Hostname: page-component-848d4c4894-v5vhk Total loading time: 0 Render date: 2024-07-02T09:41:47.817Z Has data issue: false hasContentIssue false

Clay- and zeolite-bearing Triassic sediments at Kaka Point, New Zealand: evidence of microbially influenced mineral formation from earliest diagenesis into the lowest grade of metamorphism

Published online by Cambridge University Press:  09 July 2018

C. V. Jeans
Affiliation:
Department of Earth Sciences, Downing Street, Cambridge CB2 3EQ, UK
A. E. Fallick
Affiliation:
Scottish Universities Research & Reactor Centre, East Kilbride, Glasgow G75 OQF, UK
M. J. Fisher
Affiliation:
Nevis Associates Ltd., Helensburgh, Dumbartonshire G84 8DD, UK
R. J. Merriman
Affiliation:
British Geological Survey, Nicker Hill Keyworth, Nottinghamshire NG12 5GG, UK
R. M. Corfield
Affiliation:
Department of Earth Sciences, Parks Road, Oxford OX1 3PR, UK
B. Manighetti
Affiliation:
Department of Earth Sciences, Downing Street, Cambridge CB2 3EQ, UK

Abstract

The distribution, mineralogy, petrology and bulk and stable isotope chemistry of altered volcanic ash beds in the marine sediments of Mid-Triassic age (Etalian) at Kaka Point, New Zealand, are described and related to lithofacies and the geological processes controlling their development.

Three varieties of altered ash occur in the Kaka Point sediments — porcellanite, claystone (bentonite) and albite-rich. Porcellanites are quartz-rich and may contain analcime and heulandite: they are restricted mainly to the on-shore facies. Claystones are rich in smectitic clay minerals and occur in both the on-shore and off-shore facies. They often contain diagenetic nodules of analcime, quartz, apatite and carbonates. The authigenic carbonates of the on-shore facies are variable in composition (sideritic, rhodochrositic, calcitic), whereas in the off-shore facies they consist only of calcite. The albite-rich lithology is very rare and is known only from the off-shore facies.

The development of the porcellanite and albite-rich lithologies was restricted to slowly deposited, relatively coarse-grained ash sediments in which extensive interchange took place between the sediment's pore-waters and ambient seawater, resulting in enhanced microbial activity and high pH throughout the pore-waters of the suboxic zone beneath the water-sediment interface. The high pH increased the rate of volcanic ash hydrolysis and provided the conditions necessary for the precipitation of zeolite, feldspar and quartz. The development of smectitic claystones was associated with more rapid deposition and limited interchange between the pore-waters of the parent ash and ambient seawater. The pore-water alkalinity was generally lower and enhanced microbial activity and high pHs were restricted to patches of sediment at which quartz, analcime, apatite and carbonates formed diagenetic nodules. Modelling of the stable isotopes of the smectitic clays (δ18O, δD) and diagenetic carbonates (δ18O, δ13C) suggest that: (1) ash argillization in the on-shore facies took place in brackish water (∼25% meteoric water) at an average temperature of ∼50°C and in the off-shore facies in marine pore-waters (∼10% meteoric waters) at ∼40°C and (2) diagenetic carbonate precipitation in the near-shore facies took place at ∼30°C and in the off-shore facies at 60–80°C.

The pattern of ash alteration in the marine Triassic sediments at Kaka Point is considered to represent an early stage in the development of the zeolite pattern associated with the classic area of zeolite facies metamorphism in the Taringatura and Hokonui Hills.

Type
Research Article
Copyright
Copyright © The Mineralogical Society of Great Britain and Ireland 1997

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

Bigeleisen, J., Perlman, M.L. & Prosser, H.C. (1952) Conversion of hydrogenic materials to hydrogen for isotopic analysis. Anal. Chem. 24, 1356–1357.CrossRefGoogle Scholar
Boles, J.R. (1974) Structure, stratigraphy and petrology of mainly Triassic rocks, Hokonui Hills, Southland, New Zealand. N.Z. J. Geol. Geophys. 17, 324334.CrossRefGoogle Scholar
Boles, J.R. & Coombs, D.S. (1975) Mixed reactions in zeolitic Triassic tuff, Hokonui Hills, New Zealand. Bull. Geol. Soc. Amer. 86, 163173.2.0.CO;2>CrossRefGoogle Scholar
Boles, J.R. & Coombs, D.S. (1977) Zeolite facies alteration of sandstones in the Southland Syncline, New Zealand. Am. J. Sci. 277, 9821012.CrossRefGoogle Scholar
Borthwick, J. & Harmon, R.S. (1982) A note regarding ClF3 as an alternative to BrF5 for oxygen isotope analysis. Geochim. Cosmochim. Acta, 46, 16651668.CrossRefGoogle Scholar
Campbell, J.D. & Coombs, D.S. (1966) Murihiku Supergroup (Triassic-Jurassic) of Southland and South Otago. N.Z J. Geol. Geophys. 9, 393398.CrossRefGoogle Scholar
Capuano, R.M. (1992) The temperature dependence of hydrogen isotope fractionation between clay minerals and water: evidence from a geopressured system. Geochim. Cosmochim. Acta, 56, 2547–2554.Google Scholar
Clayton, C.J. (1986) The chemical environment of flint formation in Upper Cretaceous chalks. Pp. 43–54 in: The Scientific Study of Flint and Chert, (de C. Siveking, G. & Hart, M.B., editors). Cambridge University Press, Cambridge.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
Coleman, M.L. (1985) Geochemistry of diagenetic nonsilicate minerals: kinetic considerations. Phil. Trans. R. Soc. A31g, 39-56.Google Scholar
Colman, S.M. & Dethier, D.P. (1986) Rates of Chemical Weathering of Rocks and Minerals. Academic Press, New York & London.Google Scholar
Coombs, D.S. (1954) The nature and alteration of some Triassic sediments from Southland, New Zealand. Trans. R. Soc. N.Z. 82, 65109.Google Scholar
Coombs, D.S. (1965) Sedimentary analcime rocks and sodium-rich gneisses. Mineral. Mag. 34, 144–158.Google Scholar
Coombs, D.S. & Cox, S.C. (1991) Low- and very low grade metamorphism in southern New Zealand and its geological setting. Geol. Soc. N. Z. M&c. Publ. 58. 79 pp.Google Scholar
Corfield, R.M. & Cartlidge, J.E. (1992) Oceanographic and climatic implications of the Palaeocene carbon isotope maximum. Terra Nova, 4, 443–455.Google Scholar
Cornford, C. (1990) Source rocks and hydrocarbons of the North Sea. Pp. 294–361 in: Introduction to the Petroleum Geology of the North Sea, 3rd ed., (Glennie, K.W., editor), Blackwell Scientific Publications, Oxford.Google Scholar
Correns, C.W. (1961) The experimental chemical weathering of silicates. Clay Miner. Bull. 4, 249265.Google Scholar
Doremus, R.H. (1975) Interdiffusion of hydrogen and alkali ions in a glass surface. J. Non-Cryst. Solids, 19, 137144.Google Scholar
Drever, J.I. (1985) The Chemistry of Weathering. NATO ASI Series, Reidel, Dordrecht.Google Scholar
Duchaufor, P., Bonneau, M. & Souchier, B. (1977) Pidologie 1. Pedogenèse et Classification. Masson, Paris.Google Scholar
Fisher, M.J. (1980) Kerogen distribution and depositional environments in the Middle Jurassic of Yorkshire, U.K. Proc. lVth Int. Palynological Confi, Lucknow 2, 574-580.Google Scholar
Hemley, J.J. (1962) Alteration studies in the system Na2o-Al2O3-SiO2-H2O and K2O-Al2O3-SiO2-H2O. Geol. Soc. Amer. Spec. Papers, 68, 322.Google Scholar
Hemley, J.J. & Janes, W.R. (1964) Chemical aspects of hydrothermal alteration with emphasis on hydrogen metasomatism. Econ. Geol. 59, 538–569.Google Scholar
Hogg, A.J.C., Pearson, M.J. & Fallick, A.E. (1993) Pretreatment of Fithian illite for oxygen isotope analysis. Clay Miner. 28, 149152.Google Scholar
Höller, H. & Wirsching, U. (1978) Experiments on the formation of zeolites by hydrothermal alteration of volcanic glasses. Pp. 329–336 in: Natural Zeolites, (Sand, L.B. & Mumpton, F.A., editors), Pergamon Press, Oxford.Google Scholar
Houser, C.A., Herman, J.S., Tsong, I.S.T., White, W.B. & Landford, W.A. (1980) Sodium-hydrogen interdiffusion in sodium silicate glasses. J. Non-Cryst. Solids, 41, 8998.Google Scholar
IAEA (1992) Statistical treatment of data on environmental isotopes in precipitation. Technical Report Series No. 331, IAEA, Vienna.Google Scholar
Jeans, C.V. (1980) Early submarine lithification in the Red Chalk and Lower Chalk of Eastern England: a bacterial control mechanism and its implications. Proc. Yorks. Geol. Soc. 43, 81157.Google Scholar
Jeans, C.V., Merriman, R.J. & Mitchell, J.G. (1977) Origin of Middle Jurassic and Lower Cretaceous fuller's earths in England. Clay Miner. 12, 11–44.CrossRefGoogle Scholar
Kisch, H.J. (1987) Correlation between indicators of very low-grade metamorphism. Pp. 227–304 in: Low Temperature Metamorphism, (Frey, M., editor), Blackie, Glasgow.Google Scholar
Kisch, H.J. (1991) Illite crystallinity: recommendations on sample preparation X-ray diffraction settings and inter-laboratory samples. J. Met. Geol. 9, 665–670.Google Scholar
Lanford, W.A., Davis, K., Lamarche, P., Laursen, T. & Groleau, R. (1979) Hydration of soda-lime glass. J. Non-Cryst. Solids, 33, 249266.Google Scholar
Le Maitre, R.W. (Ed.) (1989) A Classification of Igneous Rocks and Glossary of Terms. Blackwell Scientific Publications, Oxford.Google Scholar
Longstaffe, F.J. (1989) Stable isotopes as tracers in clastic diagenesis. Pp. 201–277 in: A Short Course in Burial Diagenesis, 15 (Hutcheon, I.E., editor). Min. Assoc. Canada Short Course Series.Google Scholar
Lovering, T.G. & Patten, L.E. (1962) The effect of CO2 at low temperature and pressure on solutions supersaturated with silica in the presence of limestone and dolomite. Geochim. Cosmochim. Acta, 26, 787–796.Google Scholar
Mariner, R.H. & Surdam, R.C. (1970) Alkalinity and formation of zeolites in saline alkaline lakes. Science, 170, 977980.CrossRefGoogle ScholarPubMed
McCrea, J.M. (1950) On the isotopic chemistry of carbonates and a paleotemperature scale. J. Chem. Phys. 18, 849857.CrossRefGoogle Scholar
Mortimer, N. (1993) Geology of the Otago Schist and Adjacent Rocks, IGNS Geological Map 7. Institute of Geological and Nuclear Sciences Ltd., Lower Hutt, New Zealand.Google Scholar
Okamoto, G., Okura, T. & Goto, K. (1957) Properties of silica in water. Geochim. Cosmochim. Acta, 12, 123132.Google Scholar
Roberts, B., Merriman, R.J. & Pratt, W. (1991) The influence of strain, lithology and stratigraphical depth of white mica (illite) crystallinity in mudrocks from the vicinity of the Corris Slate Belt, Wales: implications for the timing of metamorphism in the Welsh Basin. Geol. Mag. 128, 633645.Google Scholar
Rosenbaum, J. & Sheppard, S.M.F. (1986) An isotopic study of siderites, dolomites and ankerites at high temperature. Geochim. Cosmochim. Acta, 50, 11471150.Google Scholar
Shackleton, N.J. & Kennett, J.P. (1975) Paleotemperature history of the Cenozoic and the initiation of Antarctic glaciation: oxygen and carbon analyses in DSDP sites 277, 279, 281. Pp. 653-659 in: Initial Report DSDP 24, (Kennett, J.P. & Howtz, R.E., editors). Washington.Google Scholar
Smets, B.M. & Lommen, T.P.A. (1982) The leaching of sodium aluminosilicate glasses studied by secondary ion mass spectrometry. Phys. Chem. Glasses, 23, 8387.Google Scholar
Smith, A.G., Smith, D.G. & Funnell, B.M. (1994) Atlas of Mesozoic and Cenozoic Coastlines. Cambridge University Press, Cambridge.Google Scholar
Suggate, R.P., Stevens, G.R. & Te Punga (editors) (1978) The Geology of New Zealand, 2 vols. Government Printers, Wellington.Google Scholar
Surdam, R.C. & Sheppard, R.A. (1978) Zeolites in saline, alkaline-lake deposits. Pp. 145-174 in: Natural Zeolites, (Sand, L.B. & Mumpton, F.A., editors). Pergamon Press, Oxford.Google Scholar
Tanner, C.B. & Jackson, M.L. (1947) Nomographs of sedimentation times for soil particles under gravity or centrifugal acceleration. Proc. Soil Sci. Soc. Am. 12, 6065.Google Scholar
Taylor, H.P. (1974) The application of oxygen and hydrogen isotope studies to problems of hydrother- mal alteration and ore deposition. Econ. Geol. 69, 843–883.Google Scholar