Hostname: page-component-76fb5796d-r6qrq Total loading time: 0 Render date: 2024-04-25T14:48:39.768Z Has data issue: false hasContentIssue false

Compositional trends of a Cretaceous foreland basin shale (Belle Fourche Formation, Western Canada Sedimentary Basin): diagenetic and depositional controls

Published online by Cambridge University Press:  09 July 2018

P. de Caritat
Affiliation:
Department of Geology and Geophysics, The University of Calgary, Calgary, Alberta T2N 1N4, Canada
J. D. Bloch
Affiliation:
Institute of Sedimentary and Petroleum Geology, Geological Survey of Canada, Calgary, Alberta T2L 2A7, Canada
I. E. Hutcheon
Affiliation:
Department of Geology and Geophysics, The University of Calgary, Calgary, Alberta T2N 1N4, Canada
F. J. Longstaffe
Affiliation:
Department of Earth Sciences, The University of Western Ontario, London, Ontario N6A 5B7, Canada

Abstract

Compositional trends of the Cenomanian Belle Fourche Formation, a marine shale unit in the Western Canada Sedimentary Basin, have been investigated on a regional scale using bulk-rock geochemistry and mineralogy, clay mineral compositions and oxygen isotope geochemistry of shale and bentonite core samples. Smectitic illite-smectite found in the matrix of immature, hemipelagic samples is compositionally and isotopically consistent with an origin from low-temperature alteration of volcanic ash in the central Western Interior Seaway, where the basin received minimal detrital input. The origin of the more illitic matrix in the deeply buried, western, pro-deltaic shales can be interpreted in terms of either diagenetic ‘illitization’ of a smectitic precursor, or depositional mixing of abundant, detrital, illitic material with minor amounts of ashfall-derived smectite. It is concluded that: (1) documented silicate mineral reactions during deep diagenesis of the Belle Fourche Formation took place in a relatively closed system, with no significant import or export of mobile species at the formation scale; and (2) diagenesis and depositional mixing can have similar effects in terms of bulk-rock and oxygen isotope geochemistry, and mineral compositions and assemblages.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 1994

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

Baedecker, P.A. (editor) (1987) Methods for Geochemical Analysis. US Geol. Surv. Bull. 1770.Google Scholar
Berner, R.A. & Raiswell, R. (1984) C/S method for distinguishing freshwater from marine sedimentary rocks. Geology 12, 365368.Google Scholar
Berner, R.A. (1970) Sedimentary pyrite formation. Am. J. Sci. 268, 123.CrossRefGoogle Scholar
Bhattacharya, J. & Walker, R.G. (1991a) Allostratigraphic subdivision of the Upper Cretaceous Dunvegan, Shaftesbury, and Kaskapau formations, in the northwestern Alberta subsurface. Bull. Can. Petrol. Geol. 39, 145164.Google Scholar
Bhattacharya, J. & Walker, R.G. (1991b) River- and wave-dominated depositional systems of the Upper Cretaceous Dunvegan Formation, northwestern Alberta. Bull. Can. Petrol. Geol. 39, 165191.Google Scholar
Bjørlykke, K. & Aagaard, P. (1992) Clay minerals in North Sea sandstones. Pp. 65-80 in: Origin, Diagenesis, and Petrophysics of Clay Minerals in Sandstones (Houseknecht, D.W. & Pittman, E.D., editors). SEPM Special Publication 47.Google Scholar
Beoch, J. & Hutcheon, I.E. (1994) Shale diagenesis: a study from the Albian Harmon Member (Peace River Formation), western Canada. Clays Clay Miner. 40, 682–699.Google Scholar
Bloch, J. & Krouse, H.R. (1992) Sulfide diagenesis and sedimentation in the Abian Harmon Member, Western Canada. J. Sed. Pet. 62, 235249.Google Scholar
Bloch, J., Schroder-Adams, C., Leckie, D.A., Mcintyre, D.J., Craig, J. & Staniland, M. (1993) Revised stratigraphy of the lower Colorado Group (Albian to Turonian), western Canada. Bull. Can. Petrol. Geol. 41, 325–248.Google Scholar
Boles, J.R. (1982) Active albitization of plagioclase, Gulf Coast Tertiary. Am. J. Sci. 282, 165180.Google Scholar
Cadrin, A.A.J., Kyser, T.K., Caldwell, W.G.E. & Loncstaffe, F.J. (1994) Isotopic and chemical compositions of bentonites as paleoenvironmental indicators of the Cretaceous western interior seaway. Geochim. Cosmochim. Acta (submitted).Google Scholar
Caritat, P. DE, Bloch, J. & Hutcheon, I. (1994) LPNORM: a linear programming normative analysis code. Computers & Geosciences 20, 313347.Google Scholar
Clayton, R.N. & Mayeda, T.K. (1963) The use of bromine pentafluoride in the extraction of oxygen from oxides and silicates for isotopic analysis. Geochim. Cosmochim. Acta 27, 4352.Google Scholar
Craig, H. (1961) Standards for reporting concentrations of deuterium and oxygen-18 in natural waters. Science 133, 18331834.CrossRefGoogle ScholarPubMed
Davis, H.R. (1987) Deposition of the Lower Cretaceous Mowry Shale. PhD thesis, Univ. Wisconsin-Madison, USA.Google Scholar
Dean, W.E. & Arthur, M.A. (1989) Iron-sulfur-carbon relationships in organic-carbon-rich sequences I: Cretaceous Western Interior Seaway. Am. J. Sci. 289, 708743.Google Scholar
Eslinger, E.V. & Yen, H.-W. (1986) Oxygen and hydrogen isotope geochemistry of Cretaceous bentonites and shales from the Disturbed Belt, Montana. Geochim. Cosmochim. Acta 50, 5968.Google Scholar
Espitalié, J., Laporte, J.L., MADECM., Marquis, F., Leplai, P., Paulet, J. & Boutefeu, A. (1977) Méthode rapide de caractérisation des roches mères, de leur potentiel pétrolier et de leur degré d'évolution. Rev. Inst. franç. pét. 32, 2342.Google Scholar
Gautier, D.L. (1987) Isotope composition of pyrite: relationship to organic matter type and iron availability in some North American Cretaceous shales. Chem. Geol. 65, 293303.Google Scholar
Griffen, G.M. (1962) Regional clay-mineral facies— products of weathering intensity and current distribution in the northeastern Gulf of Mexico. Bull. Geol. Soc. Amer. 73, 737768.Google Scholar
Hayes, M.J. & Boles, J.R. (1992) Volumetric relations between dissolved plagioclase and kaolinite in sandstones: implications for aluminium mass transfer in the San Joachin Basin, California. Pp. 111-123 in: Origin, Diagenesis, and Petrophysics of Clay Minerals in Sandstones (Houseknecht, D.W. & Pittman, E.D., editors). SEPM Special Publication 47.Google Scholar
Huggett, J.M. (1992) Petrography, mineralogy and diagenesis of overpressured Tertiary and Late Cretaceous mudrocks for the east Shetland Basin. Clay Miner. 27, 487506.Google Scholar
Jiang, W.-T., Peacor, D.R., Merriman, R.J. & Roberts, B. (1990) Transmission and analytical electron microscopic study of mixed-layer illite/smectite formed as an apparent replacement product of diagenetic illite. Clays Clay Miner. 38, 449468.Google Scholar
Johnson Ibach, L.E. (1982) Relationship between sedimentation rate and total organic content in ancient marine sediments. Am. Ass. Petrol. Geol. Bull. 66, 170188.Google Scholar
Kyser, T.K., Caldwell, W.G.E., Whittaker, S.G. & Cadrin, A.J. (1993) Paleoenvironment and geochemistry of the northern portion of the Western Interior Seaway during Late Cretaceous time. Geol. Soc. Canada Spec. Paper 39, 355378.Google Scholar
Land, L.S. & Ducron, S.P. (1978) Cementation of a Pennsylvanian deltaic sandstone: isotopic data. J. Sed. Pet. 48, 11671176.Google Scholar
Leckie, D.A. (1989) Upper Zuni sequence: Middle Cretaceous to Lower Tertiary. Pp. 269-286 in: Western Canada Sedimentary Basin: A Case History (Ricketts, B.D., editor). Canadian Society of Petroleum Geologists Special Publication.Google Scholar
Lindgreen, H. & Hansen, P.L. (1991) Ordering of illitesmectite in Upper Jurassic claystones from the North Sea. Clay Miner. 26, 105125.Google Scholar
Lonostafee, F.J. & Avalon, A. (1987) Oxygen-isotope studies of clastic diagenesis in the Lower Cretaceous Viking Formation, Alberta: implications for the role of meteoric water. Pp. 277-296 in: The Diagenesis of Sedimentary Sequences (Marshall, J.D., editor). Geological Society Special Publication 36.Google Scholar
Longstaffe, F.J., Tilley, B.J., Ayalon, A. & Connolly, C.A. (1992) Controls on porewater evolution during sandstone diagenesis, Western Canada Sedimentry Basin: an oxygen-isotope perspective. Pp. 13-34 in: Origin, Diagenesis, and Petrophysics of Clay Minerals in Sandstones (Houseknecht, D.W. & Pittman, E.D., editors). SEPM Special Publication 47.Google Scholar
Mccloskey, W.G. & Busrln, R.M. (1992) Controls on distribution of organic matter in Cretaceous black shales, Colorado Group, Western Canada. 29th In. Geol. Congr. Abstract 1, 238.Google Scholar
Nadeau, P.H. & Reynolds, R.C. Jr. (1981) Volcanic components in pelitic sediments. Nature 294, 7274.Google Scholar
Nicholls, J. (1988) The statistics of Pearce element diagrams and the Chayes closure problem. Contrib. Mineral. Petrol. 99, 1124.Google Scholar
Pearce, T.H. (1967) A contributon to the theory of variation diagrams. Contrib. Mineral. Petrol. 19, 142157.CrossRefGoogle Scholar
Pettijohn, F.J. (1975) Sedimentary Rocks. Third edition, Harper & Row, New York.Google Scholar
Potter, P.E., Maynard, J.B. & Pryor, W.A. (1980) Sedimentology of Shale. Springer-Verlag, New York.Google Scholar
Sackerr, W.M. (1964) The depositional history and isotopic organic carbon composition of marine sediments. Mar. Geol. 2, 173185.Google Scholar
Savin, S.M. & Lee, M. (1988) Isotopic studies of phyllosilicates. Pp. 189-223 in: Hydrous Phyllosilicates (Exclusive of Micas) (Bailey, S.W., editor). Reviews in Mineralogy 19, Mineralogical Society of America, Washington, DC.Google Scholar
Schultz, L.G., Shepard, A.O., Blackman, P.D. & Starkly, H.C. (1971) Mixed-layer kaolinite-montmorillonite from the Yucatan Peninsula, Mexico. Clays Clay Miner. 19, 137150.Google Scholar
Shaw, D.B. & Weaver, C.E. (1965) The mineralogical composition of shales. J. Sed. Pet. 35, 213222.Google Scholar
Środoń, J., Elsass, F., Mchardy, W.J. & Morgan, D.J. (1992) Chemistry of illite-smectite inferred from TEM measurements of fundamental particles. Clay Miner. 27, 137158.Google Scholar
Stein, R. (1990) Organic carbon content/sedimentation rate relationship and its paleoenvironmental significance for marine sediments. Geo-Marine Letters 10, 3744.Google Scholar
Taylor, E., Burkett, P.J., Wackler, J.D. & Leonard, J.N. (1991) Physical properties and microstructual response of sediments to accretion-subduction: Barbados Forearc. Pp. 213-228 in: Microstructures of Fine-Grained Sediments, from Mud to Shale (Bennett, R.H., Bryant, W.R., & Hulbert, M.H., editors). Frontiers in Sedimentary Geology, Springer-Verlag, Berlin.Google Scholar
Thode, H.G., Kleerekoper, H. & Mcelcheran, D. (1951) Isotopic fractionation in the bacterial reduction of sulfate. Research 4, 581582.Google Scholar
Veblen, D.R., Guthrie, G.D. Jr., Livi, K.J.T. & Reynolds, R.C. Jr. (1990) High-resolution transmission electron microscopy and electron diffraction of mixed-layer illite/smectite: experimental results. Clays Clay Miner. 38, 113.Google Scholar
Yau, Y.-C., Peacor, D.R. & Mcdowell, S.D. (1987) Smectite-to-illite reactions in Salton Sea shales: a transmission and analytical electron microscopy study. J. Sed. Pet. 57, 335342.Google Scholar