Hostname: page-component-848d4c4894-2pzkn Total loading time: 0 Render date: 2024-05-22T17:49:35.674Z Has data issue: false hasContentIssue false

Geodynamic implications of the Cenozoic stress field on Seymour Island, West Antarctica

Published online by Cambridge University Press:  21 January 2008

A. Maestro*
Affiliation:
Instituto Geológico y Minero de España, Ríos Rosas, 23, 28003 Madrid, Spain Departamento de Geología y Geoquímica, Facultad de Ciencias, Universidad Autónoma de Madrid, 28049 MadridSpain
J. López-Martínez
Affiliation:
Departamento de Geología y Geoquímica, Facultad de Ciencias, Universidad Autónoma de Madrid, 28049 MadridSpain
F. Bohoyo
Affiliation:
Instituto Geológico y Minero de España, Ríos Rosas, 23, 28003 Madrid, Spain
M. Montes
Affiliation:
Instituto Geológico y Minero de España, Ríos Rosas, 23, 28003 Madrid, Spain
F. Nozal
Affiliation:
Instituto Geológico y Minero de España, Ríos Rosas, 23, 28003 Madrid, Spain
S. Santillana
Affiliation:
Instituto Antártico Argentino, Cerrito 1248, 1010 Buenos Aires, Argentina
S. Marenssi
Affiliation:
Instituto Antártico Argentino, Cerrito 1248, 1010 Buenos Aires, Argentina

Abstract

Palaeostress inferred from brittle mesostructures in Seymour (Marambio) Island indicates a Cenozoic to Recent origin for an extensional stress field, with only local compressional stress states. Minimum horizontal stress (σ3) orientations are scattered about two main NE–SW and NW–SE modes suggesting that two stress sources have been responsible for the dominant minimum horizontal stress directions in the north-western Weddell Sea. Extensional structures within a broad-scale compressional stress field can be linked to both the decrease in relative stress magnitudes from active margins to intraplate regions and the rifting processes that occurred in the northern Weddell Sea. Stress states with NW–SE trending σ3 are compatible with back-arc extension along the eastern Antarctic Peninsula. We interpret this as due to the opening of the Larsen Basin during upper Cretaceous to Eocene and to the spreading, from Pliocene to present, of the Bransfield Basin (western Antarctic Peninsula), both due to former Phoenix Plate subduction under the Antarctic Plate. NE–SW σ3 orientations could be expressions of continental fragmentation of the northern Antarctic Peninsula controlling eastwards drifting of the South Orkney microcontinent and other submerged continental blocks of the southern Scotia Sea.

Type
Earth Sciences
Copyright
Copyright © Antarctic Science Ltd 2008

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

Acosta, J. & Uchupi, E. 1996. Transtensional tectonics along the South Scotia Ridge, Antarctica. Tectonophysics, 267, 3156.CrossRefGoogle Scholar
Ashcroft, W.A. 1972. Crustal structure of the South Shetland Islands and Bransfield Strait. British Antarctic Survey Scientific Report, No. 66, 43 pp.Google Scholar
Barber, P.L., Barker, P.F. & Pankhurst, R.J. 1991. Dredged rocks from Powell basin and the South Orkney microcontinent. In Thomson, M.R.A., Crame, J.A. & Thomson, J.W., eds. Geological evolution of the Antarctica. Cambridge: Cambridge University Press, 361367.Google Scholar
Barker, D.H.N. & Austin, J.A. 1994. Crustal diapirism in Bransfield Strait, West Antarctica: evidence for distributed extension in marginal-basin formation. Geology, 22, 657660.2.3.CO;2>CrossRefGoogle Scholar
Barker, P.F. 1982. The Cenozoic subduction history of the Pacific margin of the Antartic Peninsula: ridge crest-trench interactions. Journal of the Geological Society of London, 139, 787801.CrossRefGoogle Scholar
Barker, P.F. & Lawver, L.A. 1988. South American–Antarctic plate motion over the past 50 Myr, and the evolution of the South American–Antarctic ridge. Geophysical Journal, 94, 377386.CrossRefGoogle Scholar
Barker, P.F., Dalziel, I.W.D. & Storey, B.C. 1991. Tectonic development of the Scotia Arc Region. In Tingey, R.J., ed. Antarctic geology. Oxford: Oxford University Press, 215248.Google Scholar
Barker, P.F., Kennett, J.P. & Shipboard Scientific Party. 1988. Weddell Sea palaeoceanography: preliminary results of ODP leg 113. Palaeogeography, Palaeoclimatology, Palaeoecology, 67, 75102.CrossRefGoogle Scholar
Bohoyo, F. 2004. Fragmentación continental y desarrollo de cuencas oceánicas en el sector medridional del Arco de Scotia, Antártida. PhD thesis, University of Granada, Spain, 236 pp. [Unpublished.]Google Scholar
Bohoyo, F., Galindo-Zaldívar, J., Maldonado, A., Schreider, A.A. & Suriñach, E. 2002. Basin development subsequent to ridge-trench collision: the Jane Basin, Antarctica. Marine Geophysical Researches, 23, 413421.CrossRefGoogle Scholar
Bohoyo, F., Galindo-Zaldívar, J., Jabaloy, A., Maldonado, A., Rodríguez-Fernández, J., Schreider, A.A. & Suriñach, E. 2007. Extensional deformation and development of deep basins associated with the sinistral transcurrent fault zone of the Scotia–Antarctic plate boundary. In Cunningham, D. & Mann, P., eds. Tectonics of strike-slip restraining and releasing bends. Geological Society of London Special Publication, 290.Google Scholar
Bott, M.H.P. 1959. The mechanics of oblique slip faulting. Geological Magazine, 96, 109117.CrossRefGoogle Scholar
Coren, F., Ceccone, G., Lodolo, E., Zanolla, C., Zitellini, N., Bonazzi, C. & Centonze, J. 1997. Morphology, seismic structure and tectonic development of the Powell Basin, Antarctica. Journal of the Geological Society, London, 154, 849862.CrossRefGoogle Scholar
DelValle, R.A., Elliot, D.H. & Macdonald, D.I.M. 1992. Sedimentary basins of the east flank of the Antarctic Peninsula: proposed nomenclature. Antarctic Science, 4, 477478.CrossRefGoogle Scholar
Dietrich, R., Dach, R., Engelhardt, G., Heck, B., Kutterer, H., Lindner, K., Mayer, M., Menge, F., Mikolaiski, H.W., Niemeier, W., Pohl, M., Salbach, H., Schenke, H.W., Schöne, T., Seeber, G. & Soltau, G. 1996. The SCAR 95 GPS campaign: objectives, data analysis and final solution. In Dietrich, R., ed. The Geodetic Antarctic Project GAP95 German Contributions to the SCAR 95 Epoch Campaign. Deutsche Geodätische Kommission bei der Bayerischen Akademie der Wissenschaften, Reihe B, Heft Nr 304, S914.Google Scholar
Dietrich, R., Rülke, A., Ihde, J., Lindner, K., Miller, H., Niemeier, W., Schenke, H.-W. & Seeber, G. 2004. Plate kinematics and deformation status of the Antarctic Peninsula based on GPS. Global and Planetary Change, 42, 313321.CrossRefGoogle Scholar
Doglioni, C. 1995. Geological remarks on the relationships between extension and convergent geodynamics settings. Tectonophysics, 252, 253267.CrossRefGoogle Scholar
Dziewonski, A.M., Chou, T.A. & Woodhouse, J.H. 1981. Determination of earthquake source parameters from waveform data for studies of global and regional seismicity. Journal of Geophysical Research, 86, 28252852.CrossRefGoogle Scholar
Eagles, G. & Livermore, R.A. 2002. Opening history of Powell Basin, Antarctic Peninsula. Marine Geology, 185, 195202.CrossRefGoogle Scholar
Eagles, G., Livermore, R. & Morris, P. 2006. Small basins in the Scotia Sea: the Eocene Drake Passage gateway. Earth and Planetary Science Letters, 242, 343353.CrossRefGoogle Scholar
Elliot, D.H. & Trautman, T.A. 1982. Lower Tertiary strata on Seymour Island, Antarctic Peninsula. In Craddock, C., ed. Antarctic geoscience. Madison, WI: University of Wisconsin, 287297.Google Scholar
Elliot, D.H., Askin, R.A., Kyte, F.T. & Zinsmeister, W.J. 1994. Iridium and dinocysts at the Cretaceous–Tertiary boundary on Seymour Island, Antarctica: implications for the K–T event. Geology, 22, 675678.2.3.CO;2>CrossRefGoogle Scholar
Feldmann, R.M. & Woodburne, M.O.eds. 1988. Geology and Paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America Memoir, No. 169, 566 pp.Google Scholar
Galindo-Zaldívar, J., Bohoyo, F., Maldonado, A., Schreider, A., Suriñach, E. & Vázquez, J.T. 2006b. Propagating rift during the opening of a small oceanic basin: the Protector Basin (Scotia Arc, Antarctica). Earth and Planetary Science Letters, 241, 398412.CrossRefGoogle Scholar
Galindo-Zaldívar, J., Gamboa, L., Maldonado, A., Nakao, S. & Bochu, Y. 2006a. Tectonic development of the Bransfield Basin and its prolongation to the South Scotia Ridge, northern Antarctic Peninsula. Marine Geology, 206, 267282.CrossRefGoogle Scholar
Galindo-Zaldívar, J., Jabaloy, A., Maldonado, A. & Sanz de Galdeano, C. 1996. Continental fragmentation along the South Scotia Ridge transcurrent plate boundary (Antarctic Peninsula). Tectonophysics, 259, 275301.CrossRefGoogle Scholar
Gambôa, L.A.P. & Maldonado, P.R. 1990. Geophysical investigations in the Bransfield Strait and in the Bellingshausen Sea, Antarctica. In John, B.S., ed. Antarctica as an exploration frontier - hydrocarbon potential, geology and hazards. American Association of Petroleum Geologists Studies in Geology, No. 31, 127142.Google Scholar
Garrett, S.W. & Storey, B.C. 1987. Lithospheric extension on the Antarctic Peninsula during Cenozoic subduction. In Coward, M.P., Dewey, J.F. & Hancock, P.L., eds. Continental extension tectonics. Geological Society of London, Special Publications, 28, 419431.Google Scholar
Gazdzicki, A. & Webb, P.N. 1996. Foraminifera from the Pecten Conglomerate (Pliocene) of Cockburn Island. Antarctic Peninsula. Paleontologia Polonica, 55, 147174.Google Scholar
Ghidella, M.E., Yáñez, G. & LaBrecque, J.L. 2002. Revised tectonic implications for the magnetic anomalies of the western Weddell Sea. Tectonophysics, 347, 6586.CrossRefGoogle Scholar
Giner-Robles, J.L., Gonzalez-Casado, J.M., Gumiel, P., Martin-Velazquez, S. & Garcia-Cuevas, C. 2003. A kinematic model of the Scotia plate (SW Atlantic Ocean). Journal of South American Earth Sciences, 16, 179191.CrossRefGoogle Scholar
Gónzalez-Casado, J.M., Giner, J.L. & López-Martínez, J. 2000. Bransfield Basin, Antarctic Peninsula: not a normal backarc basin. Geology, 28, 10431046.2.0.CO;2>CrossRefGoogle Scholar
González-Casado, J.M., López-Martínez, J. & Durán, J.J. 1999. Active tectonics and morphostructure at the nothern margin of the Central Bransfield Basin, Hurd Peninsula, Livingston Island (South Shetland Islands). Antarctic Science, 11, 323331.CrossRefGoogle Scholar
González-Ferrán, O. 1983. The Larsen Rift: An active extension fracture in West Antarctica. In Oliver, R.L., James, P.R. & Jago, J.B., eds. Antarctic earth science. Camberra: Australian Academy of Science, 344346.Google Scholar
Guterch, A., Grad, M., Janik, T. & Perchuc, E. 1991. Tectonophysical models of the crust between the Antarctic Peninsula and the South Shetland trench. In Thomson, M.R.A., Crame, J.A. & Thomson, J.W., eds. Geological evolution of Antarctica. Cambridge: Cambridge University Press, 409504.Google Scholar
Hathway, B. 2000. Continental rift to back-arc basin: Jurassic–Cretaceous stratigraphical and structural evolution of the Larsen Basins, Antarctic Peninsula. Journal of the Geological Society, London, 157, 417432.CrossRefGoogle Scholar
Hampel, A. & Hetzel, R. 2006. Response of normal faults to glacial-interglacial fluctuations of ice and water masses on Earth's surface. Journal of Geophysical Research, 11, doi: 10.1029/2005JB004124.Google Scholar
Hancock, P.L. & Engelder, T. 1989. Neotectonic joints. Bulletin Geological Society of America, 101, 11971208.2.3.CO;2>CrossRefGoogle Scholar
Hobbs, G.J. 1968. The geology of South Shetland Islands. IV. The geology of Livingstone Island. British Antarctic Survey Scientific Report, No. 47, 34 pp.Google Scholar
Hyden, G. & Tanner, W.G. 1981. Late Paleozoic–Early Mesozoic fore-arc basin sedimentary rocks at the Pacific Margin in western Antarctica. Geologische Rundschau, 70, 529541.CrossRefGoogle Scholar
Ineson, J.R. 1989. Coarse-grained submarine fan and slope apron deposits in a Cretaceous back-arc basin. Sedimentology, 36, 793819.CrossRefGoogle Scholar
Ivany, L.C., VanSimaey, S., Domack, E.W. & Samson, S.D. 2006. Evidence for an earliest Oligocene ice-sheet on the Antarctic Peninsula. Geology, 34, 377380.CrossRefGoogle Scholar
Jeffers, J.D., Anderson, J.B. & Lawver, L.A. 1991. Evolution of the Brandsfield Basin, Antarctic Peninsula. In Thomson, M.R.A., Crame, J.A. & Thomson, J.W., eds. Geological evolution of Antarctica. Cambridge: Cambridge University Press, 481485.Google Scholar
Johnston, A.C. 1987. Suppresion of earthquakes by large continental ice sheets. Nature, 330, 467469.CrossRefGoogle Scholar
King, E., Leitchenkov, G., Galindo-Zaldívar, J. & Maldonado, A. 1994. Basement distribution in Powell Basin: understanding the tectonic controls on sedimentation. Terra Antartica, 1, 307308.Google Scholar
King, E.C. & Barker, P.F. 1988. The margins of the South Orkney microcontinent. Journal of the Geological Society, London, 145, 317331.CrossRefGoogle Scholar
Kovacs, L.C., Morris, P., Brozena, J. & Tikku, A. 2002. Seafloor spreading in the Weddell Sea from magnetic and gravity data. Tectonophysics, 347, 4364.CrossRefGoogle Scholar
Larter, R.D. & Barker, P.F. 1991. Effects of ridge crest-trench interaction on Antarctic–Phoenix spreading: forces on a young subducting plate. Journal of Geophysical Research, 96, 1958319607.CrossRefGoogle Scholar
Lawver, L.A., Williams, T. & Sloan, B. 1994. Seismic stratigraphy and heat flow of Powell Basin. Terra Antarctica, 1, 309310.Google Scholar
Lawver, L.A., Sloan, B.J., Barker, D.H.N., Ghidella, M.E., VonHerzen, R.F., Keller, R.A., Klinkhamer, G.P. & Chin, C.S. 1996. Distributed active extension in Bransfield Basin, Antarctic Peninsula: evidence from multibeam bathymetry. GSA Today, 6(11), 16.Google Scholar
Lee, J.C. & Angelier, J. 1994. Paleostress trajectory maps based on the results of local determinations: the “lissage” program. Computers & Geosciences, 20, 161191.CrossRefGoogle Scholar
Lee, D.K., Jin, Y.K., Kim, Y. & Nam, S.H. 2000. Seismicity and tectonics around the northern Antarctic Peninsula from King Sejong station data. Antarctic Science, 12, 196204.CrossRefGoogle Scholar
Lisle, R.J., Orife, T. & Arlegui, L. 2001. A stress inversion method requiring only faults slip sense. Journal of Geophysical Research, 106, 22812289.CrossRefGoogle Scholar
Livermore, R.A. & Woollett, R.W. 1993. Seafloor spreading in the Weddell Sea and southwest Atlantic since the Late Cretaceous. Earth and Planetary Science Letters, 117, 475495.CrossRefGoogle Scholar
Livermore, R.A., Nankivell, A., Eagles, G. & Morris, P. 2005. Paleogene opening of Drake Passage. Earth and Planetary Science Letters, 236, 459470.CrossRefGoogle Scholar
Macdonald, D.I.M., Barker, P.F., Garrett, S.W., Ineson, J.R., Pirrie, D., Storey, B.C., Whitham, A.C., Kinghorn, R.R.F. & Marshall, J.E.A. 1988. A preliminary assessment of the hydrocarbon potential of the Larsen Basin, Antarctica. Marine and Petroleum Geology, 5, 3453.CrossRefGoogle Scholar
Maestro, A., Somoza, L., Rey, J., Martínez-Frías, J. & López-Martínez, J. 2007. Active tectonics, fault patterns, and stress field of Deception Island: a response to oblique convergence between the Pacific and Antarctic plates. Journal of South American Earth Sciences, 23, 256268.CrossRefGoogle Scholar
Maldonado, A., Larter, R.D. & Aldaya, F. 1994. Forearc tectonic evolution of the South Shetland Margin, Antarctic Peninsula. Tectonics, 13, 13451370.CrossRefGoogle Scholar
Maldonado, A., Balanyá, J.C., Barnolas, A., Galindo-Zaldívar, J., Hernández, J., Jabaloy, A., Livermore, R.A., Martínez, J.M., Rodríguez-Fernández, J., Sanz deGaldeano, C., Somoza, L., Suriñach, E. & Viseras, C. 2000. Tectonics of an extinct ridge-transform intersection, Drake Passage (Antarctica). Marine Geophysical Researches, 21, 4368.CrossRefGoogle Scholar
Marenssi, S.A., Santillana, S.N. & Rinaldi, C.A. 1998. Stratigraphy of La Meseta Formation (Eocene), Marambio Island, Antarctica. In Casadío, S., ed. Paleógeno de America del Sur y de la Península Antártica. Revista de la Asociación Paleontológica Argentina, Publicación Especial, 5, 137146.Google Scholar
Orife, T., Arlegui, L. & Lisle, R.J. 2002. DIPSLIP: a QuickBasic stress inversion program for analysisng sets of faults without slip lineations. Computers & Geosciences, 28, 775781.CrossRefGoogle Scholar
Parra, J.C., GonzálezFerrán, O. & Bannister, J. 1984. Aeromagnetic survey over the South Shetland Islands, Bransfield Strait and part of the Antarctic Peninsula. Revista Geológica de Chile, 23, 320.Google Scholar
Pelayo, A. & Wiens, D. 1989. Seismotectonics and relative plate motions in the Scotia Sea region. Journal of Geophysical Research, 96, 72937320.CrossRefGoogle Scholar
Pirrie, D. 1994. Petrology and provenance of the Marambio Group, Vega Island, Antarctica. Antarctic Science, 6, 517527.CrossRefGoogle Scholar
Pirrie, D., Crame, J.A., Lomas, S.A. & Riding, J.B. 1997. Late Cretaceous stratigraphy of the Admiralty Sound region, James Ross Basin, Antarctica. Cretaceous Research, 18, 109137.CrossRefGoogle Scholar
Porebski, S.J. 2000. Shelf-valley compound fill produced by fault subsidence and eustatic sea level changes, Eocene La Meseta Formation, Seymour Island, Antarctica. Geology, 28, 147150.2.0.CO;2>CrossRefGoogle Scholar
Rinaldi, C.A., Massabie, A., Morelli, J., Rosenmann, L.H. & Del Valle, R.A. 1978. Geología de la isla Vicecomodoro Marambio, Antártida. Instituto Antártico Argentino, Contribución, 217, 137.Google Scholar
Roach, P.J. 1978. The nature of back-arc extension in Brandsfield Strait. Royal Astronomical Society Geophysical Journal, 53, 165.Google Scholar
Robertson, S.D., Wiens, D.A., Shore, P.J., Vera, E. & Dorman, L.M. 2003. Seismicity and tectonics of the South Shetland Islands and Bransfield Strait from a regional broadband seismograph deployment. Journal of Geophysical Research, 108, doi: 10.1029/2003JB002416.Google Scholar
Rodríguez-Fernández, J., Balanya, J.C., Galindo-Zaldívar, J. & Maldonado, A. 1997. Tectonic evolution and growth patterns of a restricted ocean basin: the Powell Basin (northeastern Antarctic Peninsula). Geodinamica Acta, 10, 159174.CrossRefGoogle Scholar
Sadler, P.M. 1988. Geometry and stratification of uppermost Cretaceous and Paleogene units of Seymour Island, northern Antarctic Peninsula. In Feldman, R.M. & Woodburne, M.O., eds. Geology and paleontology of Seymour Island, Antarctic Peninsula. Geological Society of America, No. 169, 303320.CrossRefGoogle Scholar
Simón, J.L. 1986. Análisis of gradual change in stress regime (example from the eastern Iberian Chain, Spain). Tectonophysics, 124, 3753.Google Scholar
Smith, W.H.F. & Sandwell, D.Y. 1997. Global seafloor topography from satellite altimetry and ship depth sounding. Science, 277, 19571962.CrossRefGoogle Scholar
Thomas, C., Livermore, R. & Pollitz, F. 2003. Motion of the Scotia Sea plates. Geophysical Journal International, 155, 789804.CrossRefGoogle Scholar
Wallace, R.E. 1951. Geometry of shearing stress and relation to faulting. Journal of Geology, 59, 118130.CrossRefGoogle Scholar
Willan, R.C.R. & Kelley, S.P. 1999. Mafic dike swarms in the South Shetland Islands volcanic arc: unravelling multiepisodic magmatism related to subduction and continental rifting. Journal of Geophysical Research, 104, 2305123068.CrossRefGoogle Scholar
Willan, R.C.R., Pankhurst, R.J. & Hervé, F. 1994. A probable Early Triassic age for the Miers Bluff Formation, Livingston Island, South Shetlands Islands. Antarctic Science, 6, 401408.CrossRefGoogle Scholar
Zinsmeister, W.J. & De Vries, T. 1983. Quaternary glacial marine deposits on Seymour Island. Antarctic Journal of the United States, 18(5), 6465.Google Scholar