Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-05-31T15:59:47.472Z Has data issue: false hasContentIssue false

Precambrian rift-related magmatism and sedimentation, south Victoria Land, Antarctica

Published online by Cambridge University Press:  16 August 2007

Y.A. Cook
Affiliation:
Geology Department, University of Otago, Dunedin, New Zealand

Abstract

Precambrian continental extension is described in detail for the first time in the Victoria Land segment of the Transantarctic Mountains and is comparable with plume related intercontinental rifting of the Afar area, Africa. The Baronick Formation comprises igneous-derived conglomerate, marble and volcanic to sub-volcanic igneous layers. Volcanic and carbonate horizons were eroded in a fluvial or marine environment and provided debris for mass flow and slump deposits which formed in a marginal marine basin in the Precambrian. Clasts in these deposits include basalt, trachyte and comendite, and along with the interbedded volcanic layers of basalt, trachyte and quartz syenite, indicate proximity and contemporaneity of volcanic activity. Igneous layers and source rocks for clasts of the Baronick Formation have an enriched MORB chemistry and underwent albitization of calcic feldspar before erosion and conglomerate deposition. The Highway Suite forms a kilometre-scale body of gabbro and dolerite plugs and is interpreted as a slice of transitional continental oceanic crust. The chemistry of all igneous rocks suggests a continental rift environment and the associated sediments are consistent with such a setting. The Baronick Formation was locally intruded by sills of the Highway Suite; however, the main body of the Highway Suite was juxtaposed against the Baronick Formation during greenschist facies shearing before c. 551 Ma.

Type
EARTH SCIENCES
Copyright
Copyright © Antarctic Science Ltd 2007

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

Allibone, A.H. 1992. Low pressure/high temperature metamorphism of Koettlitz Group metasediments, Taylor Valley and upper Ferrar Glacier area, South Victoria Land, Antarctica. New Zealand Journal of Geology and Geophysics, 35, 115128.CrossRefGoogle Scholar
Bailey, D.K. 1983. The chemical and thermal evolution of rifts. Tectonophysics, 94, 585597.CrossRefGoogle Scholar
Baker, B.H. 1987. Outline of the geology of the Kenya rift alkaline province. In Fitton, J.G. & Upton, B.G.J., eds. Alkaline igneous rocks. Geological Society of London, Special Publication No 3, 293311.Google Scholar
Baker, B.H., Goles, G.G., Leeman, W.P. & Lindstrom, M.M. 1977. Geochemistry and petrogenesis of a basalt-benmoreite-trachyte suite from the southern part of the Gregory Rift, Kenya. Contributions to Mineralogy and Petrology, 64, 303332.CrossRefGoogle Scholar
Bamber, T. 1974. A series of magmatism related to the formation of spilite. In Amstutz, G.C., ed. Spilites and spilitic rocks. Berlin: Springer, 83112.CrossRefGoogle Scholar
Barberi, F., Santacroce, R. & Varent, J. 1974. Silicic peralkaline volcanic rocks of the Afar Depression (Ethiopia). Bulletin Volcanologique, 38, 755790.CrossRefGoogle Scholar
Barberi, F., Ferrara, G., Santacroce, R., Treuil, M. & Varet, J. 1975. A transitional basalt-pentellerite sequence of fractional crystallisation, the Boina Centre (Afar Rift, Ethiopia). Journal of Petrology, 16, 2256.CrossRefGoogle Scholar
Basaltic Volcanism Study Project. 1981. Basaltic volcanism on the terrestrial planets. New York: Pergamon Press, 1286 pp.Google Scholar
Bell, R.T. & Jefferson, C.W. 1987. A hypothesis for an Australian–Canadian connection in the Late Proterozoic and the birth of the Pacific Ocean. In Proceedings, Pacific Rim Congress 1987. Parkville, Victoria, Australian Institute of Mining and Metallurgy, 3950.Google Scholar
Blank, H.R., Cooper, R.A., Wheeler, R.H. & Willis, I.A.G. 1963. Geology of the Koettlitz-Blue Glacier region, southern Victoria Land, Antarctica. Transactions of the Proceedings of the Royal Society of New Zealand, 2, 79100.Google Scholar
Bond, G.C., Nickeson, P.A. & Kominz, M.A. 1984. Breakup of a supercontinent between 625 Ma and 555 Ma: new evidence and implications for continental histories. Earth and Planetary Science Letters, 70, 325345.CrossRefGoogle Scholar
Borg, S.G., DePaolo, D.J. & Smith, B.M. 1990. Isotopic structure and tectonics of the central Transantarctic Mountains. Journal of Geophysical Research, 95, 6647–67.CrossRefGoogle Scholar
Cook, Y.A. 1997. The Skelton Group and the Ross orogeny. PhD thesis, University of Otago, 502 pp.Google Scholar
Cook, Y.A. & Craw, D. 2001. Amalgamation of disparate crustal fragments in the Walcott Bay–Foster Glacier area, south Victoria Land, Antarctica. New Zealand Journal of Geology and Geophysics, 44, 403416.CrossRefGoogle Scholar
Cook, Y.A. & Craw, D. 2002. Neoproterozoic crustal slices in the Ross orogen, Skelton Glacier area, South Victoria Land, Antarctica. New Zealand Journal of Geology and Geophysics, 45, 133143.CrossRefGoogle Scholar
Crawford, A.J., Stevens, B.P.J. & Fanning, M. 1997. Geochemistry and tectonic setting of some Neoproterozoic and Early Cambrian volcanics in western New South Wales. Australian Journal of Earth Sciences, 44, 831852.CrossRefGoogle Scholar
Dalziel, I.W.D. 1991. Pacific margins of Laurentia and East Antarctica-Australia as a conjugate rift pair; evidence and implications for an Eocambrian supercontinent. Geology, 19, 598601.2.3.CO;2>CrossRefGoogle Scholar
Dalziel, I.W.D. 1992. Antarctica: a tale of two supercontinents? Annual Review Earth Planetary Sciences, 20, 501526.CrossRefGoogle Scholar
Davies, G.R. & McDonald, R. 1987. Crustal influences in the petrogenesis of the Naivasha basalt-comendite complex: combined trace element and Sr–Nd–Pb isotope constraints. Journal of Petrology, 28, 10091031.CrossRefGoogle Scholar
Deniel, C., Vidal, P., Coulon, C., Vellutini, P. & Piguet, P. 1994. Temporal evolution of mantle sources during continental rifting: the volcanism of Djibouti (Afar). Journal of Geophysical Research, 99, 28532869.CrossRefGoogle Scholar
Encarnacion, J. & Grunow, A. 1996. Changing magmatic and tectonic styles along the paleo-Pacific margin of Gondwana and the onset of early Paleozoic magmatism in Antarctica. Tectonics, 15, 13251341.CrossRefGoogle Scholar
Ewart, A. 1982. The mineralogy and petrology of Tertiary–Recent orogenic volcanic rocks: with special reference to the andesitic-basaltic compositional range. In Thorpe, R.S., ed. Andesites: orogenic andesites and related rocks. Chichester: John Wiley, 2687.Google Scholar
Findlay, R.H., Skinner, D.N.B. & Craw, D. 1984. Lithostratigraphy and structure of the Koettlitz Group, McMurdo Sound. New Zealand Journal of Geology and Geophysics, 27, 513536.CrossRefGoogle Scholar
Gibson, I.L. 1972. The chemistry and petrogenesis of a suite of pantellerites from the Ethiopian Rift. Journal of Petrology, 13, 3144.Google Scholar
Goodge, J.W., Myrow, P., Williams, I.S. & Bowring, S.A. 2002. Age and provenance of the Beardmore Group, Antarctica: constraints on Rodinia supercontinent breakup. Journal of Geology, 110, 393407.CrossRefGoogle Scholar
Gunn, B.M. & Warren, G. 1962. Geology of Victoria Land between Mawson and Mulock Glaciers, Antarctica. New Zealand Geological Survey Bulletin, No. 71, 175 pp.Google Scholar
Harris, C. 1983. The petrology of lavas and associated plutonic inclusions of Ascension Island. Journal of Petrology, 24, 424470.CrossRefGoogle Scholar
Hawksworth, C.J., van Calsteren, P.W., Rogers, N.W., Ellam, R. & Menzies, M.A. 1987. Isotope variations in recent volcanics: a trace element perspective. In Menzies, M.A. & Hawksworth, C.J., eds. Mantle metasomatism. London: Academic Press, 365389.Google Scholar
Hoffman, P.E. 1991. Did the breakout of Laurentia turn Gondwanaland inside-out? Science, 252, 14091412.CrossRefGoogle ScholarPubMed
Humphris, S.E., Thompson, G., Schilling, J.-G. & Kingsley, R.H. 1985. Petrological and geochemical variations along the Mid-Atlantic Ridge between 46°S and 32°S: influence of the Tristan da Cunha mantle plume. Geochimica et Cosmochimica Acta, 49, 14451464.CrossRefGoogle Scholar
Leat, P.T., Thompson, R.N., Morrison, M.A., Hendry, G.L. & Dickin, A.P. 1988. Compositionally-diverse Miocene–Recent rift-related magmatism in northwest Colorado: partial melting and mixing of mafic magmas from three different asthenospheric and lithospheric mantle sources. Journal of Petrology Special Lithosphere Issue, 351377.Google Scholar
Le Roex, A.P. 1987. Source regions of mid-ocean ridge basalts: evidence for enrichment processes. In Menzies, M.A. & Hawksworth, C.J., eds. Mantle metasomatism. London: Academic Press, 389423.Google Scholar
Le Roex, A.P., Dick, H.J.B., Reid, A.M., Frey, F.A., Erlank, A.J. & Hart, S.R. 1983. Petrology and geochemistry, mineralogy and petrogenesis of lavas erupted along the southwest Indian Ridge between the Bouvet triple junction and 11 degrees East. Journal of Petrology, 24, 267318.CrossRefGoogle Scholar
Lopatin, B.G. 1972. Basement complex of the McMurdo ‘oasis’, southern Victoria Land. In Adie, R.J., ed. Antarctic geology and geophysics. Oslo: Universitetsforlaget, 287292.Google Scholar
MacDonald, R., Davies, G.R., Bliss, C.M., Leat, P.T., Bailey, D.K. & Smith, R.L. 1987. Geochemistry of high silica peralkaline comendites, Naivasha, Kenya Rift Valley. Journal of Petrology, 28, 9791008.CrossRefGoogle Scholar
Middlemost, E.A.K. 1994. Naming materials in the magma/igneous rock system. Earth Science Reviews, 37, 215224.CrossRefGoogle Scholar
Millar, I.L. & Storey, B.C. 1995. Early Palaeozoic rather than Neoproterozoic volcanism and rifting within the Transantarctic Mountains. Journal of the Geological Society of London, 152, 417420.CrossRefGoogle Scholar
Miyashiro, A. 1978. Nature of alkalic volcanic rock series. Contributions to Mineralogy and Petrology, 66, 91104.CrossRefGoogle Scholar
Moores, E.M. 1991. Southwest U.S.–East Antarctic (SWEAT) connection: a hypothesis. Geology, 19, 425–28.2.3.CO;2>CrossRefGoogle Scholar
Pankhurst, R.J., Storey, B.C., Millar, I.L., Macdonald, D.I.M. & Vennum, W.R. 1988. Cambrian–Ordovician magmatism in the Thiel Mountains, Transantarctic Mountains, and implications for the Beardmore orogeny. Geology, 16, 246249.2.3.CO;2>CrossRefGoogle Scholar
Patwardhan, A.M. & Bhandari, A. 1974. Petrogenesis of Spilites occurring at Mandi, Himachal Pradesh, India. In Amstutz, G.C., ed. Spilites and spilitic rocks. Berlin: Springer, 175190.CrossRefGoogle Scholar
Rowell, A.J., Van Schmus, W.R., Storey, B.C., Fetter, A.H. & Evans, K.R. 2001. Latest Neoproterozoic to Mid-Cambrian age for the main deformation phases of the Transantarctic Mountains: new stratigraphic and isotopic constraints from the Pensacola Mountains, Antarctica. Journal of the Geological Society of London, 158, 295308.CrossRefGoogle Scholar
Rowell, A.J., Rees, M.N., Duebendorfer, E.M., Wallin, E.T., Van Schmus, W.R. & Smith, E.I. 1993. An active Neoproterozoic margin: evidence from the Skelton Glacier area, Transantarctic Mountains. Journal of the Geological Society of London, 150, 677682.CrossRefGoogle Scholar
Rowell, A.J., Gonzales, D.A., McKenna, L.W., Evans, K.R., Stump, E. & Van Schmus, W.R. 1997. Lower Paleozoic rocks in the Queen Maud Mountains, revised ages and significance. In Ricci, C.A., ed. The Antarctic region: geological evolution and processes. Siena: Terra Antartica Publications, 201207.Google Scholar
Skinner, D.N.B. 1982. Stratigraphy and structure of lower grade metasediments of Skelton Group, McMurdo Sound: does Teall Greywacke really exist? In Craddock, C., ed., Antarctic geoscience. Madison: University of Wisconsin Press, 555563.Google Scholar
Storey, B.C. 1993. The changing face of late Precambrian and early Palaeozoic reconstructions. Journal of the Geological Society of London, 150, 665668.CrossRefGoogle Scholar
Storey, B.C., Alabaster, T., Macdonald, D.I.M., Millar, I.L., Pankhurst, R.J. & Dalziel, I.W.D. 1992. Upper Proterozoic rift-related rocks in the Pensacola Mountains, Antarctica; precursors to supercontinent breakup? Tectonics, 11, 13921405.CrossRefGoogle Scholar
Stump, E. 1992. The Ross orogen of the Transantarctic Mountains in light of the Laurentia-Gondwana split. GSA Today, 2, 2531.Google Scholar
Sun, S.-S. 1980. Lead isotopic study of young volcanic rocks from mid-ocean ridges, ocean islands and island arcs. Philosophical Transactions of the Royal Society, A297, 409445.Google Scholar
Sun, S.-S. & McDonough, W.F. 1989. Chemical and isotopic systematics of ocean basalts: Implications for mantle composition and processes. In Saunders, A.D. & Norry, M.J. eds. Magmatism in the ocean basins. Geological Society of London Special Publication, 42, 313–46.Google Scholar
Thompson, R.N., Morrison, M.A., Hendry, G.L. & Parry, S.J. 1984. An assessment of the relative roles of a crust and mantle in magma genesis: an elemental approach. Philosophical Transactions of the Royal Society, A310, 549590.Google Scholar
Walcott, C.R. & Craw, D. 1993. Post-emplacement deformation of plutons and their metasedimentary host, Mt. Dromedary area, south Victoria Land, Antarctica. New Zealand Journal of Geology and Geophysics, 36, 487496.CrossRefGoogle Scholar
Weaver, S.D., Sceal, J.S.C. & Gibson, I.L. 1972. Trace element data relevant to the origin of trachytic and pantelleritic lavas in the East African Rift System. Contributions to Mineralogy and Petrology, 36, 181194.CrossRefGoogle Scholar
Winchester, J.A. & Floyd, P.A. 1977. Geochemical discrimination of different magma series and their differentiation products using immobile elements. Chemical Geology, 20, 325343.CrossRefGoogle Scholar