-H, Zhao, Z-J, Lu, Y-Z. & Zhang, S-B. 2008. Carbon isotope composition and correlation across the Cambrian‒Ordovician boundary in Kalpin Region of the Tarim Basin, China. Science in China: Earth Sciences 51, 1317–59.Google Scholar
Landing, E. 2007. Ediacaran–Ordovician of east Laurentia – geologic setting and controls on deposition along the New York Promontory. In Ediacaran–Ordovician of East Laurentia: S. W. Ford Memorial Volume (ed. Landing, E.), pp. 524. New York State Museum Bulletin 510.Google Scholar
Landing, E. 2012. Time-specific black mudstones and global hyperwarming on the Cambrian–Ordovician slope and shelf of the Laurentia palaeocontinent. Palaeogeography, Palaeoclimatology, Palaeoecology 367–368, 256–72.Google Scholar
Landing, E. 2013. The Great American Carbonate Bank in northeast Laurentia: its births, deaths, and linkage to continental slope oxygenation (Early Cambrian–Late Ordovician). In The Great American Carbonate Bank, Essays in Honor of James Lee Wilson (eds Derby, J. R., Fritz, R. D., Longacre, S. A., Morgan, W. A. & Sternbach, C. A.), 451–92. American Association of Petroleum Geologists Bulletin, Memoir 98.Google Scholar
Landing, E., Westrop, S. R. & Adrain, J. M. 2011. The Lawsonian Stage – the Eoconodontus notchpeakensis (Miller, 1969) FAD and HERB carbon isotope excursion define a globally correlatable terminal Cambrian stage. Bulletin of Geosciences 86 (3), 621–40.Google Scholar
Landing, E., Westrop, S. R. & Miller, J. F. 2010. Globally practical base for the uppermost Cambrian (Stage 10): FAD of the conodont Eoconodontus notchpeakensis and the Housian [sic, read ‘Lawsonian’ as the abstract text] Stage. In The 15th Field Conference of the Cambrian Stage Subdivision Working Group. Abstracts and Excursion Guide, Prague, Czech Republic, and south-eastern Germany (eds Fatka, O. & Budil, P.), p. 18. Prague: Czech Geological Survey.Google Scholar
Lavoie, D., Desrochers, A., Dix, G., Knight, I. & Hersi, O. S. 2012. The great carbonate Bank in eastern Canada: an overview. In The Great American Carbonate Bank: The Geology and Economic Resources of Cambrian‒Ordovician Sauk Megasequence of Laurentia (eds Derby, J., Fritz, R., Longcare, S., Morgan, W. & Sternbach, C.), pp. 499524. American Association of Petroleum Geologists, Memoir 98.Google Scholar
Li, X., Jenkyns, H. C., Wang, C., Hu, X., Chen, X., Wei, Y., Huang, Y. & Cui, J. 2006. Upper Cretaceous carbon- and oxygen-isotope stratigraphy of hemipelagic carbonate facies from southern Tibet, China. Journal of Geological Society of London 163, 375‒82.Google Scholar
Li, D., Zhang, X., Chen, K., Zhang, G., Chen, X., Huang, W., Peng, S. & Shen, Y. 2017. High-resolution C-isotope chemostratigraphy of the uppermost Cambrian stage (Stage 10) in South China: implications for defining the base of Stage 10 and palaeoenvironmental change. Geological Magazine 1, 112.Google Scholar
Machel, H. G. & Burton, E. A. 1991. Factors governing cathodoluminescence in calcite and dolomite, and their implications for studies of carbonate diagenesis. In Luminescence Microscopy and Spectroscopy, Qualitative and Quantitative Applications (eds Barker, C. A. & Kopp, O. C.), pp. 3757. Society of Economic Paleontologists and Mineralogists (SEPM) Short Course Notes 25.Google Scholar
Miller, J. F., Evans, K. R., Freeman, R. L., Ripperdan, R. L. & Taylor, J. F. 2011. Proposed stratotype for the base of the Lawsonian Stage (Cambrian Stage 10) at the first appearance datum of Econodontus notchpeakensis (Miller) in the house Range, Utah, USA. Bulletin of Geosciences 86, 595620.Google Scholar
Miller, J. F., Repetski, J. E., Nicoll, R. S., Nowlan, G. & Ethington, R. L. 2014. The conodont Iapetognathus and its value for defining the base of the Ordovician System. GFF 136, 185–8.Google Scholar
Playton, T. E., Janson, X. & Kerans, C. 2010. Carbonate slopes. In Facies Models 4 (eds James, N. P. & Dalrymple, R. W.). GEOtext 6. St John's: Geological Association of Canada, 586 pp.Google Scholar
Ripperdan, R. L., Magaritz, M. & Kirschvink, J. L. 1993. Carbon isotope and magnetic polarity evidence for non-depositional events within the Cambrian‒Ordovician boundary section at Dayangcha. Jilin Province, China. Geological Magazine 130, 443–52.Google Scholar
Ripperdan, R. L., Magaritz, M., Nicoll, R. S. & Shergold, J. H. 1992. Simultaneous changes in carbon isotopes, sea level, and conodont biozones within the Cambrian-Ordovician boundary interval at Black Mountain, Australia. Geology 20, 1039–41.Google Scholar
Rush, P. F. & Chafetz, H. S. 1990. Fabric retentive, non-luminescent brachiopods as indicators of original δ13C and δ18O compositions: a test. Journal of Sedimentary Petrology 60, 968–81.Google Scholar
Sial, A. N., Peralta, S., Gaucher, C., Alonso, R. N. & Pimentel, M. A. 2008. Upper Cambrian carbonate sequences of the Argentine Precordillera and the Steptoean C-isotope positive excursion (SPICE). Gondwana Research 13, 437–52.Google Scholar
Śliwiński, M. G., Whalen, M. T. & Day, J. 2010. Trace element variations in the middle Frasnian punctata zone (late Devonian) in the western Canada sedimentary basin – changes in oceanic bioproductivity and paleoredox spurred by a pulse of terrestrial afforestation? Geologica Belgica 13, 459–82.Google Scholar
Stouge, S., Bagnoli, G. & Azmy, K. 2016. The Cambrian HERB excursion (Furongian) from the Martin Point Member of the Cow Head Group, western Newfoundland, Canada. In Palaeo Down Under 2, Adelaide, July 2016 (eds Laurie, J. R., Kruse, P. D., Garcia-Bellido, D. C. & Holmes, J. D.), p. 56. Geological Society of Australia Abstracts 117.Google Scholar
Terfelt, F., Eriksson, M. E. & Schmitz, B. 2014. The Cambrian–Ordovician transition in dysoxic facies in Baltica – diverse faunas and carbon isotope anomalies. Palaeogeography, Palaeoclimatology, Palaeoecology 394, 5973.Google Scholar
Veizer, J. 1983. Chemical diagenesis of carbonates. In Theory and Application of Trace Element Technique: Stable Isotopes in Sedimentary Geology (eds Arthur, M. A., Anderson, T. F., Kaplan, I. R., Veizer, J. & Land, L. S.), pp. III-1–III-100. Society of Economic Paleontologists and Mineralogists (SEPM) Short Course Notes 10.Google Scholar
Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Bruhn, F., Buhl, D., Carden, G., Diener, A., Ebneth, S., Goddris, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O. & Strauss, H. 1999. 87Sr/86Sr, δ18O and δ13C evolution of Phanerozoic seawater. Chemical Geology 161, 5988.Google Scholar
Wignall, P. B. & Twitchett, R. J. 1996. Oceanic anoxia and the end Permian mass extinction. Science 272, 1155–8.Google Scholar
Wignall, P. B., Zonneveld, J.-P., Newton, R. J., Amor, K., Sephton, M. A. & Hartley, S. 2007. The end Triassic mass extinction record of Williston Lake, British Columbia. Palaeogeography, Palaeoclimatology, Palaeoecology 253, 385406.Google Scholar
Williams, S. H. & Stevens, R. K. 1991. Late Tremadoc graptolites from western Newfoundland. Palaeontology 34, 147.Google Scholar
Wilson, J. L., Medlock, P. L., Fritz, R. D., Canter, K. L. & Geesaman, R. G. 1992. A review of Cambro-Ordovician breccias in North America. In Paleokarst, Karst-Related Diagenesis and Reservoir Development: Examples from Ordovician–Devonian Age Strata of West Texas and Mid-Continent (eds Candelaria, M. P. & Reed, C. L.), pp. 1929. Permian Basin Section, Society of Economic Paleontologists and Mineralogists. SEPM Publication no. 92-33.Google Scholar