Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-r5zm4 Total loading time: 0 Render date: 2024-06-19T21:44:45.120Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  05 March 2016

Michal Nemčok
Affiliation:
University of Utah
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Rifts and Passive Margins
Structural Architecture, Thermal Regimes, and Petroleum Systems
, pp. 507 - 593
Publisher: Cambridge University Press
Print publication year: 2016

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

Aajdour, M., Hssaïda, T., Fedan, B., El Morabet, A., Oumalch, F., and Fakhi, S. (1999). Contribution à l’étude du Jurassique de la marge marocaine, Apports de l’analyse palynologique dans l’évolution géodynamique. 1st Colloquium on Moroccan Jurassic, Morocco, pp. 1–2.Google Scholar
Abe, N. (2001). Petrochemistry of serpentinized peridotite from the Iberia abyssal plain (ODP Leg 173): its character intermediate between suboceanic and sub-continental upper mantle. Geological Society of London Special Publications, 187, 143–59.CrossRefGoogle Scholar
Abu El Karamat, S. and Fouda, H. G. (1990). Cross fault pattern and its impact on clysmic faults in the Gulf of Suez. Egyptian 10th Exploration and Production Conference, p. 14.Google Scholar
Achauer, U. and the KRISP Teleseismic Working Group (1994). New ideas on the Kenya Rift based on the inversion of the combined dataset of the 1985 and 1989/90 seismic tomography experiments. Tectonophysics, 236, 305–30.CrossRefGoogle Scholar
Adám, Á. (1976). Results of deep electromagnetic investigations (in Pannonian Basin). Geoelectric and Geothermal Studies (East-Central Europe, Soviet Asia). In KAPG Geophysical Monograph, ed. Adám, Á. and Kiadó, Akadémiai, Budapest, pp. 547–60.Google Scholar
Adám, Á. (1997). Mantle diapir – mantle plumes in the Pannonian Basin (investigations on the asthenosphere in Hungary from the sixties till the present). In Physics and Geophysics with Special Historical Case Studies, ed. Schröder, W.. Bremen-Roennebeck, Germany, pp. 7187.Google Scholar
Adams, J. T. and Dart, C. (1998). The appearance of potential sealing faults on borehole images. In Faulting, Fault Sealing and Fluid Flow in Hydrocarbon Reservoirs, ed. Jones, G., Fisher, Q. J., and Knipe, R. J.. Geological Society of London Special Publications, 147, 7186.Google Scholar
Aguiar, L. A. M., Jinno, K., and Neto, J. B. M. (1986). Transpressional tectonics in the Foz do Amazonas Basin. Internal report, Petrobras (in Portuguese). In Tectonic Evolution of Brazilian Equatorial Continental Margin Basins, De Azevedo, R. P.. (1991). PhD thesis, Imperial College, London, pp. 1403.Google Scholar
Aguilera, R. and van Poollen, H. K. (1979). Occurrence of fractured reservoirs. Oil and Gas Journal, 77, 70–4.Google Scholar
Ahlbrandt, T. S. (2002). Assessment of global oil, gas, and NGL resources based on the total petroleum system concept. American Association of Petroleum Geologists Bulletin, 86, 1847.Google Scholar
Ahmadi, Z., Sawyers, M., Kenyon-Roberts, S., Stanworth, B., Kugler, K., Kristensen, J., and Fugelli, E. (2002). Chapter 14. Paleocene. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. D. Evans, C. Graham, A. Armour, and P. Bathurst. Geological Society of London, pp. 549–97.Google Scholar
Ahmadi, Z., Sawyers, M., Kenyon-Roberts, S., Stanworth, B., Kugler, K., Kristensen, J., and Fugelli, E. (2003). Chapter 14. Paleocene. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. Geological Society of London, pp. 235–59.Google Scholar
Ahnert, F. (1970). Functional relationships between denudation, relief, and uplift in large, mid-latitude drainage basins. American Journal of Science, 268, 243–63.CrossRefGoogle Scholar
Aizenshtat, Z., Miloslavski, I., Ashengrau, D., and Oren, A. (1999). Hypersaline depositional environments and their relation to oil generation. In Microbiology and Biogeochemistry of Hypersaline Environments: The Microbiology of Extreme and Unusual Environments, ed. Oren, A.. CRC Press, Boca Raton, FL, pp. 89108.Google Scholar
Al-Tabaa, A. and Wood, D. M. (1987). Some measurements of permeability of kaolin. Geotechnique, 37, 499503.CrossRefGoogle Scholar
Albrechtsen, T., Andersen, S. J., Dons, T., Engstrom, F., Jorgensen, O., Sorensen, F. W., and Maersk Olie og Gas AS (2001). Developing non-structurally trapped oil in North Sea chalk. Conference paper. Society of Petroleum Engineers Annual Technical Conference and Exhibition, New Orleans, Louisiana, September 30–October 3, 2001.CrossRefGoogle Scholar
Alexander, L. L. and Handschy, J. W. (1998). Fluid flow in a faulted reservoir system: fault trap analysis for the Block 330 field in Eugene Island, south addition, offshore Louisiana. American Association of Petroleum Geologists Bulletin, 82, 387411.Google Scholar
Allan, U. S. (1989). A model for the migration and entrapment of hydrocarbons within faulted structures. American Association of Petroleum Geologists Bulletin, 75, 803–21.Google Scholar
Allemand, P. (1988). Approche experimentale de la mécanique du rifting continental. Memoires et Documents du Centre Armoricain D´etude Structurale des Socles, Rennes, 38, 1192.Google Scholar
Allemand, P. and Brun, J. P. (1991). Width of continental rifts and rheological layering of the lithosphere. Tectonophysics, 188, 63–9.CrossRefGoogle Scholar
Allemand, P., Brun, J. P., Davy, P., and van den Driessche, J. (1989). Symetrie et asymetrie des rifts et mecanismes d’amincissement de la lithosphere. Symmetry and asymmetry of rifts and mechanism of lithospheric thinning. Bulletin de la Societe Geologique de France, Huitieme Serie, 5, 445–51.Google Scholar
Allen, P. A. and Allen, J. R. (1990). Basin Analysis: Principles and Applications. Blackwell Scientific Publications, Oxford, p. 451.Google Scholar
Allen, P. A. and Densmore, A. L. (2000). Sediment flux from an uplifting fault block. Basin Research, 12, 367–80.CrossRefGoogle Scholar
Allis, R. G., Armstrong, P. A., and Funnell, R. H. (1995). Implications of high heat flow anomaly around New Plymouth. New Zealand Journal of Geology and Geophysics, 38, 121–30.CrossRefGoogle Scholar
Allix, P., Grosdidier, E., Jardine, S., Legoux, O., and Popoff, M. (1981). Decouverte d’Aptien superieur a Albien inferieur date par microfossiles dans la serie detritique cretacee du fosse de la Benoue (Nigeria). Upper Aptian and lower Albian deposits, dated from microfossils, in the Cretaceous detrital series of the Benue Valley, Nigeria. Comptes-Rendus des Seances de l’Academie des Sciences, Serie 2: Mecanique-Physique, Chimie, Sciences de l’Univers, Sciences de la Terre, 292, 1291–4.Google Scholar
Almeida, F. F. M., Dal Re Carneiro, C., Machado, D. D. L. Jr., and Dehira, L. K. (1988). Magmatismo Pos-Paleozoico no nordeste oriental do Brasil. Post-Paleozoic magmatism of the eastern part of northeastern Brazil. Revista Brasileira de Geociencias, 18, 451–62.Google Scholar
Al’Subbary, A. K., Nichols, G. N., Bosence, D. W. J., and Al-Kadasi, M. (1998). Pre-rift doming, peneplanation or subsidence in the southern Red Sea?: Evidence from the Medj-Zir Formation (Tawilah Group) of western Yemen. In Sedimentation and Tectonics in Rift Basins: Red Sea–Gulf of Aden Case, ed. B. Purser and D. Bosence. Chapman and Hall, London, pp. 119–34.Google Scholar
Alt, J. C. (1995). Subseafloor processes in mid-oceanic ridge hydrothermal systems. American Geophysical Monograph, 91, 85114.Google Scholar
Alvarez, F., Virieux, J., and Le Pichon, X. (1984). Thermal consequences of lithosphere extension: the initial stretching phase. Geophysical Journal of the Royal Astronomic Society, 78, 389411.CrossRefGoogle Scholar
Ambrose, G. (2000). The geology and hydrocarbon habitat of the Sarir sandstone, SE Sirt Basin, Libya. Journal of Petroleum Geology, 23, 165–92.CrossRefGoogle Scholar
Ambrose, W. A., Jones, R. H., Wawrzyniec, T. F., Fouad, K., Dutton, S. P., Jennette, D. C., Elshayeb, T., Sanchez-Barreda, L., Solis, H., Meneses-Rocha, J. L., Lugo, J. E., Aguilera, L., Berlanga, J., Miranda, L., Ruiz Morales, J., and Rojas, R. (2002). Upper Miocene and Pliocene shallow-marine and deepwater, gas-producing systems in the Macuspana Basin, southeastern Mexico. In Gulf Coast Association of Geological Societies and Gulf Coast Section SEPM, Technical Papers and Abstracts, ed. Dutton, S. P., Ruppel, S. C., and Hentz, T. F.. Transactions – Gulf Coast Association of Geological Societies, 52, 312.Google Scholar
Aminzadeh, F. and Connolly, D. (2002). Looking for gas chimneys and faults. American Association of Petroleum Geologists Explorer, 23, 20–1.Google Scholar
Aminzadeh, F. and Connolly, D. (2007). Distinguishing hydrocarbon migration pathways in seismic data. Bulletin of the South Texas Geological Society, 47, 1516.Google Scholar
Amrani, A., Lewan, M. D., and Aizenshtat, Z. (2005). Stable sulfur isotope partitioning during simulated petroleum formation as determined by hydrous pyrolysis of Ghareb Limestone, Israel. Geochimica et Cosmochimica Acta, 69, 5317–31.CrossRefGoogle Scholar
Anderholm, S. K. (1983). Hydrogeology of the Socorro and La Jencia Basins, Socorro County, New Mexico. Field Conference Guidebook, New Mexico Geological Society, 34, 303–10.Google Scholar
Anderholm, S. K. (1987). Hydrogeology of the Socorro and La Jencia Basins, Socorro County, New Mexico. U.S. Geological Survey Water Resources Investigation Report, 844342.Google Scholar
Anders, D. E. and Gerrild, P. M. (1984). Hydrocarbon generation in lacustrine rocks of Tertiary age, Uinta Basin, Utah – organic carbon, pyrolysis yield, and light hydrocarbons. In Hydrocarbon Source Rocks of the Greater Rocky Mountain Region, ed. Woodward, J., Meissner, F. F., and Clayton, J. L.. Rocky Mountain Association of Geologists, Denver, pp. 513–29.Google Scholar
Anders, D. E., Palacas, J. G., and Johnson, R. C. (1992). Thermal maturity of rocks and hydrocarbon deposits, Uinta Basin, Utah. In Hydrocarbon and Mineral Resources of the Uinta Basin, Utah and Colorado, ed. Fouch, T. D., Nuccio, V. F., and Chidsey, T. C.. Utah Geological Association Guidebook, 20, 5376.Google Scholar
Anders, M. H. and Schlische, R. W. (1994). Overlapping faults, intrabasin highs, and the growth of normal faults. Journal of Geology, 102, 165–79.CrossRefGoogle Scholar
Anders, M. H., Spiegelman, M., Rodgers, D. W., and Hagstrum, J. T. (1993). The growth of fault-bounded tilt blocks. Tectonics, 12, 1451–9.CrossRefGoogle Scholar
Anderson, D. L. (1994). The sublithospheric mantle as the source of continental flood basalts: the case against the continental lithosphere and plume head reservoirs. Earth and Planetary Science Letters, 123, 269–80.CrossRefGoogle Scholar
Anderson, D. L. (2000). The thermal state of the upper mantle: no role for mantle plumes. Geophysical Research Letters, 27, 3623–6.CrossRefGoogle Scholar
Anderson, D. L., Zhang, Y. S., and Tanimoto, T. (1992). Plume heads, continental lithosphere, flood basalts and tomography. In Magmatism and the Causes of Continental Break-up, ed. Storey, B. C., Alabaster, T., and Pankhurst, R. J.. Geological Society of London Special Publications, 68, 99124.Google Scholar
Anderson, R. N., Flemings, P. B., Losh, S., Austin, J., and Woodhams, R. (1994a). Gulf of Mexico growth fault drilled, seen as oil, gas migration pathway. Oil and Gas Journal, 94, 97104.Google Scholar
Anderson, R. N., Flemings, P. B., Losh, S., Whelan, J., Billeaud, J., Austin, J., and Woodhams, R. (1994b). The Pathfinder drilling program into a major growth fault in Eugene Island 330, Gulf of Mexico: implications for behavior of hydrocarbon migration pathways. CD-ROM prepared under Department of Energy contract DE-FC22-93BC14961.Google Scholar
Anderson, R. N., Hobart, M. A., and Langseth, M. G. (1979). Geothermal convection through oceanic crust and sediments in the Indian Ocean. Science, 204, 828–32.CrossRefGoogle ScholarPubMed
Andrews-Speed, C. P., Oxburgh, E. R., and Cooper, B. A. (1984). Temperatures and depth-dependent heat flow in western North Sea. American Association of Petroleum Geologists Bulletin, 68, 1764–81.Google Scholar
Angelier, J. (1985). Extension and rifting: the Zeit region, Gulf of Suez. Journal of Structural Geology, 7, 605–12.CrossRefGoogle Scholar
Angelier, J. and Colletta, B. (1983). Tension fractures and extensional tectonics. Nature, 301, 4951.CrossRefGoogle Scholar
Angevine, C. L., Turcotte, D. L., and Furnish, M. D. (1982). Pressure solution lithification as a mechanism for the stick-slip behavior of faults. Tectonophysics, 1, 151–60.Google Scholar
Anhaeusser, C. R., Mason, R., Vilhoen, M. J., and Viljoen, R. P. (1969). A reappraisal of some aspects of Precambrian shield geology. Geological Society of America Bulletin, 80, 2175–200.CrossRefGoogle Scholar
Anisimov, P. D., Dixon, J. E., and Langmuir, C. H. (2004). A hydrous melting and fractionation model for mid-ocean ridge basalts: application to the Mid- Atlantic Ridge near Azores. Geochemistry, Geophysics, Geosystems, 5.Google Scholar
ANP (2013). http://www.anp.gov.br/ accessed on March 1, 2013. Lithostratigraphic chart of the Santos Basin accessed at: http://www.anp.gov.br/brasil-rounds/round4/round4/workshop/restrito/portugues/Santos_port.pdf. Lithostratigraphic chart of the Campos Basin accessed at: http://www.anp.gov.br/brasil-rounds/round4/round4/workshop/restrito/portugues/Campos_port.PDF.Google Scholar
Antonellini, M. and Aydin, A. (1994). Effect of faulting on fluid flow in porous sandstones: petrophysical properties. American Association of Petroleum Geologists Bulletin, 78, 355–77.Google Scholar
Antonellini, M. and Mollema, P. N. (2000). A natural analog for a fractured and faulted reservoir in dolomite: Triassic Sella Group, northern Italy. American Association of Petroleum Geologists Bulletin, 84, 314–44.Google Scholar
AOA Geophysics (2001). Boujdour permit area, Morocco. Phase 1 investigation AOA Geophysics, Houston.Google Scholar
Arabasz, W. J., Pechmann, J. C., and Brown, E. D. (1992). Observational seismology and the evaluation of earthquake hazards and risk in the Wasatch Front area, Utah. In Assessment of Regional Earthquake Hazards and Risk along the Wasatch Front, Utah, ed. Gori, P. L. and Hays, W. W.. U.S. Geological Survey Professional Paper, 1500-A-J, pp. 136.Google Scholar
Arad, A. and Bein, A. (1986). Saline versus freshwater contribution to the thermal waters of the northern Jordan Rift Valley. Journal of Hydrology, 83, 4966.CrossRefGoogle Scholar
Arch, J. and Maltman, A. J. (1990). Anisotropic permeability and tortuosity in deformed wet sediments. Journal of Geophysical Research, 95, 9035–47.CrossRefGoogle Scholar
Arch, J., Maltman, A. J., and Knipe, R. J. (1988). Shear-zone geometries in experimentally deformed clays: the influence of water content, strain rate and primary fabric. Journal of Structural Geology, 10, 91–9.CrossRefGoogle Scholar
Areshev, E. G., Dong, T. L., San, N. T., and Shnip, O. A. (1992). Reservoirs in fractured basement on the continental shelf of Southern Vietnam. Journal of Petroleum Geology, 15, 451–64.CrossRefGoogle Scholar
Argent, J. D., Stewart, S. A., Green, P. F., and Underhill, J. R. (2002). Heterogeneous exhumation in the Inner Moray Firth, UK North Sea: constraints from new AFTA and seismic data. Journal of the Geological Society of London, 159, 715–29.CrossRefGoogle Scholar
Argent, J. D., Stewart, S. A., and Underhill, J. R. (2000). Controls on the Lower Cretaceous Punt Sandstone member, a massive deep-water clastic deposystem, Inner Moray Firth, UK North Sea. Petroleum Geoscience, 6, 275–85.CrossRefGoogle Scholar
Armijo, R., Tapponnier, P., Mercier, J. L., and Han, T. L. (1986). Quaternary extension in southern Tibet: field observations and tectonic implications. Journal of Geophysical Research, 91, 13803–72.CrossRefGoogle Scholar
Armstrong, L. A., Ten Have, A., and Johnson, H. D. (1987). The geology of the Gannet fields, Central North Sea, UK Sector. In Petroleum Geology of North West Europe, ed. Brooks, J. and Glennie, K. W.. Graham and Trotman, London, pp. 533–48.Google Scholar
Armstrong, P. A. and Chapman, D. S. (1998). Beyond surface heat flow: an example from a tectonically active sedimentary basin. Geology, 26, 183–6.2.3.CO;2>CrossRefGoogle Scholar
Armstrong, P. A., Ehlers, T. A., Chapman, D. S., Farley, K. A., and Kamp, P. J. J. (2003). Exhumation of the central Wasatch Mountains, Utah: 1. Patterns and timing of exhumation deduced from low-temperature thermochronology data. Journal of Geophysical Research, 108.CrossRefGoogle Scholar
Armstrong, P. A., Kamp, P. J. J., Allis, R. G., and Chapman, D. S. (1997). Thermal effects of intrusion below the Taranaki Basin (New Zealand): evidence from combined apatite fission track age and vitrinite reflectance data. Basin Research, 9, 151–69.CrossRefGoogle Scholar
Arndt, N. T. and Christensen, U. (1992). The role of the lithospheric mantle in continental flood volcanism: thermal and geochemical constraints. Journal of Geophysical Research, 97, 10967–81.CrossRefGoogle Scholar
Arrowsmith, J. R. (1995). Coupled Tectonic Deformation and Geomorphic Degradation along the San Andreas Fault System. Stanford University, Stanford, California, p. 347.Google Scholar
Artemieva, I. M. and Mooney, W. D. (2001). Thermal thickness and evolution of Precambrian lithosphere: a global study. Journal of Geophysical Research, 106, 16387–414.CrossRefGoogle Scholar
Artemieva, I. M. and Mooney, W. D. (2002). On the relations between cratonic lithosphere thickness, plate motions, and basal drag. Tectonophysics, 358, 211–31.CrossRefGoogle Scholar
Artemjev, M. E. and Artyushkov, E. V. (1971). Structure and isostasy of the Baikal Rift and the mechanism of rifting. Journal of Geophysical Research, 76, 1197–211.CrossRefGoogle Scholar
Artyushkov, E. V., Morner, N. A., and Tarling, D. H. (2000). The cause of loss of lithospheric rigidity in areas far from plate tectonic activity. Geophysical Journal International, 143, 752–76.CrossRefGoogle Scholar
Asimow, P. D., Dixon, J. E., and Langmuir, C. H. (2004). A hydrous melting and fractionation model for mid-ocean ridge basalts: application to the Mid-Atlantic Ridge near Azores. Geochem. Geophys. Geosyst., 5.CrossRefGoogle Scholar
Aslanian, D., Moulin, M., Olivet, J.-L., Unternehr, P., Matias, L., Bache, F., Rabineau, M., Nouzé, H., Klingelhoefer, F., Contrucii, I., and Labails, C. (2009). Brazilian and African passive margins of the Central Segment of the South Atlantic Ocean: kinematic constraints. Tectonophysic, 468, 98112.CrossRefGoogle Scholar
Athy, L. F. (1930). Density, porosity and compaction of sedimentary rocks. American Association of Petroleum Geologists, 14, 124.Google Scholar
Atwood, G. I. and Forman, M. J. (1959). Nature and growth of southern Louisiana salt domes and its effect on petroleum accumulation. American Association of Petroleum Geologists Bulletin, 43, 2595–622.Google Scholar
Audet, D. M. and McConell, J. D. C. (1992). Establishing resolution limits for tectonic subsidence curves by forward basin modeling. Marine and Petroleum Geology, 11, 400–11.Google Scholar
Aulbach, S., Rudnick, R. L., and McDonough, W. F. (2008). Li-Sr-Nd isotope signatures of the plume and cratonic lithospheric mantle beneath the margin of the rifted Tanzanian craton (Labait). Contribution to Mineralogy and Petrology, 155, 7992.CrossRefGoogle Scholar
Austin, J. A., Stoffa, P. L., Phillips, J. D., Oh, J., Sawyer, D. S., Purdy, G. M., Reiter, E., and Makris, J. (1990). Crustal structure of the southeast Georgia Embayment–Carolina Trough: preliminary results of a composite seismic image of a continental suture (?) and a volcanic passive margin. Geology, 18, 1023–7.2.3.CO;2>CrossRefGoogle Scholar
Avdulov, M. V. (1970). Stroyeniye zemnoy kory Krymskogo poluostrova po rezul’tatam geofizicheskikh issledovaniy. Structure of the Earth’s crust of the Crimea Peninsula according to geophysical survey results. Kompleksnyye issledovaniya Chernomorskoy vpadiny. Akademia Nauk SSSR, Mezhduved. Geofiz. Kom., Rezul’t. Issled. Mezhdunar. Geofiz. Proyekt., Moscow.Google Scholar
Aydin, A. (1977). Small faults formed as deformation bands in sandstone. In Proceedings of Conference II: Experimental Studies of Rock Friction with Application to Earthquake Prediction, ed. Evernden, J. F.. U.S. Geological Survey, Office of Earthquake Studies, Menlo Park, California, pp. 617–53.Google Scholar
Aydin, A. (1978a). Faulting in sandstone. Doctoral dissertation, Stanford University.Google Scholar
Aydin, A. (1978b). Small faults formed as deformation bands in sandstone. Pure and Applied Geophysics, 116, 913–30.Google Scholar
Aydin, A., Borja, R. I., and Eichhubl, P. (2006). Geological and mathematical framework for failure modes in granular rock. Journal of Structural Geology, 28, 8398.CrossRefGoogle Scholar
Aydin, A. and Johnson, A. M. (1978). Development of faults as zones of deformation bands and as slip surfaces in sandstone. Pure and Applied Geophysics, 116, 931–42.Google Scholar
Aydin, A. and Johnson, A. M. (1983). Analysis of faulting in porous sandstones. Journal of Structural Geology, 5, 1935.CrossRefGoogle Scholar
Aydin, A. and Nur, A. (1985). The types and role of stepovers in strike-slip tectonics. In Strike-Slip Deformation, Basin Formation and Sedimentation, ed. Biddle, K. T. and Christie-Blick, N.. Special Publication – Society of Economic Paleontologists and Mineralogists, pp. 3544.CrossRefGoogle Scholar
Babuška, V. and Plomerová, J. (1988). Subcrustal continental lithosphere: a model of its thickness and anisotropic structure. Physics of the Earth and Planetary Interiors, 51, 130–2.CrossRefGoogle Scholar
Bachler, D., Kohl, T., and Rybach, L. (2003). Impact of graben-parallel faults on hydrothermal convection – Rhine Graben case study. Physics and Chemistry of the Earth, 28, 431–41.Google Scholar
Bachu, S. (1985). Influence of lithology and fluid flow on the temperature distribution in a sedimentary basin: a case study from the Cold Lake area, Alberta, Canada. Tectonophysics, 120, 257–84.CrossRefGoogle Scholar
Bachu, S. (1991). On the effective thermal and hydraulic conductivity of binary heterogeneous sediments. Tectonophysics, 190, 299314.CrossRefGoogle Scholar
Bachu, S., Ramon, J. C., Villegas, M. E., and Underschultz, J. R. (1995). Geothermal regime and thermal history of the Llanos Basin, Colombia. American Association of Petroleum Geologists Bulletin, 79, 116–29.Google Scholar
Bacoccoli, G. (1991). O futuro da exploração de petróleo na Bacia de Sergipe-Alagoas: Relatório Interno Petrobrás, Depex, Rio de Janeiro. Soil structure and its effects on hydraulic conductivity. Soil Science, 114, 417–22.Google Scholar
Bailey, C. C., Price, I., and Spencer, A. M. (1981). The Ula oil field, Block 7/12, Norway. In The Sedimentation of the North Sea Reservoir Rocks. Norwegian Symposium of Exploration, Report of the Norwegian Petroleum Society, Oslo, 18, pp. 126.Google Scholar
Bailey, N. J. L., Burwood, R., and Harriman, G. E. (1990). Application of pyrolysate carbon isotope and biomarker technology to organofacies definition and oil correlation problems in North Sea basins. Organic Geochemistry, 16, 1157–72.CrossRefGoogle Scholar
Bailey, R. C. (1990). Trapping of aqueous fluids in the deep crust. Geophysical Research Letters, 17, 1129–32.CrossRefGoogle Scholar
Baird, R. A. (1986). Maturation and source rock evaluation of Kimmeridge Clay, Norwegian North Sea. American Association of Petroleum Geologists Bulletin, 7, 111.Google Scholar
Baker, A. (1933). Geology and oil possibilities of the Moab district, Grand and San Juan Counties. Geological Survey Bulletin, Utah, p. 841.Google Scholar
Baker, J., Chazot, G., Menzies, M. A., and Thirlwall, M. (2002). Lithospheric mantle beneath Arabia; a Pan-African protolith modified by the Afar and older plumes, rather than a source for continental flood volcanism? In Volcanic Rift Margins, ed. Menzies, M. A., Klemperer, S. L., Ebinger, C., and Baker, J.. Geological Society of America Special Paper, 362, pp. 6580.Google Scholar
Baker, J., MacPhearson, C. J., Menzies, M. A., Thirwall, M. F., Al-Kadasi, M., and Mattey, D. P. (2000). Resolving crustal and mantle contributions to continental flood volcanism, Yemen: constraints from mineral oxygen isotope data. Journal of Petrology, 41(12), 1805–20.CrossRefGoogle Scholar
Baker, J., Snee, L., and Menzies, M. A. (1996a). A brief Oligocene period of flood volcanism in Yemen: implications for the duration and rate of continental flood volcanism at the Afro-Arabian triple junction. Earth and Planetary Science Letters, 138, 3956.CrossRefGoogle Scholar
Baker, J., Thirwall, M. F., and Menzies, M. A. (1996b). Sr-Nd_Pb isotopic and trace element evidence for crustal contamination of a mantle plume: Oligocene flood volcanism in western Yemen. Geochemica et Cosmochimica Acta, 60, 2559–81.CrossRefGoogle Scholar
Bakun, W. H. and Lindh, A. G. (1985). The Parkfield, California earthquake prediction experiment. Science, 229, 619–24.CrossRefGoogle ScholarPubMed
Balkwill, H. R. and McMillan, N. J. (1990). Geology of the Labrador Shelf, Baffin Bay and Davis Strait: Part 1. Mesozoic–Cenozoic geology of the Labrador Shelf. In Geology of the Continental Margin of Eastern Canada, ed. Keen, M. J. and Williams, G. L.. Geological Survey of Canada, 2, pp. 295324.Google Scholar
Balling, N. (1985). Thermal structure of the lithosphere beneath the Norwegian-Danish Basin and the southern Baltic shield: a major transition zone. Terra cognita, 5, pp. 377–8.Google Scholar
Bally, A. W. (1982). Musings over sedimentary basin evolution. In Evolution of Sedimentary Basins, ed. Kent, P., Botts, M. H. P., McKenzie, D. P., and Williams, C. A.. Philosophical Transactions of the Royal Society of London, Mathematical and Physical Sciences, 305, pp. 325–38.Google Scholar
Bally, A. W., Bernoulli, D., Davis, G. A., and Montadert, L. (1981). Listric normal faults. Oceanologica Acta, 4, 87101.Google Scholar
Bally, A. W., Gordy, P. L., and Stewart, G. A. (1966). Structure, seismic data, and orogenic evolution of southern Canadian Rocky Mountains. Bulletin of Canadian Petroleum Geology, 14, 337–81.Google Scholar
Banda, E., Gallart, J., Garcia-Duenas, V., Danobeitia, J. J., and Makris, J. (1993). Lateral variation of the crust in the Iberian Peninsula: new evidence from the Betic Cordillera. Tectonophysics, 221, 5366.CrossRefGoogle Scholar
Bankey, V. A., Cevas, D., Daniels, A. A., Finn, I., Hermandez and Project Members (2002). Digital data grids for the magnetic anomaly map of North America: [Nova Scotia aeromagnetic field map]. U.S. Geological Survey, Open-File Report, 02-0414.Google Scholar
Barbarand, J., Carter, A., Wood, I., and Hurford, T. (2003). Compositional and structural control of fission-track annealing in apatite. Chemical Geology, 198, 107–37.CrossRefGoogle Scholar
Barnard, P. C., Collins, A. G., and Cooper, B. S. (1981). Identification and distribution of kerogen facies in a source rock horizon: examples from the North Sea Basin. In Organic Maturation Studies and Fossil Fuel Exploration, ed. Brook, J.. Academic Press, London, pp. 271–82.Google Scholar
Barnard, P. C. and Cooper, B. S. (1981). Oils and source rocks in the North Sea area. In Petroleum Geology of the Continental Shelf of Northwest Europe, ed. Illing, L. V. and Hobson, G. D.. Heyden and Son, London, pp. 169–75.Google Scholar
Barnouin-Jha, K., Parmentier, E. M., and Sparks, D. W. (1997). Buoyant mantle upwelling and crustal production at oceanic spreading centers: on-axis segmentation and off-axis melting. Journal of Geophysical Research, 102, 11979–89.CrossRefGoogle Scholar
Barr, D. (1987). Lithospheric stretching, detached normal faulting and footwall uplift. Geological Society Special Publications, 28, 7594.CrossRefGoogle Scholar
Barroll, M. W. and Reiter, M. (1990). Analysis of the Socorro hydrogeothermal system: central New Mexico. Journal of Geophysical Research, 95, 21949–63.CrossRefGoogle Scholar
Barton, A. J. and White, R. S. (1997a). Crustal structure of Edoras Bank continental margin and mantle thermal anomalies beneath the North Atlantic. Journal of Geophysical Research, 102, 3109–29.CrossRefGoogle Scholar
Barton, A. J. and White, R. S. (1997b). Volcanism on the Rockall continental margin. Journal of the Geological Society, 154, 531–6.CrossRefGoogle Scholar
Barton, C. A. and Moss, D. (1988). Analysis of macroscopic fractures in the Cajon Pass scientific drillhole: over the interval 1829–2115 meters. Geophysical Research Letters, 15, 1013–16.CrossRefGoogle Scholar
Barton, C. A. and Zoback, M. D. (1992). Self-similar distribution and properties of macroscopic fractures at depth in crystalline rock on the Cajon Pass scientific drill hole. Journal of Geophysical Research, 97, 5181–200.CrossRefGoogle Scholar
Barton, C. A., Zoback, M. D., and Moos, D. (1995). Fluid flow along potentially active faults in crystalline rock. Geology, 23, 683–6.2.3.CO;2>CrossRefGoogle Scholar
Basak, P. (1972). Soil structure and its effects on hydraulic conductivity. Soil Science, 114, 417–22.CrossRefGoogle Scholar
Basile, C., Mascle, J., Benkhelil, J., and Bouillin, J. P. (1998). Geodynamic evolution of the Côte d’Ivoire–Ghana transform margin: an overview of Leg 159 results. In Proc. ODP, Sci. Results, 159, 101–110, ed. J. Mascle, G. P. Lohmann and M. Moullade. Available at: http://www.odp.tamu.edu/publications/159_SR/CHAPTERS/CHAP_11.PDF accessed December 5, 2001.Google Scholar
Baskin, D. K. and Peters, K. E. (1992). Early generation characteristics of a sulfur-rich Monterey kerogen. American Association of Petroleum Geologists Bulletin, 76, 113.Google Scholar
Baskin, V. N., Dmitrieva, L. S., Makhov, G. T., Ponomareva, T. V., and Polyakov, V. A. (1979). Effect of the duration of storage of standard specimens on their chemical composition. Industrial Laboratory (USSR), 45, 159–60.Google Scholar
Bassi, G., Keen, C. E., and Potter, P. (2003). Contrasting styles of rifting: models and examples from the eastern Canadian margin. Tectonics, 12, 639–55.Google Scholar
Bauer, K., Neben, S., Schrekenberger, B., Emmerman, R., Hinz, K., Jokat, W., Schulze, A., Trumbull, R. B., and Weber, K. (2000). Deep structure of the Namibia continental margin as derived from integrated geophysical studies. Journal of Geophysical Research, 105, 25829–53.CrossRefGoogle Scholar
Baum, M. S., Withjack, M. O., and Schlische, R. W. (2003). Controls of structural geometries associated with rift-basin inversion. 2003 AAPG Annual Convention with SEPM, Extended Abstracts, Tulsa, Oklahoma, 12, 10.Google Scholar
Baum, M. S., Withjack, M. O., and Schlische, R. W. (2008). The ins and outs of buttress folds: examples from the Fundy rift basin, Nova Scotia and New Brunswick, Canada. In Central Atlantic Conjugate Margins, ed. Brown, D. E.. Dalhousie University, Halifax, Nova Scotia, pp. 5361.Google Scholar
Baxter, K., Cooper, G. T., Hill, K. C., and O’Brien, G. W. (1999). Late Jurassic subsidence and passive margin evolution in the Vulcan Sub-basin, north-west Australia: constraints from basin modelling. Basin Research, 11, 97111.CrossRefGoogle Scholar
Bear, J. (1972). Dynamics of Fluids in Porous Media. Elsevier, New York.Google Scholar
Beaumont, C., Ellis, S., and Pfiffner, A. (1999). Dynamics of sediment subduction-accretion at convergent margins: short-term modes, long-term deformation, and tectonic implications. Journal of Geophysical Research, 104, 17573–601.CrossRefGoogle Scholar
Beaumont, C., Keen, C. E., and Boutilier, R. (1982). On the evolution of rifted continental margins: comparison of models and observations for the Nova Scotian margin. Geophysical Journal of the Royal Astronomical Society, 70, 667715.CrossRefGoogle Scholar
Beck, W. C. and Chapin, C. E. (1994). Structural and tectonic evolution of the Joyita Hills, central New Mexico: implications of basement control on Rio Grande Rift. Special Paper – Geological Society of America, 291, 187205.CrossRefGoogle Scholar
Becker, H., Horan, M. F., Walker, R. J., Gao, S., Lorand, J. P., and Rudnick, R. L. (2006). Highly siderophile element composition of the Earth’s primitive upper mantle: constraints from new data on peridotite massifs and xenoliths. Geochimica et Cosmochimica Acta, 70, 4528–50.CrossRefGoogle Scholar
Bederke, E. (1966). The development of European rifts. Geological Survey of Canada, 66, 213–19.Google Scholar
Bedrosian, P. A., Unsworth, M. J., and Egbert, G. D. (2002). Magnetotelluric imaging of the creeping segment of the San Andreas Fault near Hollister. Geophysical Research Letters, 29, 1-1-1-4.CrossRefGoogle Scholar
Bedrosian, P. A., Unsworth, M. J., Egbert, G. D., and Thurber, C. H. (2004). Geophysical images of the creeping segment of the San Andreas Fault: implications for the role of crustal fluids in the earthquake process. Tectonophysics, 385, 137–58.CrossRefGoogle Scholar
Beekman, F. (1994). Tectonic modeling of thick-skinned compressional intraplate deformation. PhD thesis, Vrije University, Amsterdam, p. 152.Google Scholar
Beer, J. A., Allmendinger, R. W., Figueroa, D. E., and Jordan, T. E. (1990). Seismic stratigraphy of a Neogene piggyback basin, Argentina. American Association of Petroleum Geologists Bulletin, 74, 1183–202.Google Scholar
Begg, G. C., Griffin, W. L., Natapov, L. M., O’Reilly, S. Y., Grand, S. P., O’Neill, C. J., Hronsky, J. M. A., Poudjom Djomani, Y., Swain, C. J., Deen, T., and Bowden, P. (2009). The lithospheric architecture of Africa: seismic tomography, mantle petrology, and tectonic evolution. Geosphere, 5, 2350.CrossRefGoogle Scholar
Beherens, W. E. (1988). Geology of a continental slope oil seep, Gulf of Mexico. American Association of Petroleum Geologists Bulletin, 72, 105–14.Google Scholar
Behn, M. D. and Lin, J. (2000). Segmentation in gravity and magnetic anomalies along the U.S. East Coast passive margin: implications for incipient structure of the oceanic lithosphere. Journal of Geophysical Research, 105, 25769–90.CrossRefGoogle Scholar
Behrendt, J. C., Hamilton, R. M., Ackermann, H. D., and Henry, V. J. (1981). Cenozoic faulting in the vicinity of the Charleston, South Carolina, 1886 earthquake. Geology, 9, 117–22.2.0.CO;2>CrossRefGoogle Scholar
Beitler, B., Chan, M. A., and Parry, W. T. (2003). Bleaching of Jurassic Navajo Sandstone on Colorado Plateau Laramide highs: evidence of exhumed hydrocarbon supergiants? Geology, 31, 1041–4.CrossRefGoogle Scholar
Bekins, B. A., McCaffrey, A. M., and Dreiss, S. J. (1994). The influence of kinetics on the smectite to illite transition in the Barbados accretionary prism. Journal of Geophysical Research, 99, 18145–58.CrossRefGoogle Scholar
Belgasem, B. A. (1990). An evaluation of an oil-bearing granite reservoir from well logs. The Log Analyst, 32, 404.Google Scholar
Belgasem, B. A., Barlai, Z., and Rez, F. (1991). Interpretation of well logs in the basement oil-bearing reservoir, Nafoora-Augile Field, Libya. The Log Analyst, 32, 42–3.Google Scholar
Bellani, S., Brogi, A., Lazzarotto, A., Liotta, D., and Ranalli, G. (2004). Heat flow, deep temperatures and extensional structures in the Larderello geothermal field (Italy): constraints on geothermal fluid flow. Journal of Volcanology and Geothermal Research, 132, 1529.CrossRefGoogle Scholar
Belousov, V. V. (1988). Structure and evolution of the Earth’s crust and upper mantle of the Black Sea. Bolletino di Geofisica Teorica ed Applicata, 3, 109–96.Google Scholar
Ben-Avraham, Z., Ginzburg, A., and Yuval, Z. (1981). Seismic reflection and refraction investigations of Lake Kinneret-central Jordan Valley, Israel. Tectonophysics, 80, 165–81.CrossRefGoogle Scholar
Ben-Avraham, Z., Hanel, R., and Villinger, H. (1978). Heat flow through the Dead sea rift. Marine Geology, 28, 253–67.CrossRefGoogle Scholar
Ben-Avraham, Z., Hartnady, C. J. H., and Kitchin, K. A. (1997). Structure and tectonics of the Agulhas-Falkland fracture zone. Tectonophysics, 282, 8398.CrossRefGoogle Scholar
Ben-Avraham, Z. and Zoback, M. D. (1992). Transform-normal extension and asymmetric basins; an alternative to pull-apart models. Geology, 20, 423–6.2.3.CO;2>CrossRefGoogle Scholar
Bender, A. A. (1987). O comportamento termomecânico do Terciário da Bacia do Pará-Maranháo. University Ouro Preto.Google Scholar
Benes, V. and Davy, P. (1996). Modes of continental lithospheric extension: experimental verification of strain localization processes. Tectonophysics, 254, 6987.CrossRefGoogle Scholar
Benkhelil, J. (1988). Structure et evolution geodynamique du bassin intracontinental de la Benoue (Nigeria). Structure and geodynamic evolution of the intracontinental Benue Basin, Nigeria. Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine, 12, 29128.Google Scholar
Benkhelil, J., Guiraud, M., Mascle, J., Basile, C., Bouillin, J.-P., Mascle, G., and Cousin, M. (1996). Structural record of the Africa/Brazil sliding in the Cretaceous sediments of the Cote d’Ivoire-Ghana transform margin. Comptes rendus de l´Academie des sciences, Série 2, 323, 7390.Google Scholar
Bennett, R. A., Wernicke, B. P., and Davis, J. L. (1998). Continuous GPS measurements of contemporary deformation across the northern Basin and Range Province. Geophysical Research Letters, 25, 563–6.CrossRefGoogle Scholar
Benson, R. N. (2002). Age estimates of the seaward-dipping volcanic wedge, earliest oceanic crust, and earliest drift-stage sediments along the North American Atlantic continental margin. Geophysical Monograph, 136, 6175.Google Scholar
Berckhemer, H., Baier, B., Bartelsen, H., Behle, A., Burkhart, H. et al. (1975). Deep seismic sounding in the Afar region and on the highland of Ethiopia. In Afar Depression of Ethiopia, vol. 1, ed. A. Pilger and A. Roesler. E. Schweizer Verlagsbuchhandl, Suttgart.Google Scholar
Berg, R. R. (1975). Capillary pressure in stratigraphic traps. American Association of Petroleum Geologists Bulletin, 59, 939–56.Google Scholar
Berg, R. R. and Gangi, A. F. (1999). Primary migration by oil-generation microfracturing in low-permeability source rocks: application to the Austin Chalk, Texas. American Association of Petroleum Geologists Bulletin, 83, 727–56.Google Scholar
Berger, M. and Roberts, A. M. (1999). The Zeta structure: a footwall degradation complex formed by gravity sliding on the western margin of the Tampen Spur, Northern North Sea. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, The Geological Society of London, pp. 107–16.Google Scholar
Bergslien, D. (2002). Balder and Jotun – Two sides of the same coin? A comparison of two Tertiary oil fields in the Norwegian North Sea. Petroleum Geoscience, 8, 349–63.CrossRefGoogle Scholar
Berkowitz, N. (1997). Fossil Hydrocarbons: Chemistry and Technology. Academic Press, San Diego, California.Google Scholar
Berman, A. (2008). Three super-giant fields discovered offshore Brazil. World Oil Magazine, 229. http://www.worldoil.com/Magazine/MAGAZINE_DETAIL.asp?ART_ID=3450&MONTH_YEAR=Feb-2008.Google Scholar
Bertotti, G. (1991). Early Mesozoic extension and Alpine shortening in the western Southern Alps: the geology of the area between Lugano and Menaggio (Lombardy, northern Italy). Memorie di Scienze Geologiche di Padova, 43, 17123.Google Scholar
Bertotti, G., Gouiza, M., Foeken, J., and Andriessen, P. (2010). Late Jurassic–Early Cretaceous Tectonics and Exhumation Onshore Morocco: implications for Terrigenous Sand Reservoirs in the Offshore of NW-Africa. Adapted from poster presentation at AAPG Annual Convention and Exhibition, New Orleans, Louisiana, April 11–14, 2010.Google Scholar
Bertotti, G., Siletto, G. B., and Spalla, M. I. (1993). Deformation and metamorphism associated with crustal rifting: the Permian to Liassic evolution of the Lake Lugano–Lake Como area (Southern Alps). Tectonophysics, 226, 271–84.CrossRefGoogle Scholar
Bertotti, G. and ter Voorde, M. (1994). Thermal effects of normal faulting during rifted basin formation, 2, The Lugano-Val Grande normal fault and the role of pre-existing thermal anomalies. Tectonophysics, 240, 145–57.CrossRefGoogle Scholar
Bertram, G. T. and Milton, N. J. (1989). Reconstructing basin evolution from sedimentary thickness: the importance of palaeobathymetric control, with reference to the North Sea. Basin Research, 1, 247–57.Google Scholar
Bertrand, R. and Achab, A. (1990). Equivalents between the reflectance of vitrinite, zooclasts (chitinozoans, graptolites and scolenodonts) and the color alteration of palynomorphs (spores and acritarchs). Palynology, 15, 13280.Google Scholar
Beslier, M. O., Cornen, G., Agrinier, P., Feraud, G., and Girardeau, J. (1996). Structure and evolution of a passive continental margin: main results of ODP Leg 149 in the ocean–continent transition of the Iberia abyssal plain. In First Eurocolloquium of the Ocean Drilling Program, ed. Ristedt, H.. Oldenburg, Federal Republic of Germany, pp. 21–2.Google Scholar
Best, P. et al. (1985). Ghana Project – Mid Term Report. In Tectonic Evolution of Brazilian Equatorial Continental Margin Basins, ed. De Azevedo, R. P.. PhD thesis, Imperial College, London, p. 455.Google Scholar
Bethke, C. M. (1985). A numerical model of compaction-driven groundwater flow and heat transfer and its application to the paleohydrology of intracratonic sedimentary basins. Journal of Geophysical Research, 90, 6817–28.CrossRefGoogle Scholar
Bethke, C. M. and Corbet, T. F. (1985). Linear and nonlinear solutions for one-dimensional compaction flow in sedimentary basins. Water Resources Research, 24, 461–7.Google Scholar
Bethke, C. M. and Marshak, S. (1990). Brine migrations across North America – the plate tectonics of groundwater. Annual Reviews of Earth and Planetary Sciences, 18, 287315.CrossRefGoogle Scholar
Betts, P. G., Giles, D., and Lister, G. S. (2003). Tectonic environment of shale-hosted massive sulfide Pb-Zn-Ag deposits of Proterozoic northeastern Australia. Economic Geology, 98, 557–76.CrossRefGoogle Scholar
Beydoun, Z. R. (1997). Introduction to the revised Mesozoic stratigraphy and nomenclature for Yemen. Marine and Petroleum Geology, 14, 617–29.CrossRefGoogle Scholar
Beydoun, Z. R., As-Saruri, M. A. L., El-Nakhal, H., Al-Ganad, I. N., Baraba, R. S., Nani, A. A. S. O., and Al-Aawah, M. H. (1998). International Lexicon of Stratigraphy, v. III. Asia, Republic of Yemen, International Union of Geological Sciences Publication, 34, 245.Google Scholar
Biddle, K. T. and Wielchowski, C. C. (1994). Hydrocarbon traps. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists, 60, 219–35.Google Scholar
Bielik, M., Blizkovský, M., Burda, M., Fusan, O., Huebner, M., Herrmann, H., Novotný, A., Suk, M., Tomek, C., and Vyskočil, V. (1994). Density models of the earth’s crust along seismic profiles. In Crustal Structure of the Bohemian Massif and the West Carpathians, ed. Burda, M., Suk, M., Šefara, J., and Pliva, G.. Academy of Sciences Czech Republic, Geophysical Institute, CZ-14131 Prague 4, Czechoslovakia CSK, pp. 177–87.Google Scholar
Bielik, M., Šefara, J., Bezák, V., and Kubeš, P. (1995). Deep-seated models of the Western Carpathians. Proceedings of the 1st Slovak Geophysical Conference, GISAS, pp. 7–11.Google Scholar
Biggar, N. E. and Adams, J. A. (1987). Dates derived from Quaternary strata in the vicinity of Canyonlands National Park. In Geology of Cataract Canyon and Vicinity, ed. Campbel, J. A.. Four Corners Geological Society, 10th Field Conference Guidebook, pp. 127–36.Google Scholar
Bijwaard, H. and Spakman, W. (1999). Tomographic evidence for a narrow whole mantle plume below Iceland. Earth and Planetary Science Letters, 166, 121–6.CrossRefGoogle Scholar
Biles, N. E., Hannan, A. E., Jamieson, G. A., and Bain, J. E. (2000). Geologic overview of the NE Mississippi fan and shelf to West Florida terrace region, offshore Gulf of Mexico. In Transactions – Gulf Coast Association of Geological Societies, ed. Ragsdale, J. A. and Rosen, N. C.. Gulf Coast Association of Geological Societies and Gulf Coast Section SEPM, 50th Annual Meeting, 50, 129–36.Google Scholar
Bill, M., O’Dogherty, L., Guex, J., Baumgartner, P. O., and Masson, H. (2001). Radiolarite ages in Alpine-Mediterranean ophiolites: constraints on the oceanic spreading and the Tethys-Atlantic connection. Geological Society of America Bulletin, 113, 129–43.2.0.CO;2>CrossRefGoogle Scholar
Billiris, H., Paradissis, D., Veis, G., England, P., Featherstone, W., Parsons, B., Cross, P., Rands, P., Rayson, M., Sellers, P., Ashkenazi, V., Davison, M., Jackson, J., and Ambraseys, N. (1991). Geodetic determination of tectonic deformation in central Greece from 1900 to 1988. Nature, 350, 124–9.CrossRefGoogle Scholar
Binks, R. M. and Fairhead, J. D. (1992). A plate tectonic framework for the evolution of the Cretaceous rift basins in West and Central Africa. In Geodynamics of Rifting. Volume II. Case history studies on rifts: North and South America and Africa, ed. Ziegler, P. A.. Tectonophysics, 213, 141–51.CrossRefGoogle Scholar
Biot, M. A. (1941). General theory of three dimensional consolidation. Journal of Applied Physics, 12, 155–64.CrossRefGoogle Scholar
Biot, M. A. (1956). Theory of propagation of elastic waves in a fluid-saturated porous solid: I. Low-frequency range. Journal of the Acoustical Society of America, 28, 168–78.Google Scholar
Birch, F. and Clark, H. (1940). The thermal conductivity of rocks and its dependence upon temperature and composition. American Journal of Science, 238, 529–58.Google Scholar
Bird, P. (1978a). Finite element modeling of lithosphere deformation: the Zagros collision orogeny. Tectonophysics, 50, 307–36.CrossRefGoogle Scholar
Bird, P. (1978b). Initiation of intracontinental subduction in the Himalayas. Journal of Geophysical Research, 10, 4975–87.Google Scholar
Bischoff, M., Endrum, B., and Meier, T. (2006). Lower crustal anisotropy in Central Europe deduced from dispersion analysis of Love and Rayleigh waves. Geophysical Research Abstracts, 8.Google Scholar
Bishop, M. G. (2001). South Sumatra Basin province, Indonesia: the Lahat/Talang Akar-Cenozoic total petroleum system. USGS Open-File Report, 99-50-S, p. 22.Google Scholar
Bishop, R. S., Gehman, J. H. M., and Young, A. (1984). Concepts for estimating hydrocarbon accumulation and dispersion. In Petroleum Geochemistry and Basin Evaluation, ed. Demaison, G. and Murris, R. J.. American Association of Petroleum Geologists, 35, 4152.Google Scholar
Bissada, K. K. (1982). Geochemical constraints on petroleum generation and migration – a review. Proceeding of the Second ASCOPE Conference, Manilla, October 1981, pp. 69–87.Google Scholar
Biteau, J. J., Choppin de Janvry, G., and Perrodon, A. (2003). How the petroleum system relates to the petroleum province. Oil and Gas Journal, 101, 46–9.Google Scholar
Bjorlykke, K., Ramm, M., and Saigal, G. C. (1989). Sandstone diagenesis and porosity modification during basin evolution. Geologische Rundschau, 78, 243–68.Google Scholar
Bjorlykke, M., Dypvik, H., and Finstad, K. G. (1975). The Kimmeridge shale, its composition and radioactivity. In Proceedings from NPS Jurassic Northern North Sea Symposium, ed. Finstad, K. G. and Selley, R. C.. Norwegian Petroleum Society, Bergen, pp. 12.1–12.20.Google Scholar
Blackwell, D. D. (1978). Heat flow and energy loss in the Western United States. In Cenozoic Tectonics and Regional Geophysics of the Western Cordillera, ed. Eaton, G. P. and Smith, R. B.. Geological Society of America Memoir, 152, 209–50.Google Scholar
Blackwell, D. D. and Richards, M. C. (2004). Geothermal Map of North America. Scale 1:6,500,000. AAPG, Tulsa, Oklahoma.Google Scholar
Blackwell, D. D. and Steele, J. L. (1989). Thermal conductivity of sedimentary rocks: measurement and significance. In Thermal History of Sedimentary Basins – Methods and Case Histories, ed. Naeser, N. D. and McCulloh, T. H.. Heidelberg, Springer, pp. 1436.Google Scholar
Blackwell, D. D., Steele, J. L., and Brott, C. A. (1981). Heat flow in the Pacific Northwest. In Physical Properties of Rocks and Minerals, Vol. II-2, ed. Toulokian, Y. S., Judd, W. R., and Roy, R. F.. McGraw-Hill, New York, pp. 1510.Google Scholar
Blanc, P. and Connan, J. (1993). Crude oils in reservoirs: the factors influencing their composition. In Applied Petroleum Geochemistry, ed. Bordenave, M. L.. Technip, Paris, pp. 151–74.Google Scholar
Blanc, P. and Connan, J. (1994). Preservation, degradation, and destruction of trapped oil. In The Petroleum System – from Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 237–47.Google Scholar
Blanpied, M. L., Lockner, D. A., and Byerlee, J. D. (1995). Frictional slip of granite at hydrothermal conditions. Journal of Geophysical Research, 100, 13045–64.CrossRefGoogle Scholar
Bodell, J. M. and Chapman, D. S. (1982). Heat flow in the north-central Colorado Plateau. Journal of Geophysical Research, 87, 2869–84.CrossRefGoogle Scholar
Bodenlos, A. J. (1970). Cap rock development and salt stock movement. In The Geology and Technology of Gulf Coast Salt, ed. Kupfer, D. H.. Symposium Proceedings, Louisiana State University, Baton Rouge, p. 192.Google Scholar
Boehm, A. and Moore, J. C. (2002). Fluidized sandstone intrusions as an indicator of paleostress orientation, Santa Cruz, California. Geofluids, 2, 147–61.CrossRefGoogle Scholar
Boettcher, S. S., James, A., Wenger, L., Gross, O., Harrison, S., and Hood, K. E. (1998). Integration of seismic and surface geochemistry data for evaluation of hydrocarbon systems in the Gulf of Mexico. AAPG International Conference and Exhibition, 82, p. 1893.Google Scholar
Bohacs, K. M. (1998). Contrasting expressions of depositional sequences in mudrocks from marine to nonmarine environs. In Shales and Mudstones I, ed. Schieber, J., Zimmerle, W., and Sethi, P.. Schweizerbartische Verlangsbuchhandlung, Stuttgart, pp. 3378.Google Scholar
Bohacs, K. M., Carroll, A. R., Neal, J. E., and Mankiewicz, P. J. (2000). Lake-basin type, source potential, and hydrocarbon character: an integrated sequence- stratigraphic-geochemical framework. In Lake Basins through Space and Time, ed. Gierlowski-Kordesh, E. H. and Kelts, K. R.. American Association of Petroleum Geologists Studies in Geology, 46, 334.Google Scholar
Bohacs, K. M. and Suter, J. (1997). Sequence stratigraphic distribution of coaly rocks: fundamental controls and paralic examples. American Association of Petroleum Geologists Bulletin, 81, 1612–39.Google Scholar
Bohannon, R. G., Naeser, C. W., Schmidt, D. L., and Zimmermann, R. A. (1989). The timing of uplift, volcanism and rifting peripheral to the Red Sea: a case for passive rifting? Journal of Geophysical Research, 94, 1683–701.CrossRefGoogle Scholar
Boillot, G., Girardeau, J., and Kornprobst, J. (1989). Rifting of the West Galicia continental margin: a review. Bulletin de la Societe Geologique de France, 8, 393400.CrossRefGoogle Scholar
Boillot, G., Grimaud, S., Mauffret, A., Mougenot, D., Kornprobst, J., Mergoil-Daniel, J., and Torrent, G. (1980). Ocean–continent boundary off the Iberian margin: a serpentinite diapir west of the Galicia Bank. Earth and Planetary Science Letters, 48, 2334.CrossRefGoogle Scholar
Boillot, G., Recq, M., Winterer, E., Meyer, A. W., Applegate, J., Baltuck, M., Bergen, J. A., Comas, M. C., Davis, T. A., Dunham, K., Evans, C. A., Girardeau, J., Goldberg, D. G., Haggerty, J., Jansa, L. F., Johnson, J. A., Kasahara, J., Loreau, L.-P., Luna-Sierra, E., Moullade, M., Ogg, J., Sarti, M., Thurow, J., and Williamson, M. (1987). Tectonic denudation of the upper mantle along passive margins: a model based on drilling results (ODP Leg 103, western Galicia margin, Spain). Tectonophysics, 132, 335–42.CrossRefGoogle Scholar
Bolchert, G., Weimer, P., and McBride, B. C. (2000). Structural and stratigraphic controls on petroleum seeps, Green Canyon and Ewing Bank, Northern Gulf of Mexico: implications for petroleum migration. Gulf Coast Association of Geological Societies Transactions, 50, 6574.Google Scholar
Boles, J. R., Clark, J. F., Washburn, L., and Leifer, I. (2001). Temporal variation in natural methane seep rate due to tides and other factors, Coal Point area, California. Journal of Geophysical Research, 106, 27077–86.CrossRefGoogle Scholar
Boles, J. R., Eichhubl, P., Garven, G., and Chen, J. (2004). Evolution of a hydrocarbon migration pathway along basin-bounding faults: evidence from fault cement. American Association of Petroleum Geologists Bulletin, 88, 947–70.CrossRefGoogle Scholar
Bonatti, E. (1976). Serpentinite protrusions in the oceanic crust. Earth and Planetary Science Letters, 32, 107–13.CrossRefGoogle Scholar
Bonatti, E. (1978). Vertical tectonism in oceanic fracture zones. Earth and Planetary Science Letters, 37, 369–79.CrossRefGoogle Scholar
Bonatti, E. (1985). Punctiform initiation of seafloor spreading in the Red Sea during transition from a continental to an oceanic rift. Nature, 316, 33–7.CrossRefGoogle Scholar
Bonatti, E., Colantoni, P., Della Vedova, B., and Taviani, M. (1984). Geology of the Red Sea transitional zone (22°N–25°N). Oceanologica Acta, 7, 385–98.Google Scholar
Boness, N. L. and Zoback, M. D. (2004). Multi-scale crustal seismic anisotropy in the region surrounding the San Andreas Fault near Parkfield California. American Geophysical Union, Fall Meeting 2004, Abstract, T11F-05.Google Scholar
Bos, B. (2000). Faults, fluids and friction: effect of pressure solution and phyllosilicates on fault slip behaviour, with implications for crustal rheology. PhD thesis, Universiteit Utrecht, Utrecht, p. 58.Google Scholar
Bos, B. and Spiers, C. J. (2002). Frictional-viscous flow of phyllosilicate-bearing fault rock: microphysical model and implications for crustal strength profiles. Journal of Geophysical Research, 107, 113.CrossRefGoogle Scholar
Bosence, D. W. J. (1998). Stratigraphic and sedimentological models of rift basins. In Sedimentation and Tectonics of Rift Basins: Red Sea-Gulf of Aden, ed. Purser, B. H. and Bosence, D. W. J. Chapman and Hall, London, pp. 925.CrossRefGoogle Scholar
Bostick, N. H. (1979). Maturation of Organic Matter and Generation of Petroleum in Tertiary Oil Basins. Geological Survey Professional Paper, Washington, DC, p. 1150.Google Scholar
Bostick, N. H. and Alpern, B. (1977). Principles of sampling, preparation and constituent selection for microphotometry in measurement of maturation of sedimentary organic matter. Journal of Microscopy, 109, 41–7.CrossRefGoogle Scholar
Bosworth, W. (1985). Geometry of propagating continental rifts. Nature, 316, 625–7.CrossRefGoogle Scholar
Bosworth, W. (1992). Mesozoic and early Tertiary rift tectonics in East Africa. Tectonophysics, 209, 115–37.CrossRefGoogle Scholar
Bosworth, W. (1995). A high-strain rift model for the southern Gulf of Suez (Egypt). Geological Society Special Publications, 80, 75102.CrossRefGoogle Scholar
Bott, M. H. P. and Dean, D. S. (1973). Stress diffusion from plate boundaries. Nature, 243, 339–41.CrossRefGoogle Scholar
Bott, M. H. P. and Kusznir, N. J. (1984). The origin of tectonic stress in the lithosphere. Tectonophysics, 105, 113.CrossRefGoogle Scholar
Bottero, C. J., Desneulin, J., Joubert, J. B., and Martine, J. P. (1989). Final geological report of (the) Matondo 1 well, Equatorial Guinea, Elf Aquitaine.Google Scholar
Bouillin, J. P., Basile, C., Labrin, E., and Mascle, J. (1998). Thermal constraints on the Cote D’Ivoire–Ghana transform margin: evidence from apatite fission track ages. Proceedings of the Ocean Drilling Program, Scientific Results, 159, 43–8.Google Scholar
Bowen, B. B., Martini, B. A., Chan, M. A., and Parry, W. T. (2007). Reflectance spectroscopic mapping of diagenetic heterogeneities and fluid-flow pathways in the Jurassic Navajo Sandstone. American Association of Petroleum Geologists Bulletin, 91, 173–90.CrossRefGoogle Scholar
Bown, J. W. and White, R. S. (1994). Variation with spreading rate of oceanic crustal thickness and geochemistry. Earth and Planetary Science Letters, 121, 435–49.CrossRefGoogle Scholar
Bown, J. W. and White, R. S. (1995). Effect of finite extension rate on melt generation at rifted continental margins. Journal of Geophysical Research, 100, 18011–29.CrossRefGoogle Scholar
Boyd, F. R. (1989). Compositional distinction between oceanic and cratonic lithosphere. Earth and Planetary Science Letters, 96, 1526.CrossRefGoogle Scholar
Brace, W. F. (1961). Dependence of fracture strength of rocks on grain size. Bulletin of the Mineral Industries Experiment Station, Mining Engineering Series, Rock Mechanics, 76, 99103.Google Scholar
Brace, W. F. and Kohlstedt, D. L. (1980). Limits on lithospheric stress imposed by laboratory experiments. Journal of Geophysical Research, 85, 6248–52.CrossRefGoogle Scholar
Brace, W. F., Paulding, B. W., and Scholz, C. (1966). Dilatancy in the fracture of crystalline rocks. Journal of Geophysical Research, 71, 3939–53.CrossRefGoogle Scholar
Bradley, W. H. and Eugster, H. P. (1969). Geochemistry and paleolimnology of the trona deposits and associated authigenic minerals of the Green River Formation of Wyoming. USGS Professional Paper, 496, 186.Google Scholar
Brandt, J. L. (1965). Stratigraphic clay-mineral distribution in the Cretaceous Colorado Group near Saskatoon (Saskatchewan). Master’s thesis, University of Saskatchewan, Saskatoon.Google Scholar
Brantley, S. and Glicken, H. (1986). Volcanic debris avalanches. Earthquakes and Volcanoes, U.S. Geological Survey Report, 18, 195206.Google Scholar
Braun, J. and Beaumont, C. (1987). Styles of continental rifting: results from dynamic models of lithosphere extension. In Sedimentary Basins and Basin-Forming Mechanisms, ed. Beaumont, C. and Tankard, A. J.. Canadian Society of Petroleum Geologists, 12, 241–58.Google Scholar
Braun, J. and Beaumont, C. (1989). A physical explanation of the relation between flank uplifts and the breakup unconformity at rifted margins. Geology, 17, 760–4.2.3.CO;2>CrossRefGoogle Scholar
Brassell, S. C., Sheng, G., Fu, J., and Eglinton, G. (1988). Biological markers in lacustrine Chinese oil shales. In Advances in Organic Geochemistry 1985, ed. Leythaeuser, D. and Rulkotter, J.. Pergamon Press, Oxford, pp. 927–41.Google Scholar
Bredehoeft, J. D., Djevanshir, R. D., and Belitz, K. R. (1988). Lateral fluid flow in a compacting sand-shale sequence: South Caspian Basin. American Association of Petroleum Geologists Bulletin, 72, 416–24.Google Scholar
Bredehoeft, J. D. and Ingebritsen, S. E. (1990). Degassing of carbon dioxide as a possible source of high pore pressure in the crust. In The Role of Fluids in Crustal Processes, ed. Bredehoeft, J. D. and Norton, D. L.. National Academy Press, Washington, DC, pp. 158–64.Google Scholar
Bredehoeft, J. D. and Papadopulos, I. S. (1965). Rates of vertical groundwater movement estimated from the Earth’s thermal profile. Water Resources Research, 1, 325–8.CrossRefGoogle Scholar
Brekke, H. (2000). The tectonic evolution of the Norwegian Sea Continental Margin with emphasis on the Vøring and Møre Basins. In Dynamics of Norwegian Margin, ed. Nottvedt, A., Larsen, B. T., Olaussen, S., Tørudbakken, B., Skogseid, J., Gabnelson, R. H., Brekke, H., and Birkeland, O.. Geological Society of London Special Publications, 167, 327–8.Google Scholar
Brennand, T. P. and Van Veen, F. R. (1975). The Auk oil-field. In Petroleum and the Continental Shelf of North-West Europe, v. 1, Geology, ed. Woodland, A. W.. Halstead Press, New York, pp. 275–83.Google Scholar
Briais, A., Patriat, P., and Tapponnier, P. (1993). Updated interpretation of magnetic anomalies and seafloor spreading stages in the South China Sea: implications for the Tertiary tectonics of Southeast Asia. Journal of Geophysical Research, 98, 6299–328.CrossRefGoogle Scholar
Briais, A. and Rabinowicz, M. (2002). Temporal variations of the segmentation of slow to intermediate spreading mid-ocean ridges 1. Synoptic observations based on satellite altimetry data. Journal of Geophysical Research, 107, 3-1–317.CrossRefGoogle Scholar
Bridge, J. S. (1999). Alluvial architecture of the Mississippi Valley: predictions using a 3D simulation model. In Floodplains: Interdisciplinary Approaches, ed. B. Marriott, S. and Alexander, J.. Geological Society Special Publications, 163, 269–78.Google Scholar
Bridge, J. S. (2003). Rivers and floodplains: forms, processes, and sedimentary record. Blackwell Publishing, Oxford, p. 491.Google Scholar
Bridge, J. S. and Tye, R. S. (2000). Interpreting the dimensions of ancient fluvial channel bars, channels, and channel belts from wireline-logs and cores. American Association of Petroleum Geologists Bulletin, 84, 1205–28.Google Scholar
Brigaud, F., Chapman, D. S., and Le Douaran, S. (1990). Estimating thermal conductivity in sedimentary basins using lithologic data and geophysical well logs. American Association of Petroleum Geologists Bulletin, 74, 1459–77.Google Scholar
Briggs, D. J. (1980). An investigation into variations in mineralogical and mechanical properties of coal measure strata. PhD thesis, University of Wales, Cardiff.Google Scholar
Briole, P., Rigo, A., Lyon-Caen, H., Ruegg, J. C., Papazissi, K., Mitsakaki, C., Balodimou, A., Veis, G., Hatzfeld, D., and Deschamps, A. (2000). Active deformation of the Corinth Rift, Greece: results from repeated global positioning system surveys between 1990 and 1995. Journal of Geophysical Research, 105, 25605–25.CrossRefGoogle Scholar
Brocher, T. M. (2005). Empirical relations between elastic wavespeeds and density in the Earth’s crust. Bulletin of the Seismological Society of America, 95, 2081–92.CrossRefGoogle Scholar
Brodie, J. and White, N. (1994). Sedimentary basin inversion caused by igneous underplating: northwest European continental shelf. Geology, 22, 147–50.2.3.CO;2>CrossRefGoogle Scholar
Brodie, J. and White, N. (1995). The link between sedimentary basin inversion and igneous underplating. In Basin Inversion, ed. Buchanan, J. G. and Buchanan, P. G.. Geological Society Special Publications, 88, 2138.Google Scholar
Bromann-Klausen, M. and Larsen, H. C. (2002). East- Greenland coast-parallel dike swarm and its role in continental break-up. In Volcanic Rifted Margins, ed. Menzies, M. A., Klemperer, S. L., Ebinger, C. J., and Baker, J.. GSA Special Papers, 362, 133–58.Google Scholar
Brooke, C. M., Trimble, T. J., and Mackay, T. A. (1995). Mounded shallow gas sands from the Quaternary of the North Sea: analogues for the formation of sand mounds in deep water Tertiary sediments? Geological Society London Special Publication, 94, 95101.CrossRefGoogle Scholar
Brooks, C. K. and Nielsen, T. D. F. (1982). The E Greenland continental margin: a transition between oceanic and continental magmatism. Journal of the Geological Society of London, 136, 265–75.Google Scholar
Brooks, M., Hillier, B. V., and Miliorizos, M. (1993). New seismic evidence for a major geological boundary at shallow depth, N Devon. Journal of the Geological Society of London, 150, 131–5.CrossRefGoogle Scholar
Brooks, M., Trayner, P. M., and Trimble, T. (1988). Mesozoic reactivation of Variscan thrusting in the Bristol Channel area, UK. Journal of the Geological Society of London, 145, 439–44.CrossRefGoogle Scholar
Brophy, P. (2003). Geothermal exploration techniques. Short course, GRC Annual Meeting 2003, Morelia.Google Scholar
Brosse, E. and Huc, A. Y. (1986). Organic parameters as indicators of thermal evolution in the Viking Graben. In Thermal Modelling in Sedimentary Basins, ed. Burrus, J.. Technip, Paris, pp. 173–95.Google Scholar
Brott, C. A., Blackwell, D. D., and Ziagos, J. P. (1981). Thermal and tectonic implications of heat flow in the eastern Snake River, Idaho. Journal of Geophysical Research, 86, 11709–34.CrossRefGoogle Scholar
Brown, D. W. (1986). The Bay of Fundy: thin-skinned tectonics and resultant early Mesozoic sedimentation. Basins of Eastern Canada and Worldwide Analogues, Atlantic Geoscience Society Programme, Abstracts, Halifax, 26.Google Scholar
Brown, K. M. and Moore, J. C., 1993. Comment on “Anisotropic permeability and tortuosity in deformed wet sediments” by Arch, J. and Maltman, A. J.. Journal of Geophysical Research, 98, 17859–64.CrossRefGoogle Scholar
Brown, R. W. (1991). Backstacking apatite fission-track “stratigraphy”: a method for resolving the erosional and isostatic rebound components of tectonic uplift histories. Geology, 19, 74–7.2.3.CO;2>CrossRefGoogle Scholar
Brown, R. W., Gallagher, K., Gleadow, A. J. W., and Summerfield, M. A. (2000). Morphotectonic evolution of the South Atlantic margins of Africa and South America. In Geomorphology and Global Tectonics, ed. Summerfield, M. A.. John Wiley and Sons, Chichester, UK, pp. 255–81.Google Scholar
Brown, R. W., Rust, D. J., Summerfield, M. A., Gleadow, A. J. W., and de Wit, M. C. J. (1990). An Early Cretaceous phase of accelerated erosion on the southwestern margin of Africa: evidence from apatite fission track analysis and the offshore sedimentary record. Nuclear Tracks and Radiation Measurements, 17, 339–50.Google Scholar
Brown, S. (1990). Jurassic. In Introduction to the Petroleum Geology of the North Sea, ed. Glennie, K. W.. Blackwell Scientific Publications, Oxford, pp. 219–49.Google Scholar
Brownfield, M. E. and Charpentier, R. R. (2003). Assessment of the undiscovered oil and gas of the Senegal Province, Mauritania, Senegal, Gambia, and Guinea-Bissau, Northwest Africa. U.S. Geological Survey Bulletin, 2207, 126.Google Scholar
Brozena, J. M. and White, R. S. (1990). Ridge jumps and propagations in the South Atlantic Ocean. Nature, 348, 149–52.CrossRefGoogle Scholar
Bruce, C. H. (1984). Smectite dehydration – its relation to structural development and hydrocarbon accumulation in northern Gulf of Mexico Basin. American Association of Petroleum Geologists Bulletin, 68, 673–83.Google Scholar
Bruhn, R. L., Yonkee, W., and Parry, W. (1990). Structural and fluid-chemical properties of seismogenic normal faults. Tectonophysics, 175, 139–57.CrossRefGoogle Scholar
Brun, J. P. (1999). Narrow rifts versus wide rifts: inferences for the mechanics of rifting from laboratory experiments. Philosophical Transactions – Royal Society, Mathematical, Physical and Engineering Sciences, 357, 695712.CrossRefGoogle Scholar
Brun, J. P. and Beslier, M. O. (1996). Mantle exhumation at passive margins. Earth and Planetary Science Letters, 142, 161–73.CrossRefGoogle Scholar
Brun, J. P. and Choukroune, P. (1983). Normal faulting, block tilting, and decollement in a stretched crust. Tectonics, 2, 345–56.CrossRefGoogle Scholar
Brun, J. P., Gutscherr, M. A., and ECORS-DEKORP team (1992). Deep structure of the Rhine Graben from DEKORP-ECORS seismic reflection data, a summary. In Geodynamics of Rifting, Volume I, Case History Studies on Rifts: Europe and Asia, ed. Ziegler, P. A.. Tectonophysics, 208, 139–47.Google Scholar
Brun, J. P., Sokoutis, D., and van den Driessche, J. (1994). Analogue modeling of detachment fault systems and core complexes. Geology, 22, 319–22.2.3.CO;2>CrossRefGoogle Scholar
Bruton, D. J. and Helgeson, H. C. (1983). Calculation of the chemical and thermodynamic consequences of differences between fluid and geostatic pressure in hydrothermal systems. American Journal of Science, 283, 540–88.Google Scholar
Bryant, B. (1990). Geologic map of the Salt Lake City 30’ x 60’ quadrangle, north-central Utah, and Uinta County, Wyoming. U.S. Miscellaneous Investigations Series Map I-1944, scale 1:100,000.Google Scholar
Buck, P. (1986). Sedimentology and micropalaeontology of gravity cores from the NE Atlantic Ocean (south west of Ireland). Technical Report – Marine Geoscience Unit, Joint Geological Survey/University of Cape Town, 16, 106–12.Google Scholar
Buck, W. R. (1985). When does small-scale convection begin beneath oceanic lithosphere? Nature, 313, 775–7.CrossRefGoogle Scholar
Buck, W. R. (1986). Small scale convection induced by passive rifting: the cause for uplift of rift shoulders. Earth Planetary Science Letters, 77, 362–72.Google Scholar
Buck, W. R. (1991). Modes of continental lithospheric extension. Journal of Geophysical Research, 96, 20161–78.CrossRefGoogle Scholar
Buck, W. R. (2004). Consequences of asthenospheric variability on continental rifting. In Rheology and Deformation of the Lithosphere at Continental Margins, ed. Karner, G. D., Taylor, B., Driscolland, N. W., and Kohlstedt, D. L.. Columbia University Pages, pp. 131.Google Scholar
Buck, W. R., Lavier, L. L., and Poliakov, A. N. B. (1999). How to make a rift wide. Mathematical, Physical and Engineering Sciences, 357, 671–93.CrossRefGoogle Scholar
Buck, W. R., Lavier, L. L., and Poliakov, A. N. B. (2005). Modes of faulting at mid-ocean ridges. Nature, 43, 719–23.Google Scholar
Buck, W. R., Martinez, F., Steckler, M. S., and Cochran, J. R. (1988). Thermal consequences of lithospheric extension: Pure and simple. Tectonics, 7, 213–34.CrossRefGoogle Scholar
Buntebarth, G. (1984). Geothermics. Springer-Verlag, New York.CrossRefGoogle Scholar
Buntebarth, G. and Rybach, L. (1981). Linear relationships between petrophysical properties and mineralogical constitution: preliminary results. Tectonophysics, 75, 41–6.CrossRefGoogle Scholar
Burbank, D. W., Derry, L. A., and France-Lanord, C. (1993). Reduced Himalayan sediment production 8 Myr ago despite an intensified monsoon. Nature, 364, 4850.CrossRefGoogle Scholar
Burchfiel, B. C., Hodges, K. V., and Royden, L. H. (1987). Geology of Panamint Valley-Saline Valley pull-apart system, California: palinspastic evidence for low-angle geometry of a Neogene range-bounding fault. Journal of Geophysical Research, 92, 10422–6.CrossRefGoogle Scholar
Burchfiel, B. C. and Stewart, J. H. (1966). “Pull-apart” origin of the central segment of Death Valley, California. Geological Society of America Bulletin, 77, 439–41.CrossRefGoogle Scholar
Burford, R. O. and Harsh, P. W. (1980). Slip on the San Andreas Fault in central California from alignment array surveys. Bulletin of the Seismological Society of America, 70, 1233–61.Google Scholar
Burgess, C. F., Rosendahl, B. R., Sander, S., Burgess, C. A., Lambiase, J., Derksen, S., and Meader, N. (1988). The structural and stratigraphic evolution of Lake Tanganyika: a case study of continental rifting. In Triassic-Jurassic Rifting: Continental Breakup and the Origin of the Atlantic and Passive Margins, ed. Manspeizer, W.. Elsevier, Amsterdam, pp. 859–81.Google Scholar
Burgess, P. M. and Hovius, N. (1998). Rates of delta progradation during highstands: consequences for timing of deposition in deep-marine systems. Journal of the Geological Society of London, 155, 217–22.CrossRefGoogle Scholar
Burke, K. (1972). Longshore drift, submarine canyons and submarine fans in development of Niger delta. American Association of Petroleum Geologists Bulletin, 56, 1975–83.Google Scholar
Burley, S. D. (1984). Distribution and origin of authigenic minerals in the Triassic Sherwood Sandstone Group, UK. Clay Minerals, 19, 403–40.Google Scholar
Burley, S. D., Mullis, J., and Matter, A. (1989). Timing of diagenesis in the Tartan Reservoir (UK North Sea): constraints from combined cathodoluminescence microscopy and fluid inclusion studies. Marine and Petroleum Geology, 6, 98120.CrossRefGoogle Scholar
Burnham, A. K., Schmidt, B. J., and Braun, B. L. (1995). A test of the parallel reaction model using kinetic measurements on hydrous pyrolysis residues. Organic Geochemistry, 23, 931–9.CrossRefGoogle Scholar
Burns, B. A., Heller, P. L., Marzo, M., and Paola, C. (1997). Fluvial response in a sequence stratigraphic framework: example from the Montserrat fan delta, Spain. Journal of Sedimentary Research, 67, 311–21.Google Scholar
Burov, E. and Cloetingh, S. (1997). Erosion and rift dynamics: new thermomechanical aspects of post-rift evolution of extensional basins. Earth and Planetary Science Letters, 150, 726.CrossRefGoogle Scholar
Burov, E. B. and Diament, M. (1995). The effective elastic thickness (Te) of continental lithosphere: What does it really mean? Journal of Geophysical Research, 100, 3905–28.CrossRefGoogle Scholar
Burov, E. and Poliakov, A. (2001). Erosion and rheology controls on synrift and postrift evolution: verifying old and new ideas using a fully coupled numerical model. Journal of Geophysical Research, 106, 16461–81.CrossRefGoogle Scholar
Burov, E. B. and Watts, A. B. (2006). The long-term strength of continental lithosphere: “jelly sandwich” or “crème brulée”? GSA Today, 16.2.0.CO;2>CrossRefGoogle Scholar
Burr, T. N. and Currey, D. R. (1988). The Stockton Bar. In In the Footsteps of G. K. Gilbert – Lake Bonneville and Neotectonics of the Eastern Basin and Range Province, ed. Machettte, M. N.. Utah Geological and Mineral Survey Miscellaneous Publication, 88, 6673.Google Scholar
Burr, T. N. and Currey, D. R. (1992). The Stockton Bar Miscellaneous Publication – Utah Geological Survey, Salt Lake City.Google Scholar
Burrus, J., Schneider, F., and Wolf, S. (1994). Modeling overpressures in sedimentary basins: consequences for permeability and rheology of shales, and petroleum expulsion efficiency. American Association of Petroleum Geologists Bulletin, 78, 1137.Google Scholar
Burruss, R. C., Cercone, K. R., and Harris, P. M. (1985). Timing of hydrocarbon migration: evidence from fluid inclusions in calcite cements, tectonics and burial history. In Carbonate Cements Revisited, ed. Harris, P. M. and Schneidermann, N. M.. SEPM Special Publication, pp. 277–89.Google Scholar
Burtner, R. L., Nigrini, A., and Donelick, R. A. (1994). Thermochronology of Lower Cretaceous source rocks in the Idaho-Wyoming thrust belt. American Association of Petroleum Geologists Bulletin, 78, 1613–36.Google Scholar
Burton, K. W., Schiano, P., Birck, J. L., Allegre, C. J., Rehkamper, M., Halliday, A. N., and Dawson, J. B. (2000). The distribution and behavior of rhenium and osmium amongst mantle minerals and the age of the lithospheric mantle beneath Tanzania. Earth and Planetary Science Letters, 183, 93106.CrossRefGoogle Scholar
Burwood, R., De Witte, S. M., Mycke, B., and Paulet, J. (1995). Petroleum geochemical characterization of the Lower Congo coastal basin Bucomazi Formation. In Petroleum Source Rocks, ed. Katz, B. J.. Springer-Verlag, Berlin, pp. 235–64.Google Scholar
Burwood, R., Mycke, B., Paulet, J., Jacobs, L., and Hall, D. (1997). The Central Graben Upper Jurassic source rock system: a sequence stratigraphic approach to oil provenance. In Organic Geochemistry: Developments and Applications to Energy, Climate, Environment and Human History, ed. J. O. Grimalt and C. Dorronsoro. Selected papers from 17th International Meeting on Organic Geochemistry, San Sebastian, Spain, 4–8 September 1995, pp. 249–52.Google Scholar
Byerlee, J. D. (1968). Brittle-ductile transition in rocks. Journal of Geophysical Research, 73, 4741–50.CrossRefGoogle Scholar
Byerlee, J. (1978). Friction of rocks. Pure and Applied Geophysics, 116, 615–26.CrossRefGoogle Scholar
Caine, J. S. (1999). The Architecture and Permeability Structure of Brittle Fault Zones. Salt Lake City: University of Utah.Google Scholar
Caine, J. S., Evans, J. P., and Forster, C. B. (1996). Fault zone architecture and permeability structure. Geology, 24, 1025–8.2.3.CO;2>CrossRefGoogle Scholar
Cainelli, C. and Mohriak, W. U. (1999). Some remarks on the evolution of sedimentary basins along the eastern Brazilian continental margin. Episodes, 22, 206–16.CrossRefGoogle Scholar
Callot, J.-P. (2002). Structure et developpement des marges volcaniques: l’example du Groenland. PhD thesis, University of Paris, p. 584.Google Scholar
Callot, J.-P., Geoffroy, L., Aubourg, C., Pozzi, J.-P., and Mege, D. (2001). Magma flow directions of shallow dykes from the East-Greenland volcanic margin inferred from magnetic fabric studies. Tectonophysics, 335, 313–29.CrossRefGoogle Scholar
Callot, J.-P., Geoffroy, L., and Brun, J. P. (2002). 3D analogue modelling of volcanic passive margins. Tectonics, 21.Google Scholar
Camelbeeck, T. and Iranga, M. D. (1996). Deep crustal earthquakes and active faults along the Rukwa Trough, eastern Africa. Geophysical Journal International, 124, 612–30.CrossRefGoogle Scholar
Camp, W. K. (2000). Geologic model and reservoir description of the deepwater “p sand” at subsalt Mahogany Field, Gulf of Mexico. In Volume Integration of Geologic Models for Understanding Risk in the Gulf of Mexico, ed. R. Shoup, J. Watkins, J. Karlo, and D. Hall. Papers from Hedberg Conference, 1998, Galveston, AAPG/Databases Discovery Series 1 (CD-ROM).Google Scholar
Campanha, V. A. (1987). Análises bioestratigráficas do poço 2-AP-1-CE, ITP report, Relatório 24769. Instituto De Pesquisas Técnicas, São Paulo, p. 33.Google Scholar
Campbell, I. H. and Griffiths, R. W. (1990). Implications of mantle plume structure for the evolution of flood basalts. Earth and Planetary Science Letters, 99, 7993.CrossRefGoogle Scholar
Canales, J. P., Detrick, R. S., Lin, J., Collins, J. A., and Toomey, D. R. (2000). Crustal and upper mantle seismic structure beneath the rift mountains and across a nontransform offset at the Mid-Atlantic Ridge. Journal of Geophysical Research, 105, 2699–720.CrossRefGoogle Scholar
Cande, S. C., LaBrecque, J. L., and Haxby, W. F. (1988). Plate kinematics of the South Atlantic: Chron C34 to present. Journal of Geophysical Research, 93, 13479–92.CrossRefGoogle Scholar
Cande, S. C. and Rabinowicz, P. D. (1979). Magnetic Anomalies of the Continental Margin of Brazil. American Association of Petroleum Geologists, Tulsa, Oklahoma.Google Scholar
Cann, J. R., Blackman, D. K., Smith, D. K., McAllister, E., Janssen, B., Mello, S., Avgerinos, E., Pascoe, A. R., and Escartín, J. (1997). Corrugated slip surfaces formed at ridge-transform intersections on the Mid-Atlantic Ridge. Nature, 385, 329–32.CrossRefGoogle Scholar
Cannat, M., Rommevaux-Jestin, C., Sauter, D., Deplus, C. and Mendel, V. (1999). Formation of the axial relief at the very slow spreading Southwest Indian Ridge (49 degrees to 69 degrees E). Journal of Geophysical Research, 104.B10, 22825–22843.CrossRefGoogle Scholar
Cao, J., Yao, S. P., Jin, Z. J., Hu, W. X., Zhang, Y. J., Wang, X. L., Zhang, Y. Q., and Tang, Y. (2006). Petroleum migration and mixing in NW Junggar Basin (NW China): constraints from oil-bearing fluid inclusion analyses. Organic Geochemistry, 37, 827–46.CrossRefGoogle Scholar
Cao, J., Zhang, Y. J., Hu, W. X., Yao, S. P., Wang, X. L., Zhang, Y. Q., and Tang, Y. (2005). The Permian hybrid petroleum system in the northwest margin of the Junggar Basin. Marine and Petroleum Geology, 22, 331–49.CrossRefGoogle Scholar
Cappetti, G., Celati, R., Cigni, U., Squarci, P., Stefani, G. C., and Taffi, L. (1985). Development of deep exploration in the geothermal areas of Tuscany, Italy. International Symposium on Geothermal Energy, Kailua-Kona, Hawaii, International Volume, pp. 303–9.Google Scholar
Capuano, R. M. (1990). Hydrochemical constraints on fluid-mineral equilibria during compaction diagenesis of kerogen-rich geopressured sediments. Geochimica et Cosmochimica Acta, 54, 1283–99.CrossRefGoogle Scholar
Capuano, R. M. (1993). Evidence of fluid flow in microfractures in geopressured shales. American Association of Petroleum Geologists Bulletin, 77, 1303–14.Google Scholar
Carey, S. W. (1958). The tectonic approach to continental drift. In Continental Drift – A Symposium, ed. Carey, S. W.. University of Tasmania, Hobart, pp. 177363.Google Scholar
Carlson, W. D., Donelick, R. A., and Ketcham, R. A. (1999). Variability of apatite fission-track annealing kinetics: I. Experimental results. American Mineralogist, 84, 1213–23.CrossRefGoogle Scholar
Carlsson, A. and Olsson, T. (1979). Hydraulic conductivity and its stress dependence. Proceedings: Workshop on Low-Flow, Low Permeability Measurements in Largely Impermeable Rocks, Paris, pp. 249–59.Google Scholar
Carroll, A. R. and Bohacs, K. M. (1995). A stratigraphic classification of lake types and hydrocarbon source potential: balancing climatic and tectonic controls. First International Limno-Geological Congress, Abstract volume, pp. 1819.Google Scholar
Carroll, A. R. and Bohacs, K. M. (1999). Stratigraphic classification of ancient lakes: balancing tectonic and climatic controls. Geology, 27, 99102.2.3.CO;2>CrossRefGoogle Scholar
Carroll, A. R. and Bohacs, K. M. (2001). Lake-type controls on petroleum source rock potential in nonmarine basins. American Association of Petroleum Geologists Bulletin, 85, 1033–53.Google Scholar
Carroll, A. R., Brassell, S. C., and Graham, S. A. (1992). Upper Permian lacustrine oil shales, southern Junggar basin, northwest China. American Association of Petroleum Geologists Bulletin, 76, 1874–902.Google Scholar
Carruthers, A., McKie, T., Price, J., Dyer, R., Williams, G., and Watson, P. (1996). The application of sequence stratigraphy to the understanding of Late Jurassic turbidite plays in the Central north Sea, UKCS. In Geology of the Humber Group: Central Graben and Moray Firth, UKCS, ed. Hurst, A., Johnson, H. D., Burley, S. D., Canham, A. C., and Mackertich, D. S.. Special Publication of Geological Society of London, 114, 2945.Google Scholar
Carslaw, H. S. and Jaeger, J. C. (1959). Conduction of Heat in Solids. Oxford University Press, New York.Google Scholar
Carter, N. L. (1976). Steady-state flow in rocks. Reviews of Geophysics and Space Physics, 14, 301–60.CrossRefGoogle Scholar
Carter, N. L. and Tsenn, M. C. (1987). Flow properties of continental lithosphere. Tectonophysics, 136, 2763.CrossRefGoogle Scholar
Cartwright, J. A. and Mansfield, C. S. (1998). Lateral displacement variation and lateral tip geometry of normal faults in the Canyonlands National Park, Utah. Journal of Structural Geology, 20, 319.CrossRefGoogle Scholar
Cartwright, J. A., Mansfield, C. S., and Trudgill, B. D. (1996). The growth of normal faults by segment linkage. In Modern Developments in Structural Interpretation, Validation, and Modelling, ed. Buchanan, P. G. and Nieuwland, D. A.. Geological Society Special Publications, pp. 163–77.Google Scholar
Cartwright, J. A., Trudgill, B. D., and Mansfield, C. S. (1995). Fault growth by segment linkage: an explanation for scatter in maximum displacement and trace length data from the Canyonlands Grabens of SE Utah. Journal of Structural Geology, 17, 1319–26.CrossRefGoogle Scholar
Casey, B. J., Romani, R. S., and Schmitt, R. H. (1993). Appraisal geology of the Saltire field, Witch Ground Graben, North Sea. In Petroleum Geology of Northwest Europe, ed. Parker, J. R.. Proceedings of the 4th Conference, Geological Society of London, pp. 507–17.Google Scholar
Cassignol, C., Cornette, Y., David, B., and Gillot, P.-Y. (1978). Technologie potassium-argon. Rapport CEA R-4802, C.E.N. Saclay Publication, p. 37.Google Scholar
Catalan, L., Xiaowen, F., Chatzis, I., and Dullien, F. A. I. (1992). An experimental study of secondary oil migration. American Association of Petroleum Geologists Bulletin, 76, 638–50.Google Scholar
Cathles, L. M., Erendi, A. H. J., and Barrie, T. (1997). How long can a hydrothermal system be sustained by a single intrusive event? Economic Geology, 92, 766–71.CrossRefGoogle Scholar
Catuneanu, O. (2002). Sequence stratigraphy of clastic systems: concepts, merits and pitfalls. Journal of African Earth Sciences, 35, 143.CrossRefGoogle Scholar
Cayley, G. T. (1987). Hydrocarbon migration in the central North Sea. In Petroleum Geology of North West Europe, ed. Brooks, J. and Glennie, K. W.. Graham and Trotman, London, pp. 549–55.Google Scholar
Čermák, V. (1994). Results of heat flow studies in Czechoslovakia. In Crustal Structure of the Bohemian Massif and the West Carpathians, ed. Bucha, V. and Blížkovský, M.. Academia Praha, Springer-Verlag, pp. 85120.CrossRefGoogle Scholar
Čermák, V. and Bodri, L. (1992). Crustal thinning during rifting: a possible signature in radiogenic heat production. Tectonophysics, 209, 227–39.CrossRefGoogle Scholar
Čermák, V., Bodri, L., and Rybach, L. (1991). Radioactive heat production in the continental crust and its depth dependence. In Terrestrial Heat Flow and the Lithosphere Structure, ed. Čermák, V. and Rybach, L.. Springer, Berlin, pp. 2369.CrossRefGoogle Scholar
Čermák, V. and Haenel, R. (1988). Geothermal maps. In Handbook of Terrestrial Heat-Flow Density Determination: With Guidelines and Recommendations of the International Heat Flow Commission, ed. R. Haenel, Rybach, L., and Stegena, L.. Kluwer Academic Publishers, Dordrecht, pp. 261300.CrossRefGoogle Scholar
Čermák, V. and Hurtig, E. (1977). Preliminary heat flow map of Europe, 1:5,000,000; explanatory text. Potsdam: IASPEI, International Heat Flow Communication.Google Scholar
čermák, V. and Rybach, L. (1982). Thermal Conductivity and Specific Heat of Minerals and Rocks. Landolt-Bornstein – Group V Geophysics, Springer-Verlag, pp. 305–43.Google Scholar
CGMW (19851990). International Geological Map of Africa. Commission for the Geological Map of the World and UNESCO.Google Scholar
CGMW (2000). Geologic Map of South America. Brazil Ministry of Mines and Energy, Commission for the Geological Map of the World.Google Scholar
Chan, M. A., Parry, W. T., and Bowman, J. R. (2000). Diagenetic hematite and manganese oxides and fault-related fluid flow in Jurassic sandstones southeastern Utah. American Association of Petroleum Geologists Bulletin, 84, 1281–310.Google Scholar
Chang, H. K., Kowsmann, R. O., and Figueiredo, A. M. F. (1988). New concepts on the development of East Brazilian marginal basins. Episodes, 11, 194202.CrossRefGoogle Scholar
Chang, H. K., Kowsmann, R. O., Figueiredo, A. M. F., and Bender, A. A. (1992). Tectonics and stratigraphy of the East Brazil Rift System: an overview. Tectonophysics, 213, 97138.CrossRefGoogle Scholar
Chapman, D. S. and Rybach, L. (1985). Heat flow anomalies and their interpretation. Journal of Geodynamics, 4, 337.CrossRefGoogle Scholar
Chapman, D. S., Willett, S. D., and Clauser, C. (1991). Using thermal fields to estimate basin-scale permeabilities. In Petroleum Migration, ed. England, W. A. and Fleet, A. J.. Geological Society of London Special Publications, 59, 123–5.Google Scholar
Chapman, T. J. and Meneilly, A. W. (1990). Fault displacement analysis in seismic exploration. First Break, 8, 1112.CrossRefGoogle Scholar
Charlou, J.-L., Bougault, H., Aprioui, P., Nielsen, T., and Rona, P. (1991). Different TDM/CH4 hydrothermal plume signatures: TAG site at 26o N and serpentinized ultrabasic diaper at 15o 05’N on the Mid-Atlantic Ridge. Geochimica et Cosmochimica Acta, 55, 3209–22.CrossRefGoogle Scholar
Charlou, J.-L. and Donval, J.-P. (1993). Hydrothermal methane venting between 12 degrees N and 26 degrees N along the Mid-Atlantic Ridge. Journal of Geophysical Research, 98, 9625–42.CrossRefGoogle Scholar
Chekunov, A. V. (1989). Dynamic model of the geotraverse through the Dnieper-Donets paleorift, Ukrainian Shield, southern Carpathians. Geotectonics, 23, 467–75.Google Scholar
Chen, W.-P. and Molnar, P. (1983). Focal depths of intracontinental and intraplate earthquakes and their implications for the thermal and mechanical properties of the lithosphere. Journal of Geophysical Research, 88, 4183–214.CrossRefGoogle Scholar
Chen, Z., Zhou, G., and Alexander, R. (1994). A biomarker study of immature crude oils from the Shengli oilfield, People’s Republic of China. Chemical Geology, 113, 117–32.Google Scholar
Cherry, S. T. J. (1993). The interaction of structure and sedimentary processes controlling deposition of the Upper Jurassic Brae Formation conglomerate, Block 16/17, North Sea. In Petroleum Geology of Northwest Europe, ed. Parker, J. R.. Proceedings of the 4th Conference, Geological Society of London, pp. 387400.Google Scholar
Chery, J., Lucazeau, M., Daignieres, M., and Vilotte, J. P. (1992). Large uplift of rift flanks: A genetic link with lithospheric rigidity? Earth Science and Planetary Letters, 112, 195211.CrossRefGoogle Scholar
Chery, J., Zoback, M. D., and Hassani, R. (2001). An integrated mechanical model of the San Andreas Fault in Central and Northern California. Journal of Geophysical Research, 106, 22051–66.CrossRefGoogle Scholar
Chesley, J. T., Rudnick, R. L., and Lee, C. T. (1999). Re-Os systematic of mantle xenoliths from the East African Rift: age, structure, and history of the Tanzanian craton. Geochimica et Cosmochimica Acta, 63, 1203–17.CrossRefGoogle Scholar
Chester, F. M. and Logan, J. M. (1986). Composite planar fabric of gouge from the Punchbowl Fault, California. Journal of Structural Geology, 9, 621–34.Google Scholar
Chian, D., Louden, K. E., Minshull, T. A., and Whitmarsh, R. B. (1999). Deep structure of the 600 ocean–continent transition in the southern Iberia Abyssal Plain from seismic refraction 601 profiles: 1. Ocean Drilling Program (Legs 149 and 173) transect. Journal of Geophysical Research, 104, 7443–62.CrossRefGoogle Scholar
Chian, D., Louden, K. E., and Reid, I. (1995). Crustal structure of the Labrador Sea conjugate margin and implications for the formation of nonvolcanic continental margins. Journal of Geophysical Research, 100, 24239–53.CrossRefGoogle Scholar
Childs, C., Watterson, J., and Walsh, J. J. (1995). Fault overlap zones within developing normal fault systems. Journal of the Geological Society of London, 152, 535–49.CrossRefGoogle Scholar
Chilingarian, G. V. (1983). Compactional diagenesis. In Sediment Diagenesis, ed. Parker, A. and Sellwood, B. W.. Reidel Publishing Company, Dordrecht, pp. 57168.CrossRefGoogle Scholar
Chiozzi, P., Pasquale, M., and Verdoya, M. (2006). Seimicity and rheological modelling in an extensional- compressional tectonic realm. Geophysical Research Abstracts, 8, 02185.Google Scholar
Chorowicz, J. (1989). Transfer and transform fault zones in continental rifts: examples in the Afro-Arabian rift system, implications of crust breaking. Journal of African Earth Sciences, 8, 203–14.Google Scholar
Chorowicz, J. and Mukonki, M. N. B. (1980). Lineaments anciens, zones transformantes recentes et geotectonique des fosses de l’Est Africain, d’apres la teledetection et la microtectonique. Ancient lineaments, recent transform zones, and East African geotectonic trenches according to remote sensing and microtectonics. Rapport Annuel – Musee Royal de l’Afrique Centrale, Departement de Geologie et de Mineralogie, pp. 143–67.Google Scholar
Choubert, G. and Faure-Muret, A. (1985). International Geological Map of Africa. Commission for the Geological Map of the World and UNESCO, Paris.Google Scholar
Choubert, G. et al. (1988). International geological map of Africa. Commission for the Geological Map of the World and UNESCO, Paris.Google Scholar
Choudhuri, M., Guha, D., Dutta, A., Sinha, S., and Sinha, N. (2006). Spatio-temporal variations and kinematics of shale mobility in the Krishna-Godavari Basin, India. Poster presentation at AAPG Hedberg Conference, Port of Spain, Trinidad and Tobago, June 5–7, 2006.Google Scholar
Christensen, C., Nemčok, M., McCulloch, J., and Moore, J. (2002). The characteristics of productive zones in the Karaha–Telaga Bodas geothermal system. Geothermal Resources Council Transactions, 26, 623–6.Google Scholar
Christensen, N. I. (1989). Reflectivity and seismic properties of the deep continental crust. Journal of Geophysical Research, 94, 17793–804.CrossRefGoogle Scholar
Christensen, N. I. and Mooney, W. D. (1995). Seismic velocity structure and composition of the continental crust. Journal of Geophysical Research, 100, 9761–88.CrossRefGoogle Scholar
Christiansen, L. B. and Garven, G. (2003). A theoretical comparison of buoyancy-driven and compaction-driven fluid flow in oceanic sedimentary basins. Journal of Geophysical Research, 108.CrossRefGoogle Scholar
Christiansen, L. B. and Garven, G. (2004a). Transient hydrogeologic models for submarine flow in volcanic seamounts: 1. The Hawaiian Islands. Journal of Geophysical Research, 109.Google Scholar
Christiansen, L. B. and Garven, G. (2004b). Transient hydrogeologic models for submarine flow in volcanic seamounts: 2. Comparison of the Hawaiian, Canary, and Marquesas Islands. Journal of Geophysical Research, 109.Google Scholar
Christie-Blick, N. and Biddle, K. T. (1985). Deformation and basin formation along strike-slip faults. Special Publication – Society of Economic Paleontologists and Mineralogists, 37, 134.Google Scholar
Chung, H. M., Wingert, W. S., and Claypool, G. E. (1992). Geochemistry of oil from the northern Viking Graben. In Giant Oil and Gas Fields of the Decade 1978–1988, ed. M. T. Halbouty. AAPG Memoir, 54, pp. 277–86.Google Scholar
Chung, W.-Y. and Kanamori, H. (1980). Variation of seismic source parameters and stress drops within a descending slab and its implications in plate mechanics. Physics of the Earth and Planetary Interiors, 23, 134–59.CrossRefGoogle Scholar
Clark, J. A., Stewart, S. A., and Cartwright, J. A. (1998). Evolution of the NW margin of the North Permian Basin, UK North Sea. Journal of the Geological Society of London, 155, 663–76.CrossRefGoogle Scholar
Clark, Jr., S. P. (1966). Handbook of Physical Constants. Geological Society of America Memoir, New York, 97.Google Scholar
Clarke, B. J. (2002). Early Cenozoic Denudation of the British Isles: A Quantitative Stratigraphic Approach. University of Cambridge, Cambridge.Google Scholar
Clausen, S., Nemčok, M., Moore, J., Hulen, J., and Bartley, J. (2006). Mapping fractures in the Medicine Lake geothermal system. Geothermal Resources Council Transactions, 30, 383–6.Google Scholar
Clauser, C. and Villinger, H. (1990). Analysis of conductive and convective heat transfer in a sedimentary basin, demonstrated for the Rheingraben. Geophysical Journal International, 100, 393414.CrossRefGoogle Scholar
Clayton, J. L., Yang, J., King, J. D., Lillis, P. G., and Warden, A. (1997). Geochemistry of oils from the Junggar basin, northwest China. American Association of Petroleum Geologists Bulletin, 81, 1926–44.Google Scholar
Clem, K. (1985). Oil and gas production summary of the Uinta basin. In Geology and Energy Resources, Uinta Basin of Utah, ed. Picard, M. D.. Utah Geological Association, Salt Lake City, pp. 159–67.Google Scholar
Clemetz, D. M., Demaison, G. J., and Daly, A. R. (1979). Well site geochemistry by programmed pyrolysis. Proceedings of the 11th Annual Offshore Technology Conference, pp. 465–70.Google Scholar
Clemson, J., Cartwright, J., and Swart, J. (1999). The Namib Rift: a rift system of possible Karoo age, offshore Namibia. In The Oil and Gas Habitats of the South Atlantic: Geological Society, ed. N. R. Cameron, R. H. Bate, and V. S. Clure. [London] Special Publication 153, pp. 381402.Google Scholar
Clift, P. D. (1997). The thermal impact of Paleocene magmatic underplating in the Faeroe-Shetland-Rockall region. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, London, UK, pp. 585–93.Google Scholar
Clift, P. D. and Lorenzo, M. (1999). Flexural unloading and uplift along the Cote d’Ivoire Ghana transform margin, Equatorial Atlantic. Journal of Geophysical Research, 104, 25257–74.CrossRefGoogle Scholar
Clift, P. D. and Turner, J. (1998). Paleogene igneous underplating and subsidence anomalies in the Rockall-Faeroe-Shetland area. Marine and Petroleum Geology, 15, 223–43.CrossRefGoogle Scholar
Cloetingh, S. (1986). Intraplate stresses: a new tectonic mechanism for fluctuations of relative sea level. Geology, 14, 617–20.2.0.CO;2>CrossRefGoogle Scholar
Cloetingh, S. and Banda, E. (1992). Europe’s lithosphere: physical properties, mechanical structure. In A Continent Revealed – the European Geotraverse, ed. Blundell, D., Freeman, R., and Mueller, S.. Cambridge University Press, Cambridge, pp. 8091.Google Scholar
Cloetingh, S., Kooi, H., and Groenewoud, W. (1989). Intraplate stresses and sedimentary basin evolution. Geophysical Monograph, 48, 116.Google Scholar
Cloetingh, S., Tankard, A. J., Welsink, H. J., and Jenkins, W. A. M. (1989). Vail’s coastal onlap curves and their correlation with tectonic events, offshore eastern Canada. In Extensional Tectonics and Stratigraphy of the North Atlantic Margins, ed. Tankard, A. J. and Balkwill, H. R.. American Association of Petroleum Geologists Memoir, 46, 282–93.Google Scholar
Cloetingh, S., Zoetemeijer, R., and van Wees, J. D. (1995). Tectonics I. Tectonics and basin formation in convergent settings: thermo-mechanical evolution of the lithosphere and basin evolution in compressive tectonic regimes. Short Course, Vrije University, Amsterdam.Google Scholar
Coakley, B. J. and Watts, A. B. (1991). Tectonic controls on the development of unconformities: the North Slope, Alaska. Tectonics, 10, 101–30.CrossRefGoogle Scholar
Cobbold, P. R., Meisling, K. E., and Mount, V. S. (2001). Reactivation of an obliquely rifted margin, Campos and Santos Basins, southeastern Brazil. American Association of Petroleum Geologists Bulletin, 85, 1925–44.Google Scholar
Cobbold, P. R., Mourgues, R., and Boyd, K. (2004). Mechanism of thin-skinned detachment in the Amazon fan: assessing the importance of fluid overpressure and hydrocarbon generation. Marine and Petroleum Geology, 21, 1013–25.CrossRefGoogle Scholar
Coblentz, D. D., Richardson, R. M., and Sandiford, M. (1994). On the gravitational potential of the Earth’s lithosphere. Tectonics, 13, 929–45.CrossRefGoogle Scholar
Cochran, E. S., Li, Y.-G., and Vidale, J. E. (2006). Anisotropy in the shallow crust observed around the San Andreas Fault before and after the 2004 M 6.0 Parkfield earthquake. Bulletin of the Seismological Society of America, 96, 364–75.CrossRefGoogle Scholar
Cochran, J. R. (1973). Gravity and magnetic investigations in the Guiana Basin, Western Equatorial Atlantic. Geological Society of America Bulletin, 84, 3249–68.2.0.CO;2>CrossRefGoogle Scholar
Cochran, J. R. (1982). The magnetic quiet zone in the eastern Gulf of Aden: implications for the early development of the continental margin. Geophysical Journal of the Royal Astronomical Society, 68, 171201.CrossRefGoogle Scholar
Cochran, J. R. (1983). Effects of finite extension times on the development of sedimentary basins. Earth and Planetary Science Letters, 66, 289302.CrossRefGoogle Scholar
Cochran, J. R. (2005). Northern Red Sea: nucleation of an oceanic spreading center within a continental rift. Geochemistry, Geophysics, Geosystems, 6.CrossRefGoogle Scholar
Cochran, J. R. and Karner, G. D. (2007). Constraints on the deformation and rupturing of continental lithosphere of the Red Sea: the transition from rifting to drifting. In Imaging, Mapping and Modelling Continental Lithosphere Extension and Breakup, ed. Karner, G. D., Manatschal, G., and Pinheiro, L. M.. Special Publication of the Geological Society of London, 282, 265–89.Google Scholar
Cochran, J. R. and Martinez, F. (1988). Evidence from the northern Red Sea on the transition from continental to oceanic rifting. Tectonophysics, 153, 2553.CrossRefGoogle Scholar
Cockings, J. H., Kessler, L. G., Mazza, T. A., and Riley, L. A. (1992). Bathonian to mid-Oxfordian sequence stratigraphy of the South Viking Graben, North Sea. In Exploration Britain: Geological Insights into the Next Decade, ed. Hardman, R. F. P.. Special Publication of the Geological Society of London, 67, 65105.Google Scholar
Cogne, N., Gallagher, K., and Cobbold, P. R. (2011). Post-rift reactivation of the onshore margin of southwest Brazil: evidence from apatite (U-Th)/He and fission-track data. Earth and Planetary Science Letters, 309, 118–30.CrossRefGoogle Scholar
Cohee, G. V., Applin, P. L., Bass, N. W., Bell, A. H., Billings, M. P., DeFord, R. K., Dobbin, C. E., Donnell, J. R., Forrester, J. D., Gilluly, J., Hoots, H. W., Kind, P. B., Longwell, C. R., Lyons, P. L., Murray, G. E., and Waters, A. C. (1962). Tectonic Map of the United States Exclusive of Alaska and Hawaii. AAPG and USGS, Tulsa, Oklahoma.Google Scholar
Cohen, H. A. and McClay, K. R. (1996). Niger Delta shale tectonics. Marine and Petroleum Geology, 13, 313–28.CrossRefGoogle Scholar
Cole, G., Yu, A., Peel, F., Requejo, R., DeVay, J., Brooks, J., Bernard, B., Zumberge, J., and Brown, S. (2001). Constraining source and charge risk in deepwater areas. World Oil Magazine, 222, 6977.Google Scholar
Colella, A. (1988a). Pliocene-Holocene fan deltas and braid deltas in the Crati Basin, southern Italy: a consequence of varying tectonic conditions. In Fan Deltas, ed. Nemec, W. and Steel, R. J.. Blackie, London, pp. 5074.Google Scholar
Colella, A. (1988b). Marine fault-controlled Gilbert-type fan deltas. American Association of Petroleum Geologists Bulletin, 72, 995.Google Scholar
Colletta, B., Le Quellec, P., Letouzey, J., and Moretti, I. (1988). Longitudinal evolution of the Suez rift structure (Egypt). Tectonophysics, 153, 221–33.CrossRefGoogle Scholar
Collier, J. S., Sansom, V., Ishizuka, O., Taylor, R. N., Minshull, T. A., and Whitmarsh, R. B. (2008). Age of Seychelles–India break-up. Earth and Planetary Science Letters, 272, 264–77.CrossRefGoogle Scholar
Collister, J. W., Simmons, R. E., Lichtfouse, E., and Hayes, J. M. (1992). An isotopic biochemical study of the Green River oil shale. Organic Geochemistry, 19, 265–76.CrossRefGoogle Scholar
Colwell, J. B., Symonds, P. A., and Crawford, A. J. (1994). The nature of the Walaby (Curvier) Plateau and other igneous provinces of the West Australian Margin. In Geology of the Outer North West Shelf, Australia: Australian Geological Survey Organization, ed. N. Exon, AGSO Journal of Australian Geology and Geophysics, 15(1), 137–56.Google Scholar
Combarnous, M. H. and Bories, S. A. (1975). Hydrothermal convection in saturated porous media. Advances in Hydrosciences, 10, 231307.CrossRefGoogle Scholar
Combellas-Biggot, R. I. and Galloway, W. E. (2002). Depositional history and genetic sequence stratigraphic framework of the middle Miocene depositional episode south Louisiana. Annual Meeting Expanded Abstracts – American Association of Petroleum Geologists, p. 33.Google Scholar
Conceicao, J. C. de J., Zalán, P. V., and Wolff, S. (1988). Mecanismo, evolucao e cronologia do rift Sul-Atlantico. Mechanism, evolution and chronology of South Atlantic rifting. Boletim de Geociencias da PETROBRAS, 2, 255–65.Google Scholar
Coney, D., Fyfe, T. B., Retail, P., and Smith, P. J. (1993). Clair appraisal: the benefits of a co-operative approach. In Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference, ed. Parker, J. R.. The Geological Society of London, pp. 1409–20.Google Scholar
Coney, P. J. (1980). Cordilleran metamorphic core complexes: an overview. Geological Society of America Memoir, 153, 731.CrossRefGoogle Scholar
Connan, J. (1984). Biodegradation of crude oils in reservoirs. In Advances in Petroleum Geochemistry, 1, ed. Brooks, J. and Welte, D.. London, Academic Press, pp. 299335.CrossRefGoogle Scholar
Connan, J. (1993). Origin of severely biodegraded oils: a new approach using biomarker pattern of asphaltene pyrolysates. In Applied Petroleum Geochemistry, ed. Bordenave, M. L.. Technip, Paris, pp. 457–63.Google Scholar
Connolly, D. L. and Aminzadeh, F. (2003). Exploring for deep gas in the Gulf of Mexico shelf and deepwater using gas chimney processing. Gulf Coast Association of Geological Societies Transactions, 53, 135–42.Google Scholar
Connolly, D. L. and Aminzadeh, F. (2006). Prospect risking using 3-D seismic derived gas chimney volumes. Abstracts, Annual Meeting – American Association of Petroleum Geologists, 15, 21.Google Scholar
Connolly, D. L., Aminzadeh, F., and Ligtenberg, H. (2004). Reducing seal and charge risk by detecting fluid migration pathways using seismic data. Abstracts, Annual Meeting – American Association of Petroleum Geologists, 13, 27.Google Scholar
Connolly, D. L., Fraser, B., and Aminzadeh, F. (2007). Distinguishing static and dynamic faults: application for Gulf of Mexico deep gas play. Abstracts, Annual Meeting – American Association of Petroleum Geologists, 27.Google Scholar
Connolly, D. L., Sawyer, R. K., Aminzadeh, F., de Groot, P. F. M., and Ligtenberg, H. J. (2002). Gas chimney processing as a new exploration tool: a West Africa example. Abstracts, Annual Meeting – American Association of Petroleum Geologists, 34.Google Scholar
Connolly, J. A. D. and Podladchikov, Y. Y. (2000). Temperature- dependent viscoelastic compaction and compartmentalization in sedimentary basins. Tectonophysics, 324, 137–68.CrossRefGoogle Scholar
Connolly, P. and Cosgrove, J. (1999). Prediction of fracture-induced permeability and fluid flow in the crust using experimental stress data. American Association of Petroleum Geologists Bulletin, 83, 757–77.Google Scholar
Constenius, K. N. (1996). Late Paleogene extensional collapse of the Cordilleran foreland fold and thrust belt. Geological Society of America Bulletin, 108, 2039.2.3.CO;2>CrossRefGoogle Scholar
Contrucci, I., Matias, L., Moulin, M., Geli, L., Klingelhofer, F., Nouze, H., Aslanian, D., Olivet, J. L., Rehault, J. P., and Sibuet, J. C. (2004). Deep structure of the West African continental margin (Congo, Zaire, Angola), between 5u S and 8u S, from reflection/refraction seismics and gravity data. Geophysical Journal International, 158, 529–53.CrossRefGoogle Scholar
Cook, D. R. (1987). The Goban Spur: exploration in a deep water frontier basin. In Petroleum Geology of North West Europe, ed. Brooks, J. and Glennie, K. W.. Graham and Trotman, London.Google Scholar
Coolbaugh, M., Blackwell, D., and Richards, M. (2005). Temperature gradient map of the Great Basin. Great Basin Center for Geothermal Energy, University of Nevada, Reno. www.unr.edu/Geothermal/.Google Scholar
Cooper, B. S. and Barnard, P. C. (1984). Source rocks and oils of the central and northern North Sea. In Petroleum Geochemistry and Basin Evaluation, ed. Demaison, G. and Murris, R. J.. American Association of Petroleum Geologists Memoir, 35, 303–14.Google Scholar
Cooper, H. H. (1966). The equation of groundwater flow in fixed and deforming coordinates. Journal of Geophysical Research, 71, 4785–90.CrossRefGoogle Scholar
Copestake, P., Sims, A., Crittenden, S., Hamar, G., Ineson, J., Rose, P., and Tringham, M. (2002). Chapter 12. Lower Cretaceous. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. Geological Society of London, pp. 441–88.Google Scholar
Copestake, P., Sims, A., Crittenden, S., Hamar, G., Ineson, J., Rose, P., and Tringham, M. (2003). Chapter 12. Lower Cretaceous. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. Geological Society of London, pp. 191211.Google Scholar
Corey, A. T. (1986). Mechanics of Immiscible Fluids in Porous Media. Water Resources Publications, Littleton.Google Scholar
Cornea, I., Rădulescu, F. L., and Sova, A. (1981). Deep seismic sounding in Romania. Pure and Applied Geophysics, 119, 1144–56.CrossRefGoogle Scholar
Corner, B., Cartwright, J., and Swart, R. (2002). Volcanic passive margin of Namibia: a potential fields perspective. In Volcanic Rift Margins, ed. M. A. Menzies, S. L. Klemperer, C. Ebinger, and J. Baker. Geological Society of America Special Paper, 362, pp. 203–20.Google Scholar
Cornford, C. (1994). The Mandal-Ekofisk(!) petroleum system in the central graben of the North Sea. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 537–71.Google Scholar
Cornford, C. (1998). Source rock and hydrocarbons of the North Sea. In Petroleum Geology of the North Sea, Basic Concepts and Recent Advances (fourth edition), ed. Glennie, K. W.. Blackwell Scientific Publications, Oxford, pp. 376462.CrossRefGoogle Scholar
Cornford, C., Bray, R., and Kieft, R. (2005). Deepwater source rocks – accumulation and efficiencies. The 4th HGS/PESGB International Conference on African E7P, Houston Program, 4.Google Scholar
Cornford, C., Needham, C. E. J., and de Walque, L. (1986). Geochemical habitat of North Sea oils and gases. In Habitat of Hydrocarbons on the Norwegian Continental Shelf, ed. Spencer, A. M.. Graham and Trotman, London, pp. 3954.Google Scholar
Corti, G., Bonini, M., Conticelli, S., Innocenti, F., Manetti, P., and Sokoutis, D. (2003). Analogue modelling of continental extension: a review focused on the relations between the patterns of deformation and the presence of magma. Earth-Science Reviews, 63, 169247.CrossRefGoogle Scholar
Cosgrove, J. W. (2001). Hydraulic fracturing during the formation and deformation of a basin: a factor in the dewatering of low-permeability sediments. American Association of Petroleum Geologists Bulletin, 85, 737–48.Google Scholar
Coterill, K., Tari, G., Molnar, J., and Ashton, P. R. (2002). Comparison of depositional sequences and tectonic styles among the West African deepwater frontiers of western Ivory Coast, southern Equatorial Guinea, and northern Namibia. The Leading Edge, 1103–11.CrossRefGoogle Scholar
Courtillot, V., Devaille, A., Besse, J., and Stock, J. (2003). Three distinct types of hotspots in the earth’s mantle. Earth and Planetary Science Letters, 205, 295–308.CrossRefGoogle Scholar
Coutts, S. D., Larsson, S. Y., and Rosman, R. (1996). Development of the slumped crestal area of the Brent reservoir, Brent Field: an integrated approach. Petroleum Geoscience, 2, 219–29.CrossRefGoogle Scholar
Coward, R. N., Clark, N. M., and Pinnock, S. J. (1991). The Tartan field, Block 15/16, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 377–84.Google Scholar
Cowie, P. A. (1998). A healing-reloading feedback control on the growth rate of seismogenic faults. Journal of Structural Geology, 20, 1075–87.CrossRefGoogle Scholar
Cowie, P. A., Gupta, S., and Dawers, N. H. (2000). Implications of fault array evolution for synrift depocentre development: insights from a numerical fault growth model. Basin Research, 12, 241–61.CrossRefGoogle Scholar
Cowie, P. A. and Scholz, C. H. (1992). Physical explanation for the displacement-length relationship of faults using a post-yield fracture mechanics model. Journal of Structural Geology, 14, 1133–48.CrossRefGoogle Scholar
Cowley, R. and O’Brien, G. W. (2000). Identification and interpretation of leaking hydrocarbons using seismic data: a comparative montage of examples from the major fields in Australia’s North West Shelf and Gippsland Basin. APPEA Journal, 40, 121–50.CrossRefGoogle Scholar
Cox, K. G. (1980). A model for flood basalt volcanism. Journal of Petrology, 21, 629–50.CrossRefGoogle Scholar
Cox, S. J. D., Meredith, P. G., and Stuart, C. E. (1991). Microfracturing during brittle rock failure: a model for the Kaiser effect including sub-critical crack growth. Seventh International Congress on Rock Mechanics, pp. 703–7.Google Scholar
Coyle, B. J. and Zoback, M. D. (1988). In situ permeability and fluid pressure measurements at ~2km depth in the Cajon Pass research well. Geophysical Research Letters, 15, 1029–32.CrossRefGoogle Scholar
Crampin, S. (1990). The scattering of shear waves in the crust. Pure and Applied Geophysics, 132, 6791.CrossRefGoogle Scholar
Craven, J. (2000). Petroleum system of the Ivorian/Tano Basin in the W Tano Contract area, offshore west Ghana. Abstracts Volume, Petroleum Systems and Developing Technologies in African Exploration and Production, Geol. Soc./PESGB Conference, May, 2000.Google Scholar
Creaney, S. and Allan, J. (1992). Petroleum systems in the foreland basin of western Canada. In Foreland Basin and Fold Belts, ed. Macqueen, R. W. and Leckie, D. A.. American Association of Petroleum Geologists Memoir, 55, 279308.Google Scholar
Creaney, S. and Passey, Q. R. (1993). Recurring patterns of total organic carbon and source rock quality within a sequence stratigraphic framework. American Association of Petroleum Geologists Bulletin, 77, 386401.Google Scholar
Crider, J. G. and Pollard, D. D. (1998). Fault linkage: three-dimensional mechanical interaction between echelon normal faults. Journal of Geophysical Research, 103, 24373–91.CrossRefGoogle Scholar
Criss, R. E. and Hofmeister, A. M. (1991). Application of fluid dynamics principles in tilted permeable media to terrestrial hydrothermal systems. Geophysical Research Letters, 18, 199202.CrossRefGoogle Scholar
Crittenden, M. D., Coney, P. J., and Davis, G. H. (1980). Cordilleran metamorphic core complexes. Geological Society of America Memoir, 153, 490.Google Scholar
Crossley, R. (1984). Controls of sedimentation in the Malawi Rift Valley, central Africa. Sedimentary Geology, 40, 3350.CrossRefGoogle Scholar
Crowell, J. C. (1974a). Sedimentation along the San Andreas Fault, California. In Modern and Ancient Geosynclinal Sedimentation, ed. Dott, R. H. and Shaver, R. H.. Society of Economic Sedimentologists and Mineralogists Special Publications, 19, 292303.CrossRefGoogle Scholar
Crowell, J. C. (1974b). Origin of late Cenozoic basins in southern California. Society of Economic Sedimentologists and Mineralogists Special Publications, 22, 190204.Google Scholar
Crowell, J. C. (1982). The tectonics of the Ridge Basin, southern California. In Geologic History of the Ridge Basin, ed. Crowell, J. C. and Link, M. H.. Pacific Section of the Society of Economic Geologists, Paleontologists and Mineralogists, pp. 2541.Google Scholar
Crowell, J. C. and Link, M. H. (1982). Geologic History of Ridge Basin, southern California. Los Angeles, Pacific Section, Society of Economic Paleontologists and Mineralogists, p. 304.Google Scholar
Crowley, K. D., Cameron, M., and Schaefer, R. L. (1991). Experimental studies of annealing of etched fission tracks in fluorapatite. Geochimica et Cosmochimica Acta, 55, 1449–65.CrossRefGoogle Scholar
Curiale, J. A. (1994). Correlation of oils and source rocks: a conceptual and historical perspective. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 251–60.Google Scholar
Currey, D. R., Oviatt, C. G., and Czarnomski, J. E. (1984). Late Quaternary geology of Lake Bonneville and Lake Waring. In Geology of Northwest Utah, Southern Idaho and Northeast Nevada, ed. Kerns, G. J. and Kerns Jr, R. L.. Utah Geological Association Publication, 13, 227–37.Google Scholar
Currie, S. (1996). The development of the Ivanhoe, Rob Roy and Hamish Fields, Block 15/21a, UK North Sea. In Geology of the Humber Group: Central Graben and Moray Firth, UKCS, ed. Hurst, A., Johnson, H. D., Burley, S. D., Canham, A. C., and Mackertich, D. S.. Special Publication of Geological Society of London, 114, 329–41.Google Scholar
Currie, S., Gowland, S., Taylor, A., and Woodward, M. (1999). The reservoir development of the Fife Field. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A.. Proceedings of the 5th Conference, Geological Society, London, pp. 1135–46.Google Scholar
Curtis, C. D. (1983). The link between aluminum mobility and destruction of secondary porosity. American Association of Petroleum Geologists Bulletin, 67, 380–4.Google Scholar
Curtis, J. B., Kotarba, M. J., Lewan, M. D., and Wieclaw, D. (2004). Oil/source rock correlations in the Polish Flysch Carpathians and Mesozoic basement and organic facies of the Oligocene Menilite Shales: insights from hydrous pyrolysis experiments. Organic Geochemistry, 35, 1573–96.CrossRefGoogle Scholar
d’Acremont, E., Leroy, S., and Burov, E. B. (2003). Numerical modelling of a mantle plume: the plume head-lithosphere interaction in the formation of an oceanic large igneous province. Earth and Planetary Science Letters, 206, 379–96.CrossRefGoogle Scholar
D’Agostino, N., Giuliani, R., Mattone, M., and Bonci, L. (2001). Active crustal extension in the Central Apennines (Italy) inferred from GPS measurements in the interval 1994–1999. Geophysical Research Letters, 28, 2121–4.Google Scholar
Dahl, J. E. P., Moldowan, J. M., Teerman, S. C., McCaffery, M. A., Sundararaman, P., and Stelting, C. E. (1994). Source rock quality determination from oil biomarkers I: A new geochemical technique. American Association of Petroleum Geologists Bulletin, 78, 1507–26.Google Scholar
Dahl, N. and Solli, T. (1993). The structural evolution of the Snorre Field and surrounding areas. In Petroleum Geology of Northwest Europe, ed. Parker, J. R.. Proceedings of the 4th Conference, Geological Society of London, pp. 1159–66.Google Scholar
Dahlberg, E. C. (1995). Applied Hydrodynamics in Petroleum Exploration. Springer-Verlag, New York.CrossRefGoogle Scholar
Dailly, P. (2000). Tectonic and stratigraphic development of the Rio Muni Basin, Equatorial Guinea: the role of transform zones in Atlantic Basin evolution. Geophysical Monograph, 115, 105–28.Google Scholar
Dailly, P., Lowry, P., Goh, K., and Monson, G. (2002). Exploration and development of Ceiba Field, Rio Muni Basin, southern Equatorial Guinea. The Leading Edge, 1140–6.CrossRefGoogle Scholar
Daly, R. A., Manger, G. E., and Clark, S. P. Jr. (1966). Density of rocks. Handbook of Physical Constants, 4, 1926.CrossRefGoogle Scholar
Damtoft, K., Nielsen, L. H., Johannessen, P. N., Thomsen, E., and Andersen, P. R. (1992). Hydrocarbon plays of the Danish Central Trough. In Generation, Accumulation and Production of Europe’s Hydrocarbons I., ed. Spencer, A. M.. Special Publication of the European Association of Petroleum Geoscientists, 2, 3558.Google Scholar
Damuth, J. E. (1994). Neogene gravity tectonics and depositional processes on the deep Niger Delta continental margin. Marine and Petroleum Geology, 11, 320–46.CrossRefGoogle Scholar
Darcy, J. (1856). Les Fontaines Publiques de la Ville de Dijon. Dalmont, Paris.Google Scholar
Dart, C. J., Collier, R. E. L., Gawthorpe, R. L., Keller, J. V. A., and Nichols, G. (1994). Sequence stratigraphy of (?) Pliocene-Quaternary synrift, Gilbert-type fan deltas, northern Peloponnesos, Greece. Marine and Petroleum Geology, 11, 545–60.CrossRefGoogle Scholar
Davidson, G. R., Bassett, R. L., Hardin, E. L., and Thompson, D. L. (1998). Geochemical evidence of preferential flow of water through fractures in unsaturated tuff, Apache Leap, Arizona. Applied Geochemistry, 13, 185–95.CrossRefGoogle Scholar
Davies, G. F. (1994). Thermomechanical erosion of the lithosphere by mantle plumes, J. Geophys. Res., 99(B8), 15709–22.Google Scholar
Davies, P., Williams, A. T., and Bomboe, P. (1991). Numerical modelling of Lower Lias rock failures in the coastal cliffs of South Wales. In Coastal Sediments ’91: Proceedings of a Special Conference on Quantitative Approaches to Coastal Processes, ed. Kraus, N. C., Gingerich, K. J., and Kriebel, D. L.. American Society of Civil Engineers, New York, pp. 1599–612.Google Scholar
Davies, R., England, P., Parsons, B., Billiris, H., Paradissis, D., and Veis, G. (1997). Geodetic strain of Greece in the interval 1892–1992. Journal of Geophysical Research, 102, 24571–88.CrossRefGoogle Scholar
Davies, S. J., Dawers, N. H., McLeod, A. E., and Underhill, J. R. (2000). The structural and sedimentological evolution of early synrift successions: the Middle Jurassic Tarbert Formation, North Sea. Basin Research, 12, 343–65.CrossRefGoogle Scholar
Davis, G. A. (1980). Problems of Intraplate Extensional Tectonics, Western United States. National Academy of Sciences, Washington, DC.Google Scholar
Davis, G. A., Anderson, J. L., Frost, E. G., and Shackelford, T. J. (1980). Mylonitization and detachment faulting in the Whipple-Buckskin-Rawhide Mountains terrane, southeastern California and western Arizona. Geological Society of America, 153, p. 79.CrossRefGoogle Scholar
Davis, G. H. (1983). Shear-zone model for the origin of metamorphic core complexes. Geology, 11, 342–7.2.0.CO;2>CrossRefGoogle Scholar
Davis, M. and Kusznir, N. (2002). Are buoyancy forces important during the formation of rifted margins? Geophysical Journal International, 149, 524–33.CrossRefGoogle Scholar
Davis, M. and Kusznir, N. (2004). Depth-dependent lithospheric stretching at rifted continental margins. In Rheology and Deformation of the Lithosphere at Continental Margins MARGINS Theoretical and Experimental Earth Science Series, ed. Karner, G. D., Taylor, B., Driscoll, N. W., and Kohlstedt, D. L.. Columbia University Press, New York, pp. 92137.Google Scholar
Davis, P. J., McKenzie, J. A., Palmer-Julson, A., and colleagues (1991). Explanatory notes. Proceedings of the Ocean Drilling Program, Initial Reports, 133, 3158.Google Scholar
Davis, R. W. (1991). Integration of geological data into hydrodynamic analysis of hydrocarbon movement. In Petroleum Migration, ed. England, W. A. and Fleet, A. J.. Geological Society of London Special Publications, 59, 127–35.Google Scholar
Davison, I. (1989). Extensional domino fault tectonics: kinematics and geometrical constraints. Annales Tectonicae, 3, 1224.Google Scholar
Davison, I. (1999). Tectonics and hydrocarbon distribution along the Brazilian South Atlantic margin. In The Oil and Gas Habitats of the South Atlantic, ed. Cameron, N. R., Bate, R. H., and Clure, V. S.. Geological Society of London Special Publications, 153, 133–51.Google Scholar
Davison, I. (2005). Central Atlantic margin basins of northwest Africa: geology and hydrocarbon potential (Morocco to Guinea). Journal of African Earth Sciences, 43, 254–74.CrossRefGoogle Scholar
Davison, I., Al-Kadasi, M., Al-Khirbash, S., Al-Subbary, A., Baker, J., Blakely, S., Bosence, D., Dart, C., Owen, L., Menzies, M., McClay, K., Nichols, G., and Yelland, A. (1994). Geological evolution of the Southern red Sea rift margin-republic Yemen. Geological Society of American Bulletin, 106, 1474–93.2.3.CO;2>CrossRefGoogle Scholar
Davison, I., Taylor, M., and Longacre, M. (2002). Northwest Africa – Iberia – USA – Canada. Central Atlantic reconstruction, Scale 1: 5 000 000, Earthmoves, Ltd. Richmond International Geoscience, MBL, Inc.Google Scholar
Davy, P. (1986). Modélisation thermo-mécanique de la collision continentale. University of Rennes, Rennes.Google Scholar
Davy, P. and Cobbold, P. R. (1991). Experiments on shortening of a 4-layer model of the continental lithosphere. Tectonophysics, 125.CrossRefGoogle Scholar
Daw, G. P., Howell, F. T., and Woodward, F. A. (1974). The effect of applied stress upon the permeability of some Permian and Triassic sandstones of northern England: advances in rock mechanics. Proceedings of the Third International Congress of Rock Mechanics, II, 537–42.Google Scholar
Dawers, N. H. and Anders, M. H. (1995). Displacement-length scaling and fault linkage. Journal of Structural Geology, 17, 607–14.CrossRefGoogle Scholar
Dawers, N. H., Berge, A. M., Haeger, K. O., Puigdefabregas, C., and Underhill, J. R. (1999). Controls on Late Jurassic, subtle sand distribution in the Tampen Spur area, northern North Sea. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of 5th conference, London, UK, pp. 827–38.Google Scholar
Dawers, N. H. and Underhill, J. R. (2000). The role of fault interaction and linkage in controlling synrift stratigraphic sequences: Late Jurassic, Statfjord East area, northern North Sea. American Association of Petroleum Geologists Bulletin, 84, 4564.Google Scholar
de Almeida, F. F. M. (1976). The system of continental rifts bordering the Santos Basin, Brazil. Anais da Academia Brasileira de Ciencias, 48, 1526.Google Scholar
De’Ath, N. G. and Schuyleman, S. F. (1981). The geology of the Magnus oilfield. In Petroleum Geology of the Continental Shelf of North-West Europe, ed. Illing, L. V. and Hobson, G. D.. Heyden and Son, London, pp. 342–51.Google Scholar
de Azevedo, R. P. (1991). Tectonic evolution of Brazilian equatorial continental margins basins. Thesis, Imperial College, London, p. 455.Google Scholar
de Boer, J. Z. (1992). Stress configurations during and following emplacement of ENA basalts in the northern Appalachians. Special Paper – Geological Society of America, 268, 361–78.CrossRefGoogle Scholar
de Boer, J. Z. and Clifton, A. E. (1988). Mesozoic tectogenesis: development and deformation of “Newark” rift zones in the Appalachians (with special emphasis on the Hartford Basin, Connecticut). In Triassic–Jurassic Rifting, Continental Breakup and the Origin of the Atlantic Ocean Passive Margins, Part A, ed. Manspeizer, W.. Elsevier, New York, pp. 275306.CrossRefGoogle Scholar
De Bremaeker, J. C. (1983). Temperature, subsidence and hydrocarbon maturation in extensional basins: a finite element model. American Association of Petroleum Geologists Bulletin, 67, 1410–14.Google Scholar
de Charpal, O., Guennoc, P., Montadert, L., and Roberts, D. G. (1978). Rifting, crustal attenuation and subsidence in the Bay of Biscay. Nature, 275, 706–11.CrossRefGoogle Scholar
de Graciansky, P. C., Dardeau, G., Lemoine, M., and Tricart, P. (1989). The inverted margin of the French Alps and foreland basin inversion. In Inversion Tectonics, ed. Cooper, M. A. and Williams, G. D.. Geological Society of London Special Publications, 44, 87104.Google Scholar
de Graciansky, P. C., Poag, C. W., and Foss, G. (1985). Drilling on the Goban Spur: objectives, regional geological setting, and operational summary. Initial Reports of the Deep Sea Drilling Project, 80, 513.Google Scholar
de Graciansky, P. C., Poag, C. W., Cunningham, R. and colleagues (1985). Site 550. Initial Reports of the Deep Sea Drilling Project, 80, 251356.Google Scholar
De-Hua, H. and Batzle, M. (2006). Velocities of deepwater reservoir sands. The Leading Edge, pp. 260–6.Google Scholar
De Marsily, G. (1981). Hydrogeologie quantitative. Paris, Mason.Google Scholar
de Matos, R. M. D. (1987). Sistema de rifts Cretáceos do Nordeste Brasiliero, Anais, Tectos. Petrobrás-Depex, Rio de Janeiro, Brazil, pp. 126–59.Google Scholar
de Matos, R. M. D. (1992). The Northeast Brazilian rift system. Tectonics, 11, 766–91.CrossRefGoogle Scholar
de Matos, R. M. D., de Lima Neto, F. F., Alves, A. C., and Waick, R. N. (1987). O rift Potiguar, gênese, preenchimento e acumulações de hidrocarbonetos, Anais, Rifts intracontinentais. Petrobrás-Depex, Rio de Janeiro, Brazil, pp. 160–77.Google Scholar
Dean, S. M., Minshull, T. A., Whitmarsh, R. B., and Louden, K. E. (2000). Deep structure of the ocean–continent transition in the southern Iberia abyssal plain from seismic refraction profiles: the IAM-9 transect at 40 degrees 20’N. Journal of Geophysical Research, 105, 5859–85.CrossRefGoogle Scholar
DeCarlo, E. H., McMurtry, G. M., and Yeh, H.-W. (1983). Geochemistry of hydrothermal deposits from Loihi submarine volcano, Hawaii. Earth and Planetary Science Letters, 66, 438–49.Google Scholar
DeCelles, P. G., Gray, M. B., Ridgway, K. D., Cole, R. B., Srivastava, P., Pequera, N., and Pivnik, D. A. (1991). Kinematic history of a foreland uplift from Paleocene synorogenic conglomerate, Beartooth Range, Wyoming and Montana. Geological Society of America Bulletin, 103, 1458–75.2.3.CO;2>CrossRefGoogle Scholar
Decker, K., Peresson, H., and Hinsch, R. (2005). Active tectonics and Quaternary basin formation along the Vienna Basin transform fault. Quaternary Science Reviews, 24, 305–20.CrossRefGoogle Scholar
Deenen, M. H. L., Ruhl, M., Boris, N. L., Krijgsman, W., Kuerschner, W. M., Reitsma, M., and van Bergen, M. J. (2010). A new chronology for the End-Triassic Mass Extinction. Earth and Planetary Science Letters, 291, 113–25.CrossRefGoogle Scholar
Dehler, S. A. (2010). Initial rifting and break-up between Nova Scotia and Morocco: an examination of new geophysical data and models. In Conjugate Margins II, Lisbon 2010, Metedo Directo, VIII, 79–82. http://metedodirecto.pt/CM2010.Google Scholar
Dehler, S. A. and Welford, J. K. (2012). Variations in rifting style and structure of the Scotian margin, Atlantic Canada, from 3D gravity inversion. In Conjugate Divergent Margins, ed. Mohriak, W. U., Danforth, A., Post, P. J., Brown, D. E., Tari, G. M, Nemčok, M., and Sinha, S. T.. Geological Society of London Special Publications, 369, 289–300.Google Scholar
Demaison, G. J., Holck, A. J. J., Jones, R. W., and Moore, G. T. (1983). Predictive source bed stratigraphy: a guide to regional petroleum occurrences. Proceedings of the 11th World Petroleum Congress, London, pp. 113.Google Scholar
Demaison, G. J. and Huizinga, B. J. (1991). Genetic classification of petroleum systems. American Association of Petroleum Geologists Bulletin, 75, 1626–43.Google Scholar
Dembicki, Jr. H. and Anderson, M. L. (1989). Secondary migration of oil experiments supporting efficient movement of separate, buoyant oil phase along limited conduits. American Association of Petroleum Geologists Bulletin, 73, 1018–21.Google Scholar
Demercian, L. S. (1996). A halocinese na evolucao do Sul da Bacia de Santos do Aptiano ao Cretaceo Superior. MS thesis, Universidade Federal do Rio Grande do Sul, Porto Alegre, Brazil, p. 201.Google Scholar
Demercian, S., Szatmari, P., and Cobbold, P. R. (1993). Style and pattern of salt diapirs due to thin-skinned gravitational gliding, Campos and Santos basins, offshore Brazil. Tectonophysics, 228, 393433.CrossRefGoogle Scholar
Demetrescu, C. and Veliciu, S. (1991). Heat flow and lithosphere structure in Romania. In Terrestrial Heat Flow and Lithospheric Structure, ed. Čermák, V. and Rybach, L.. Springer-Verlag, Berlin, pp. 187205.CrossRefGoogle Scholar
Deming, D. (1994a). Fluid flow and heat transport in the upper continental crust. In Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins, ed. Parnell, J.. Geological Society of London Special Publications, 78, 2742.Google Scholar
Deming, D. (1994b). Overburden rock, temperature, and heat flow. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 165–86.Google Scholar
Deming, D. (1994c). Factors necessary to define a pressure seal. Bulletin of the American Association of Petroleum Geologists, 78, 1005–9.Google Scholar
Deming, D. and Chapman, D. S. (1989). Thermal histories and hydrocarbon generation: example from Utah-Wyoming thrust belt. American Association of Petroleum Geologists Bulletin, 73, 1455–71.Google Scholar
Deming, D. and Nunn, J. A. (1991). Numerical simulations of brine migration by topographically driven recharge. Journal of Geophysical Research, 96, 2485–99.CrossRefGoogle Scholar
Deming, D., Nunn, J. A., and Evans, D. G. (1990a). Thermal effects of compaction-driven groundwater flow from overthrust belts. Journal of Geophysical Research, 95, 6669–83.CrossRefGoogle Scholar
Deming, D., Nunn, J. A., Jones, S., and Chapman, D. S. (1990b). Some problems in thermal history studies. In Applications of Thermal Maturity Studies to Energy Exploration, ed. Nuccio, F. and Barker, C. E.. Rocky Mountain Section – Society of Exploration Paleontologists and Mineralogists, 6180.Google Scholar
Densmore, A. L., Ellis, M. A., and Anderson, R. S. (1998). Landsliding and the evolution of normal fault-bounded mountains. Journal of Geophysical Research, 103, 15203–19.CrossRefGoogle Scholar
Deptuck, M. E. (2011). Proximal to distal postrift structural provinces of the western Scotian Margin, offshore Eastern Canada. CNSOPB Geoscience Open File Report, 2011-001MF, p. 42.Google Scholar
Desheng, L. (1980). Geological structure and hydrocarbon occurrence of the Bohai Gulf oil and gas basin (China). In Petroleum Geology in China, ed. Mason, J. F.. Penn Well Books, Tulsa, pp. 180–92.Google Scholar
Desmurs, L., Manatschal, G., and Bernoulli, D. (2001). The Steinmann Trinity revisited: mantle exhumation and magmatism along an ocean–continent transition: the Platta Nappe, eastern Switzerland. Geological Society of Special Publications, 187, 235–66.CrossRefGoogle Scholar
Destro, N., Alkmim, F. F., Magnavita, L. P., and Szatmari, P. (2003a). The Jeremoabo transpressional transfer fault, Reconcavo-Tucano Rift, NE Brazil. Journal of Structural Geology, 25, 1263–79.CrossRefGoogle Scholar
Destro, N., Szatmari, P., Alkmim, F. F., and Magnavita, L. P. (2003b). Release faults, associated structures, and their control on petroleum trends in the Reconcavo rift, northeast Brazil. American Association of Petroleum Geologists Bulletin, 87, 1123–44.CrossRefGoogle Scholar
Devey, C. W. and Shipboard Scientific Party (2002). Hydrothermal Studies of Grimsey Field, Volcanic Studies of Kolbeinsey Ridge. University of Bremen.Google Scholar
Devlin, W. J., Cogswell, J. M., Gaskins, G. M., Isaksen, G. H., Pitcher, D. M., Puls, D. P., Stanley, K. O., and Wall, G. R. T. (1999). South Caspian Basin: young, cool, and full of promise. GSA Today, 9, 19.Google Scholar
Dewhurst, D. N., Brown, K. M., Clennell, M. B., and Westbrook, G. K. (1996a). A comparison of the fabric and permeability anisotropy of consolidated and sheared silty clay. Engineering Geology, 42, 253–67.CrossRefGoogle Scholar
Dewhurst, D. N., Clennell, M. B., Brown, K. M., and Westbrook, G. K. (1996b). Fabric and hydraulic conductivity of sheared clays. Geotechnique, 46, 761–8.CrossRefGoogle Scholar
D’Heur, M. (1984). Porosity and hydrocarbon distribution in the North Sea chalk reservoirs. Marine and Petroleum Geology, 1, 211–38.Google Scholar
Dholakia, S. K., Aydin, A., Pollard, D. D., and Zoback, M. D. (1998). Fault-controlled hydrocarbon pathways in the Monterey Formation, California. American Association of Petroleum Geologists Bulletin, 82, 1551–74.Google Scholar
Dias-Brito, D. (1987). A Bacia de Campos no Mesocretaceo: uma contribuicao a paleoceanografia do Atlantico sul primitivo. The Campos Basin of the Middle Cretaceous: a contribution to paleo-oceanography of the ancient South Atlantic. Revista Brasileira de Geociências, 17, 162–7.CrossRefGoogle Scholar
Dibblee, T. R. (1980). Geology along the San Andreas fault from Gilroy to Parkfield. In Studies of the San Andreas Fault Zone in Northern California, ed. Streitz, R. and Sherburne, R.. California Division of Mines and Geology, Sacramento, Special Report, 140, 318.Google Scholar
Dickinson, G. (1953). Geological aspects of abnormal reservoir pressures in Gulf Coast, Louisiana. American Association of Petroleum Geologists, 37, 410–32.Google Scholar
Diegel, F. A., Karlo, J. F., Schuster, D. C., Shoup, R. C., and Tauvers, P. R. (1995). Cenozoic structural evolution and tectono-stratigraphic framework of the northern Gulf Coast continental margin. In Salt Tectonics: A Global Perspective, ed. Jackson, M. P. A., Roberts, D. G., and Snelson, S.. American Association of Petroleum Geologists Memoir, 65, 109–51.Google Scholar
Dikenshteyn, G. K., Maksimov, S. P., and Ivanova, T. D. (1982). Tectonics of Oil-Gas Provinces and Regions of the U.S.S.R. (in Russian). Nedra, Moscow, p. 223.Google Scholar
Dinter, D. A. and Royden, L. (1993). Late Cenozoic extension in northeastern Greece; Strymon Valley detachment system and Rhodope metamorphic core complex. Geology, 21, 45–8.2.3.CO;2>CrossRefGoogle Scholar
Discovery 215 Working Group (1998). Deep structure in the vicinity of the ocean-continent 618 transition zone under the southern Iberia Abyssal Plain. Geology, 26, 743–6.Google Scholar
Dmitriyevskiy, A. N., Kireyev, F. A., Bochko, R. A., and Fedorova, T. A. (1993). Hydrothermal origin of oil and gas reservoirs in basement rock of the South Vietnam continental shelf. International Geology Review, 35, 621–30.CrossRefGoogle Scholar
Dodson, M. H. (1973). Closure temperature in cooling geochronological and petrological systems. Contributions to Mineralogy and Petrology, 40, 259–74.CrossRefGoogle Scholar
Doglioni, C. (1993). Some remarks on the origin of foredeeps. In Crustal Controls on the Internal Architecture of Sedimentary Basins, ed. Stephenson, R. A.. Elsevier, Amsterdam, 228, 120.Google Scholar
Doligez, B., Bessi, F., Burrus, J., Ungerer, P., and Chenet, P. Y. (1986). Integrated numerical simulation of the sedimentation heat transfer, hydrocarbon formation and fluid migration in a sedimentary basin: the Themis model. In Thermal Modeling in Sedimentary Basins, ed. Burrus, J.. Technip, Paris, pp. 173–95.Google Scholar
Donath, F. A. (1961). Experimental study of shear failure in anisotropic rocks. Geological Society of America Bulletin, 72, 985–9.CrossRefGoogle Scholar
Donelick, R. A. (1994). Selected Poland samples. Apatite fission track data. Donelick Analytical Report, 76, 1114.Google Scholar
Donelick, R. A., O’Sullivan, P. B., and Ketcham, R. A. (2005). Apatite fission-track analysis. In Low-Temperature Thermochronology: Techniques, Interpretations and Applications, ed. Reiners, P. W. and Ehlers, T. A.. Reviews in Mineralogy and Geochemistry, 58, 4994.CrossRefGoogle Scholar
Donovan, T. J., Friedman, I., and Gleason, J. D. (1974). Recognition of petroleum-bearing traps by unusual isotopic composition of carbonate-cemented surface rocks. Geology, 2, 351–4.2.0.CO;2>CrossRefGoogle Scholar
Doré, A. G., Lundin, E. R., Birkeland, O., Eliassen, P. E., and Jensen, L. N. (1997). The NE Atlantic Margin: implications of late Mesozoic and Cenozoic events for hydrocarbon prospectivity. Petroleum Geosciences, 3, 117–31.CrossRefGoogle Scholar
Dorsey, R. J., Stone, K. A., and Umhoefer, P. J. (1997a). Stratigraphy, sedimentology, and tectonic development of the southeastern Pliocene Loreto Basin, Baja California Sur, Mexico. In Pliocene Carbonates and Related Facies Flanking the Gulf of California, Baja California, Mexico, ed. Johnson, M. E. and Ledesma-Vazquez, J.. Special Paper – Geological Society of America, pp. 83109.Google Scholar
Dorsey, R. J. and Umhoefer, P. J. (2000). Tectonic and eustatic controls on sequence stratigraphy of the Pliocene Loreto Basin, Baja California Sur, Mexico. Geological Society of America Bulletin, 112, 177–99.2.0.CO;2>CrossRefGoogle Scholar
Dorsey, R. J., Umhoefer, P. J., and Falk, P. (1997b). Earthquake clustering and stacked Gilbert-type fan deltas in the Pliocene Loreto Basin, Baja California Sur, Mexico. Geology, 25, 679–82.2.3.CO;2>CrossRefGoogle Scholar
Dorsey, R. J., Umhoefer, P. J., and Renne, P. R. (1995). Rapid subsidence and stacked Gilbert-type fan deltas, Pliocene Loreto Basin, Baja California Sur, Mexico. Sedimentary Geology, 98, 181204.CrossRefGoogle Scholar
Doubre, C. (2004). Structure et mécanisme des segments de rift volcanotectoniques. PhD thesis, Université du Maine, Angers, France, p. 423.Google Scholar
Doubre, C. and Geoffroy, L. (2003). Model for stress inversions and rift-zone development around a hot-spot-related magma centre on the Isle of Skye, Scotland. Terra Nova, 15, 230–7.CrossRefGoogle Scholar
Doust, H. (1990). Petroleum geology of the Niger Delta. In Classic Petroleum Provinces, ed. Brooks, J.. Geological Society Special Publication, 50, 365.Google Scholar
Doust, H. and Omatsola, E. (1990). Niger Delta. In Divergent/Passive Margin Basins, ed. Edwards, J. D. and Santogrossi, P. A.. American Association of Petroleum Geologists Memoir, 48, 201–38.Google Scholar
Dow, W. (1977). Kerogen studies and geological interpretations. Journal of Geochemical Exploration, 7, 7999.CrossRefGoogle Scholar
Downes, H. and Vaselli, O. (1995). The lithospheric mantle beneath the Carpathian-Pannonian region: a review of trace element and isotopic evidence from ultramafic xenoliths. Acta Vulcanologica, 7, 219–29.Google Scholar
Downey, M. W. (1994). Hydrocarbon seal rocks. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 159–64.Google Scholar
Driscoll, N. W., Hogg, J. R., Christie-Blink, N., and Karner, G. D. (1995). Extensional tectonics in the Jeanne d’Arc Basin, offshore Newfoundland: implications for the timing of break-up between Grand Banks and Iberia. In The Tectonics, Sedimentation and Paleoceanography of the North Atlantic Region, ed. Scrutton, R. A.. Special Publications, pp. 128.CrossRefGoogle Scholar
Driscoll, N. W. and Karner, G. D. (1998). Lower crustal extension across the northern Carnarvon Basin, Australia: evidence for an eastward dipping detachment. Journal of Geophysical Research, 103, 4975–91.CrossRefGoogle Scholar
Droz, L., Marsset, T., Onreas, H., Lopez, M., Savoye, B., and Spy-Anderson, F.-L. (2003). Architecture of an active mud-rich turbidite system: the Zaire Fan (Congo-Angola margin Southeast Atlantic): results from ZaiAngo 1 and 2 cruises. American Association of Petroleum Geologists Bulletin, 87, 1145–68.CrossRefGoogle Scholar
Drucker, D. C. and Prager, W. (1952). Soil mechanics and plastics analysis for limited design. Quarterly of Applied Mathematics, 10, 157–65.CrossRefGoogle Scholar
Drummond, B. J. (1988). A review of crust/upper mantle structure in the Precambrian areas of Australia and implications for Precambrian crustal evolution. Precambrian Research, 4041, 101–16.Google Scholar
Du, K. E., Pai, S., Brown, J., Moore, R. M., and Simmons, M. (2000). Optimising the development of Blake field under tough economic and environmental conditions. Society of Petroleum Engineers, SPE 64714.Google Scholar
Du Rouchet, J. (1981). Stress fields. A key to oil migration. American Association of Petroleum Geologists Bulletin, 65, 7485.Google Scholar
Duan, Z., Moller, N., and Weare, J. H. (1996). A general equation of state for supercritical fluid mixtures and molecular dynamics simulation of mixture PVTX properties. Geochimica et Cosmochimica Acta, 60, 1209–16.CrossRefGoogle Scholar
Duddy, I. R., Green, P. F., Bray, R. J., and Hegarty, K. A. (1994). Recognition of the thermal effects of fluid flow in sedimentary basins. In Origin, Migration and Evolution of Fluids in Sedimentary Basins, ed. Parnell, J.. Geological Society Special Publication, 78, 325–45.CrossRefGoogle Scholar
Duddy, I. R. and Kelly, P. R. (1999). Uranium in mineral sands: measurement and uses. Australian Institute of Geoscientists Bulletin, 26, 4.Google Scholar
Duffy, C. J. and Al-Hassan, S. (1988). Groundwater circulation in a closed desert basin: topographic scaling and climatic forcing. Water Resources Research, 24, 1675–88.CrossRefGoogle Scholar
Dumitru, T. A., Hill, K. C., Coyle, D. A., Duddy, I. R., Foster, D. A., Gleadow, A. J. W., Green, P. F., Kohn, B. P., Laslett, G. M., and O’Sullivan, A. J. (1991). Fission track thermochronology: application to continental rifting of south-eastern Australia. APEA Journal, 31, 131–42.Google Scholar
Dupre, S., Bertotti, S., and Cloetingh, S. A. P. L. (2007). Tectonic history along the South Gabon Basin: anomalous early post-rift subsidence. Marine Petroleum Geology, 24, 151–72.CrossRefGoogle Scholar
Durand, B. (1988). Understanding of HC migration in sedimentary basins (present state of knowledge). Organic Geochemistry, 13, 445–59.CrossRefGoogle Scholar
Duranti, D. and Hurst, A. (2004). Fluidisation and injection in the deep-water sandstones of the Eocene Alba Formation (UK North Sea). Sedimentology, 51, 503–29.CrossRefGoogle Scholar
Durney, D. W. and Ramsay, J. G. (1973). Incremental strains measured by syntectonic crystal growths. In Gravity and Tectonics, ed. de Jong, K. A. and Scholten, R.. Wiley, New York, 6796.Google Scholar
Durrheim, R. J. and Mooney, W. D. (1991). Archean and Proterozoic crustal evolution: evidence from crustal seismology. Geology, 19, 606–9.2.3.CO;2>CrossRefGoogle Scholar
Durrheim, R. J. and Mooney, W. D. (1992). Archean and Proterozoic crustal evolution: evidence from crustal seismology. Geology, 20, 665–6.Google Scholar
Durrheim, R. J. and Mooney, W. D. (1994). Evolution of the Precambrian lithosphere: seismological and geochemical constraints. Journal of Geophysical Research, 99, 15359–74.CrossRefGoogle Scholar
Dydik, B. M., Simoneit, B. R. T., Brassell, S. C., and Eglinton, G. (1978). Organic geochemical indicators of palaeoenvironmental conditions of sedimentation. Nature, 272, 216–22.Google Scholar
Dziak, R. P., Fox, C. G., Embley, R. W., Lupton, J. E., Johnson, G. C., Chadwick, W. W., and Koski, R. A. (1996). Detection of and response to a probable volcanogenic T-wave event swarm on the western Blanco Transform Fault Zone. Geophysical Research Letters, 23, 873–6.CrossRefGoogle Scholar
Earnshaw, J. P., Hogg, A. J. C., Oxtoby, N. H., and Cawley, S. J. (1993). Petrographic and fluid inclusion evidence for the timing of diagenesis and petroleum entrapment in the Papuan Basin. In Petroleum Exploration, Development and Production in Papua New Guinea: Proceedings of the Second PNG Petroleum Convention 31st May-2nd June, ed. G. J. Carman and Z. Carman. Port Moresby, Petroleum Exploration and Development, pp. 459–75.Google Scholar
Ebdon, C. C., Granger, P. J., Johnson, H. D., and Evans, A. M. (1995). Early Tertiary evolution and sequence stratigraphy of the Faeroe-Shetland Basin; implications for hydrocarbon prospectivity. In The Tectonics, Sedimentation, and Palaeoceanography of the North Atlantic Region, ed. Scrutton, R. A., Stoker, M. S., Shimmield, G. B., and Tudhope, A. W.. Geological Society Special Publications, pp. 5169.Google Scholar
Eberli, G. P. (1988). The evolution of the southern continental margin of the Jurassic Tethys ocean as recorded in the Allgaeu Formation of the Austroalpine Nappes of Graubuenden (Switzerland). Eclogae Geologicae Helvetiae, 81, 175214.Google Scholar
Ebinger, C. J. and Casey, M. (2001). Continental breakup in magmatic provinces: an Ethiopian example. Geology, 29, 527–30.2.0.CO;2>CrossRefGoogle Scholar
Ebinger, C. J., Jackson, J. A., Foster, A. N., and Hayward, N. J. (1999). Extensional basin geometry and the elastic lithosphere. Philosophical Transactions of the Royal Society, 357, 741–65.CrossRefGoogle Scholar
Ebner, F. and Sachsenhofer, R. F. (1995). Palaeogeography, subsidence and thermal history of the Neogene Styrian Basin (Pannonian Basin system, Austria): tectonics of the Alpine-Carpathian-Pannonian region. ALCAPA Meeting on Geological Evolution of the Internal Zones in the Alps, the Carpathians and of the Pannonian Basin, Graz, Austria, July 1–3, 242, 133–50.Google Scholar
Echarfaoui, H., Hafid, M., Ait Salem, A. A., and Ait Fora, A. (2002). Seismo-stratigraphic analysis of the Abda Basin, western Morocco: a case of inverse structures during Atlantic rifting. Comptes Rendus Geoscience (Academie des Sciences), 334, 371–7.Google Scholar
Eckstein, Y. and Simmons, G. (1978). Measurements and interpretation of terrestrial heat flow in Israel. Geothermics, 6, 117–42.Google Scholar
Edwards, M. H., Kurras, G. J., Tolstoy, M., Bohnenstiehl, M. Coakley, B. J. and Cochran, J. R. (2001). Evidence of recent volcanic activity on the ultraslow-spreading Gakkel ridge. Nature, 409.6822, 808812.CrossRefGoogle ScholarPubMed
Egan, S. S. and Meredith, D. J. (2005). The structural and geodynamic evolution of the Black Sea Basin. InterMARGINS Workshop: Modeling the Extensional Deformation of the Lithosphere. (IMEDL 2004), Abstract, Pontresina, Switzerland.Google Scholar
Egloff, F., Rihm, R., Makris, J., Izzeldin, Y. A., Bobsien, M., Meier, K., Junge, I., Noman, T., and Warsi, W. (1991). Contrasting structural styles of the eastern and western margins of the southern Red Sea: the 1988 SONNE experiment. Tectonophysics, 198, 329–53.CrossRefGoogle Scholar
Ehlers, T. A., Armstrong, P. A., and Chapman, D. S. (2001). Normal fault thermal regimes and interpretation: low-temperature thermochronometer data. Physics of the Earth and Planetary Interiors, 126, 179–94.CrossRefGoogle Scholar
Ehlers, T. A. and Chapman, D. S. (1999). Normal fault thermal regimes: conductive and hydrothermal heat transfer surrounding the Wasatch fault, Utah. Tectonophysics, 312, 217–34.CrossRefGoogle Scholar
Ehlers, T. A. and Farley, K. A. (2003). Apatite (U-Th)/He’thermochronometry, methods, and applications to problems in tectonic and surface processes. Earth and Planetary Science Letters, 206, 114.CrossRefGoogle Scholar
Ehlers, T. A., Willett, S. D., Armstrong, P. A., and Chapman, D. S. (2003). Exhumation of the central Wasatch Mountains, Utah; 2, Thermokinematic model of exhumation, erosion, and thermochronometer interpretation. Journal of Geophysical Research, 108, 18.CrossRefGoogle Scholar
Eichhubl, P., Greene, H. G., Naehr, T., and Maher, N. (2000). Structural control of fluid flow: offshore seepage in the Santa Barbara Basin, California. Journal of Geochemical Exploration, 6970, 545–9.Google Scholar
Elders, W. A., Rex, R. W., Meidav, T., Robinson, P. T., and Biehler, S. (1972). Crustal spreading in southern California. Science, 178, 1524.CrossRefGoogle ScholarPubMed
Eldholm, O. and Grue, K. (1994). North Atlantic volcanic margins: dimensions and production rates. Journal of Geophysical Research, 99, 2955–68.CrossRefGoogle Scholar
Eldholm, O., Skogseid, J., Planke, S., and Gladczenko, P. (1995). Volcanic margin concept. In Rifted Ocean– Continent Boundaries, ed. Banda, E.. Kluwer Academic Publishers, p. 16.Google Scholar
Eldholm, O., Thiede, J., and Taylor, E. (1989). Evolution of the Vøring volcanic margin. In Proceedings of Ocean Drilling Program (ODP), ed. Eldholm, O., Thiede, J., and Taylor, E.. Scientific Results, 104, 1033–65.Google Scholar
Eliet, P. P. and Gawthorpe, R. L. (1995). Drainage development and sediment supply within rifts, examples from the Sperchios Basin, central Greece. Journal of the Geological Society of London, 152, 883–93.CrossRefGoogle Scholar
Eliuk, L. and Wach, G. D. (2010). Large scale mixed carbonate-siliciclastic clinoform systems: three types from the Mesozoic North American Atlantic offshore. AAPG Annual Conference, New Orleans, April 11–14.Google Scholar
Ellevset, S., Ottesen, D., Knipe, R. J., Olsen, T. S., Fisher, Q. J., and Jones, G. (1998). Fault controlled communication in the Sleipner Vest Field, Norwegian continental shelf: detailed, quantitative input for reservoir simulation and well planning. In Faulting, Fault Sealing and Fluid Flow in Hydrocarbon Reservoirs, ed. Jones, G., Fisher, Q. J., and Knipe, R. J.. Geological Society of London Special Publications, 147, 283–97.Google Scholar
Ellis, S. and Beaumont, C. (1999). Models of convergent boundary tectonics: implications for the interpretation of Lithoprobe data. Canadian Journal of Earth Sciences, 36, 1711–41.CrossRefGoogle Scholar
Embry, A. F. (1990). A tectonic origin for third-order depositional sequences in extensional basins; implications for basin modeling. In Quantitative dynamic stratigraphy, ed. Cross, T. A.. Prentice Hall, Englewood Cliffs, New Jersey.Google Scholar
Enescu, D., Danchiv, D., and Bálá, A. (1992). Lithosphere structure in Romania II: thickness of the Earth’s crust, depth dependent propagation velocity curves for P and S waves. Studii Cercetri Geofizica, 30, 319.Google Scholar
England, P. C. (1983). Constraints on extension of continental lithosphere. Journal of Geophysical Research, 88, 1145–52.CrossRefGoogle Scholar
England, P. C. and Molnar, P. (1990). Surface uplift, uplift of rocks, and exhumation of rocks. Geology, 18, 1173–7.2.3.CO;2>CrossRefGoogle Scholar
England, P. C. and Richardson, S. W. (1977). The influence of erosion upon mineral facies of rocks from different metamorphic environments. Journal of the Geological Society of London, 134, 201–13.CrossRefGoogle Scholar
England, W. A. (1994). Secondary migration and accumulation of hydrocarbons. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 211–17.Google Scholar
England, W. A., Mackenzie, A. S., Mann, D. M., and Quigley, T. M. (1987). The movement and entrapment of petroleum fluids in the subsurface. Journal of the Geological Society of London, 144, 327–47.CrossRefGoogle Scholar
Erratt, D. (1993). Relationships between basement faulting, salt withdrawal and Late Jurassic rifting, UK Central North Sea. In Petroleum Geology of Northwest Europe, Proceedings of the 4th Conference, ed. Parker, J. R.. The Geological Society of London, 1211–19.Google Scholar
Escalona, A. (2003). Regional tectonics, sequence stratigraphy and reservoir properties of Eocene clastic sedimentation, Maracaibo Bain, Venezuela. PhD thesis, University of Texas, Austin, p. 222.Google Scholar
Escartín, J., Mevel, C., MacLeod, C. J., and McCaig, A. M. (2003). Constraints on deformation conditions and the origin of oceanic detachments: the Mid-Atlantic Ridge core complex at 15 degrees 45'N. Geochemistry, Geophysics, Geosystems, 4.CrossRefGoogle Scholar
Espitalié, J., Madec, M., Tissot, B., Menning, J. J., and Leplat, P. (1977a). Source rock characterization method for petroleum exploration. Ninth Annual Offshore Technology Conference, 439–48.Google Scholar
Espitalié, J., Laporte, J. L., Madec, M., Marquis, F., Leplat, P., Paulet, J. and Boutefeu, A. (1977b). Rapid method for source rock characterization, and for determination of their petroleum potential and degree of evolution. Revue de l’Institut Francais du Petrole et Annales des Combustibles Liquides, 32, 2342.Google Scholar
Esteoule-Choux, J. (1983). Kaolinitic weathering profiles in Britany: genesis and economic importance. Contributions to Mineralogy and Petrology, 75, 129–52.Google Scholar
Eugster, H. P. and Hardie, L. A. (1975). Sedimentation in an ancient playa-lake complex: the Wilkins Peak Member of the Green River Formation of Wyoming. Geological Society of America Bulletin, 86, 319–34.2.0.CO;2>CrossRefGoogle Scholar
Evans, A. C. and Parkinson, D. N. (1983). A half-graben and tilted fault block structure in the northern North Sea. In Seismic Expression of Structural Styles: A Picture and Work Atlas; Volume 2, ed. A. W. Bally. American Association of Petroleum Geologists Studies in Geology, 15, 2.2.2-7 – 2.2.2-11.Google Scholar
Evans, D. G. and Nunn, J. A. (1989). Free thermohaline convection in sediments surrounding a salt column. Journal of Geophysical Research, 94, 12413–22.CrossRefGoogle Scholar
Evans, D. G., Nunn, J. A., and Hanor, J. S. (1991). Mechanisms driving groundwater flow near salt domes. In Crustal-Scale Fluid Transport: Magnitude and Mechanisms, ed. Torgensen, T.. Geophysical Research Letters, 18, 927–30.Google Scholar
Evans, K. F., Burford, R. O., and King, G. C. P. (1981). Propagating episodic creep and the aseismic slip behavior of the Calaveras Fault north of Hollister, California. Journal of Geophysical Research, 86, 3721–35.CrossRefGoogle Scholar
Ewart, A., Baxter, K., and Ross, J. A. (1980). The petrology and petrogenesis of the Tertiary anorogenic mafic lavas of southern and central Queensland, Australia – possible implications for crustal thickening. Contributions to Mineralogy and Petrology, 75, 129–52.CrossRefGoogle Scholar
Exon, N. and Colwell, J. B. (1994). Geological history of the outer North West Shelf of Australia: a synthesis. In Geology of the Outer North West Shelf, Australia: Australian Geological Survey Organization, ed. N. Exon. AGOSO Journal of Australian Geology and Geophysics, 15(1), 177–90.Google Scholar
Faccenna, C., Davy, P., Brun, J. P., Funiciello, R., Giardini, D., Mattei, M., and Nalpas, T. (1996). The dynamics of back-arc extension: an experimental approach to the opening of the Tyrrhenian Sea. Geophysical Journal International, 126, 781–95.CrossRefGoogle Scholar
Fachmann, S. (2001). Geologische Entwicklung im Umfeld des Mahanadi-Riftes (Indien). Universitat Freiberg.Google Scholar
Faerseth, R. B., Knudsen, B. E., Liljedahl, T., Midboe, P. S., and Soderstrom, B. (1997). Oblique rifting and sequential faulting in the Jurassic development of the northern North Sea. Journal of Structural Geology, 19, 1285–302.Google Scholar
Fainstein, R., Milliman, J. D., and Jost, H. (1975). Magnetic character of the Brazilian continental shelf and upper slope. Revista Brasileira de Geociencias, 5, 198211.Google Scholar
Fairhead, J. D. (1988a). Mesozoic plate tectonic reconstructions of the central South Atlantic Ocean: the role of the West and Central African rift system. Tectonophysics, 155, 181–91.CrossRefGoogle Scholar
Fairhead, J. D. (1988b). Late Mesozoic rifting in Africa. In Triassic-Jurassic Rifting: Continental Breakup and the Origin of the Atlantic Ocean and Passive Margins, ed. Manspeizer, W.. Developments in Geotectonics, 22, 821–31.Google Scholar
Fairhead, J. D. and Binks, R. M. (1991). Differential opening of the Central and South Atlantic Oceans and the opening of the West African rift system. Tectonophysics, 187, 191203.CrossRefGoogle Scholar
Fairhead, J. D. and Okereke, C. S. (1987). A regional gravity study of the West African rift system in Nigeria and Cameroon and its tectonic interpretation. Tectonophysics, 143, 141–59.CrossRefGoogle Scholar
Faleide, J. I., Kyrkjebo, R., Kjennerud, T., Gabrielsen, R., Jordt, J., Fanavoll, S., and Bjerke, M. D. (2002). Tectonic impact on sedimentary processes during Cenozoic evolution of the northern North Sea and surrounding areas. In Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration, ed. Dore, A. G., Cartwright, J. A., Stoker, M. S., Turner, J. P., and White, N. J.. Geological Society Special Publications, 196, 235–69.Google Scholar
Falt, U., Guerin, G., Retail, P., and Evans, M. (1992). Clair discovery: evaluation of natural fracturation in a horizontal well drilled in the basement and producing from overlying sediments. Proceedings: European Petroleum Conference: Moving the Frontiers, Sharing Solutions, Society of Petroleum Engineers, 2, 1121.Google Scholar
Farley, K. A. (2000). Helium diffusion from apatite: general behavior as illustrated by Durango fluorapa. Journal of Geophysical Research, 105, 2000.CrossRefGoogle Scholar
Farley, K. A. (2002). (U-Th)/He dating: techniques, calibrations, and applications. Reviews in Mineralogy and Geochemistry, 47, 819–43.CrossRefGoogle Scholar
Farley, K. A., Kohn, B. P., and Pillans, B. (2002). The effects of secular disequilibrium on (U-Th)/He systematics and dating of Quaternary volcanic zircon and apatite. Earth and Planetary Science Letters, 201, 114–25.CrossRefGoogle Scholar
Farley, K. A., Wolf, R. A., and Silver, L. T. (1996). The effects of long alpha-stopping distances on (U-Th)/He ages. Geochimica et Cosmochimica Acta, 60, 4223–9.CrossRefGoogle Scholar
Fassoulas, C., Kilias, A., and Mountrakis, D. (1994). Postnappe stacking extension and exhumation of high-pressure/low-temperature rocks in the Island of Crete, Greece. Tectonics, 13, 125–38.CrossRefGoogle Scholar
Faugère, E. and Brun, J. P. (1984). Modelisation experimentale de la distention continentale. Experimental models of a stretched continental crust. Compte Rendu Academie des Sciences, 299, 365–70.Google Scholar
Faugère, E., Brun, J. P., and van den Driessche, J. (1986). Bassins asymetriques en extension pure et en decrochement: modeles experimentaux. Asymmetric extension and unhooked basins: experimental models. Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine, 10, 1321.Google Scholar
Faulds, J. E., Geissman, J. W., and Mawer, C. K. (1990). Structural development of a major extensional accommodation zone in the Basin and Range Province, northwestern Arizona and southern Nevada: implications for kinematic models of continental extension. Memoir – Geological Society of America, 176, 3776.CrossRefGoogle Scholar
Faure, G. and Mensing, T. M. (2005). Isotopes. Principles and Applications. 3rd edition, John Wiley and Sons, p. 897.Google Scholar
Fechtig, H. and Kalbitzer, S. (1966). The Diffusion of Argon in Potassium-Bearing Solids. Springer, pp. 68107.CrossRefGoogle Scholar
Feighner, M. A., Kellogg, L. H., and Travis, B. J. (1995). Numerical modeling of chemically buoyant mantle plumes at spreading ridges. Geophysical Research Letters, 22, 715–18.CrossRefGoogle Scholar
Feighner, M. A. and Richards, M. A. (1995). The fluid dynamics of plume-ridge and plume-plate interactions: an experimental investigation. Earth and Planetary Science Letters, 129, 171–82.CrossRefGoogle Scholar
Feraud, G., Hofmann, C., and Alric, V. I. (1996). Successful (super 40) Ar/ (super 39) Ar laser probe dating of whole rock basalts affected by 39Ar recoil. Eos Transactions American Geophysical Union, 77, 93–4.Google Scholar
Ferentinos, G., Papatheodorou, G., and Collins, M. B. (1988). Sediment transport processes on an active submarine fault escarpment: Gulf of Corinth, Greece. Marine Geology, 83, 4361.CrossRefGoogle Scholar
Ferguson, I. J., Westbrook, G. K., Langseth, M. G., and Thomas, G. P. (1993). Heat flow and thermal models of the Barbados Ridge accretionary complex. Journal of Geophysical Research, 98, 4121–42.CrossRefGoogle Scholar
Fernàndez, M., Ayala, C., Torne, M., Vergés, J., Gómez, M., and Karpuz, R. (2005). Lithospheric structure of the Mid-Norwegian Margin: comparison between the Møre and Vøring margins. Journal of Geological Society of London, 162, 1005–12.CrossRefGoogle Scholar
Ferrara, G. C., Palmerini, G. C., and Scappini, U. (1985). Update report on geothermal development in Italy. International Symposium on Geothermal Energy, International Volume(GRCT), 95105.Google Scholar
Ferre, E. C., Bordarier, C., and Marsh, J. S. (2002). Magma flow inferred from AMS fabrics in a layered mafic sill, Insizwa, South Africa. Tectonophysics, 354, 123.CrossRefGoogle Scholar
Ferriday, I. and Hall, K. (1993). Sourcing of oils on the eastern flank of the Northern Viking Graben. Proceedings of the 16th International Meeting on Organic Geochemistry, Stavanger, Norway, September 20–4, pp. 43–5.Google Scholar
Feyzullayev, A. A., Guliyev, I. S., and Tagiyev, M. F. (2001). Source potential of the Mesozoic-Cenozoic rocks in the South Caspian basin and their role in forming the oil accumulations in the lower Pliocene reservoirs. Petroleum Geoscience, 7(4), 409–17.CrossRefGoogle Scholar
Field, J. D. (1985). Organic geochemistry in exploration of the northern North Sea. In Petroleum Geochemistry in Exploration of the Norwegian Shelf, ed. Thomas, B. M., Dore, A. G., Eggen, S. S., Home, P. C., and Mange, L. R.. Graham and Trotman, London, pp. 3957.CrossRefGoogle Scholar
Figueiredo, A. M. F. 1985. In De Azevedo, R. P. (1991). Tectonic evolution of Brazilian equatorial continental margin basins. PhD thesis, Imperial College, London, pp. 1–403.Google Scholar
Filho, J. D. S., Correa, A. C. F., Neto, E. V. S., and Trinidade, L. A. F. (2000). Alagamar-Acu petroleum system, onshore Potiguar Basin, Brazil: a numerical approach for secondary migration. In Petroleum Systems of South Atlantic Margins, ed. Mello, M. R. and Katz, B. J.. American Association of Petroleum Geologists Memoir, 73, 151–8.Google Scholar
Filjak, R., Pletikapic, Z., Nikolic, D., and Aksin, V. (1969). Geology of petroleum and natural gas from the Neogene complex and its basement in the southern part of the Pannonian basin, Yugoslavia. Conference of Institute of Petroleum and AAPG, Brighton, Proceedings, 113–30.Google Scholar
Filmer, P. E., McNutt, M. K., Webb, H. F., and Dixon, D. J. (1994). Volcanism and archipelagic aprons in the Marquesas and Hawaiian Islands. Marine Geophysical Research, 16, 385406.CrossRefGoogle Scholar
Finlayson, D. M., Collins, C. D. N., Lukaszyk, I., and Chudyk, E. C. (1998). A transect across Australia’s southern margin in the Otway Basin region: crustal architecture and the nature of rifting from wide-angle seismic profiling. Tectonophysics, 288, 177–89.CrossRefGoogle Scholar
Fisher, A. T. and Becker, K. (1995). Correlation between seafloor heat flow and basement relief: observational and numerical examples and implications for upper crustal permeability. Journal of Geophysical Research, 100, 12641–57.CrossRefGoogle Scholar
Fisher, A. T., Zwart, G., Shipley, T., Ogawa, Y., Ashi, J., Blum, P., Brueckmann, W., Filice, F., Goldberg, D., Henry, P., Housen, B., Jurado, M. J., Kastner, M., Labaume, P., Laier, T., Leitch, E., Maltman, A., Meyer, A., Moore, J. C., Moore, G., Peacock, S., Rabaute, A., Steiger, T., and To, H. (1996). Relation between permeability and effective stress along a plate-boundary fault, Barbados accretionary complex. Geology, 24, 307–10.2.3.CO;2>CrossRefGoogle Scholar
Fisher, J. B. and Boles, J. R. (1990). Water-rock interaction in Tertiary sandstones, San Joaquin Basin, California, U.S.A.: Diagenetic controls on water composition. Chemical Geology, 82, 83101.CrossRefGoogle Scholar
Fisher, M. J. and Miles, J. A. (1983). Kerogen type, organic maturation and hydrocarbon occurrences in the Moray Firth and South Viking Graben, North Sea Basin. In Petroleum Geochemistry and Exploration of Europe (second edition), ed. Brooks, J.. Geological Society of London Special Publications, 12, 195201.Google Scholar
Fisher, Q. J. and Knipe, R. J. (1998). Fault sealing processes in siliciclastic rocks. In Faulting, Fault Sealing and Fluid Flow in Hydrocarbon Reservoirs, ed. Jones, G., Fisher, Q. J., and Knipe, R. J.. Geological Society of London Special Publications, 147, 117–34.Google Scholar
Fishwick, S., Heintz, M., Kennett, B. L. N., Reading, A. M., and Yoshizawa, K. (2008). Steps in lithospheric thickness within eastern Australia: evidence from surface wave tomography. Tectonics, 27.CrossRefGoogle Scholar
Fitzgerald, P. G. (1992). The Transantarctic Mountains of southern Victoria Land: the application of apatite fission track analysis to a rift shoulder uplift. Tectonics, 11, 634–62.CrossRefGoogle Scholar
Fjellanger, E., Olsen, T. R., and Rubino, R. L. (1996). Sequence stratigraphy and paleogeography of the Middle Jurassic Brent and Vestland deltaic systems, Northern North Sea. Norsk Geologist Tiddskrift, 76, 75106.Google Scholar
Fleitout, L., Froidevaux, C., and Yuen, D. (1986). Active lithosphere thinning. Tectonophysics, 132, 271–8.CrossRefGoogle Scholar
Fleitout, L. and Yuen, D. A. (1984a). Secondary convection and the growth of the oceanic lithosphere. Physics of the Earth and Planetary Interiors, 36, 181212.CrossRefGoogle Scholar
Fleitout, L. and Yuen, D. A. (1984b). Steady state, secondary convection beneath lithospheric plates with temperature- and pressure-dependent viscosity. Journal of Geophysical Research, 89, 9227–44.CrossRefGoogle Scholar
Florensov, N. A. (1966). The Baikal Rift Zone. Nauka, Moscow.CrossRefGoogle Scholar
Flournoy, L. A. and Ferrell, R. E. (1980). Geopressure and diagenetic modifications of porosity in the Lirette field area, Terrebonne Parish, Louisiana. Gulf Coast Geological Association Transactions, 30, 341–5.Google Scholar
Flovenz, O. G. and Saemundsson, K. (1993). Heat flow and geothermal processes in Iceland. Tectonophysics, 225, 123–38.CrossRefGoogle Scholar
Forbes, P. L., Ungerer, P. M., Kuhfuss, A. B., Riis, F., and Eggen, S. (1991). Compositional modeling of petroleum generation and expulsion: trial application to a local mass balance in Smorbuk Sor field, Haltenbanken area, Norway. American Association of Petroleum Geologists Bulletin, 75, 873–93.Google Scholar
Fournier, R. O. (1991). The transition from hydrostatic to greater than hydrostatic fluid pressure in presently active continental hydrothermal systems in crystalline rock. Geophysical Research Letters, 18, 955–8.Google Scholar
Fournier, R. O. (1999). Hydrothermal processes related to movement of fluid from plastic into brittle rock in the magmatic – epithermal environment. Economic Geology, 94, 1193–212.CrossRefGoogle Scholar
Fournier, R. O., White, D. E., and Truesdell, A. H. (1974). Geochemical indicators of subsurface temperatures: Part 1. Basic assumptions. Journal of Geophysical Research, 101, 25499–509.Google Scholar
Foster, D. A. and Gleadow, A. J. W. (1994). Tertiary structural framework of basement flanking the Kenya Rift: constraints from an apatite fission-track thermal-stratigraphy. Abstracts – Geological Society of Australia, 37, 115.Google Scholar
Fowler, M. G. and McAlpine, K. D. (1994). The Egret Member: a prolific Kimmeridgian source rock from offshore eastern Canada. In Petroleum Source Rocks, ed. Katz, B. J.. Springer-Verlag, New York, pp. 111–30.Google Scholar
Francis, T. J. G., Davies, D., and Hill, M. N. (1966). Crustal structure between Kenya and the Seychelles. Royal Society of London Philosophical Transactions, 259, 240–61.Google Scholar
Francis, T. J. G. and Shor, G. G. (1966). Seismic refraction measurement in the northwest Indian Ocean. Journal of Geophysical Research, 71, 427–49.CrossRefGoogle Scholar
Frape, S. K. and Fritz, P. (1987). Geochemical trends for groundwaters from the Canadian Shield. In Saline Water and Gases in Crystalline Rocks, ed. Fritz, P. and Frape, S. K.. Geological Association of Canada Special Papers, 33, 1938.Google Scholar
Fraser, A. R. and Tonkin, P. C. (1991). The Glamis Field, Block 16/21a, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 317–22.Google Scholar
Fraser, S., Robinson, A. Q., Johnson, H., Underhill, J., Kadolsky, D., Connell, R., Johannessen, P., and Ravnas, R. (2002). Chapter 11. Upper Jurassic. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. Geological Society of London, pp. 157–89.Google Scholar
Frederiksen, S., Nielsen, S. B., and Balling, N. (2001). Post-Permian evolution of the central North Sea: a numerical model. Tectonophysics, 343, 185203.CrossRefGoogle Scholar
Freeze, R. A. and Cherry, J. A. (1979). Groundwater. Prentice Hall, Englewood Cliffs, New Jersey, p. 604.Google Scholar
Frey, F. A., McNaughton, N. J., Nelson, D. R., de Laeter, J. R., and Duncan, R. T. A. (1996). Petrogenesis of the Bunbury Basalt, Western Australia: between the Kerguelen plume and Gondwana lithosphere? Earth and Planetary Science Letters, 144(1–2), 163–83.CrossRefGoogle Scholar
Friedman, G. M. (1999). Thermal anomalies associated with forced and free ground-water convection in the Dead Sea Rift Valley: Discussion. GSA Bulletin, 111, 1098–101.2.3.CO;2>CrossRefGoogle Scholar
Fristad, T., Gorth, A., Yielding, G., and Freeman, B. (1997). Quantitative fault seal prediction – a case study from Oseberg Syd. In Hydrocarbon Seals – Importance for Exploration and Production, ed. Moller-Pederson, P. and Koestler, A. G.. Special Publication of the Norwegian Petroleum Society, 7, 107–24.Google Scholar
Frost, B. R. and Bucher, K. (1994). Is water responsible for geophysical anomalies in the deep continental crust? A petrological perspective. Tectonophysics, 231, 293309.CrossRefGoogle Scholar
Frostick, L. and Reid, I. (1987). Is structure the main control of river drainage and sedimentation in rifts? Journal of African Earth Sciences, 8, 165–82.Google Scholar
Fu, J., Shen, G., Peng, P., Brassell, S. C., Elington, G., and Liang, J. (1986). Peculiarities of salt lake sediments as potential source rocks in China. Organic Geochemistry, 10, 119–26.Google Scholar
Fuis, G. S. and Kohler, W. M. (1984). Crustal structure and tectonics of the Imperial Valley region, California. Field Trip Guidebook – Pacific Section: Society of Economic Paleontologists and Mineralogists, 40, 113.Google Scholar
Funck, T., Hopper, J. R., Larsen, H. C., Louden, K. E., Tucholke, B. E., and Holbrook, W. S. (2003). Crustal structure of the ocean–continent transition at Flemish Cap: seismic refraction results. Journal of Geophysical Research, 108, 20.CrossRefGoogle Scholar
Funck, T. and Schmincke, H. U. (1998). Growth and destruction of Gran Canaria deduced from seismic reflection and bathymetric data. Journal of Geophysical Research, 103, 15393–407.CrossRefGoogle Scholar
Fusion Oil and Gas (2002). Deepwater Northwest Africa: A Geological Tour. Presentation in Cape Town.Google Scholar
Fyfe, J. A., Gregersen, U., Jordt, H., Rundberg, Y., Eidvin, T., Evans, D., Stewart, D., Hovland, M., and Andresen, P. (2003). Oligocene to Holocene. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. The Geological Society of London, pp. 279–87.Google Scholar
Gaarenstrom, L., Tromp, R. A. J., Jong, M. C., and Brandenberg, A. M. (1993). Overpressures in the Central North Sea: implications for trap integrity and drilling safety. In Petroleum Geology of Northwest Europe, ed. Parker, J. R.. Proceedings of the 4th Conference, The Geological Society of London, 1305–13.Google Scholar
Gaina, C., Muller, R. D., Brown, B., and Ishihara, T. (2003). Micro-continent formation around Australia. In The Evolution and Dynamics of the Australian Plate, ed. Hillis, R. and Muller, R. D.. Joint Geological Association of Australia and Geological Society of America Special Paper, 22, 399410.Google Scholar
Galanis, Jr. S. P., Sass, J. H., Munroe, R. J., and Abu-Ajamieh, M. (1986). Heat flow at Zerka-Ma’in and Zara and a geothermal reconnaissance of Jordan. US Geological Survey Open-File Report, 86, 110.Google Scholar
Galbraith, R. F. (1988). Graphical display of estimates having differing standard errors. Technometrics, 30, 271–81.CrossRefGoogle Scholar
Galbraith, R. F. (1992). Statistical models for mixed ages. International Workshop on Fission Track Thermochronology, Philadelphia.Google Scholar
Gallagher, K. and Brown, R. (1999). The Mesozoic denudation history of the Atlantic margins of southern Africa and southeast Brazil and the relationship to offshore sedimentation. In The Oil and Gas Habitats of the South Atlantic, ed. Cameron, N. R., Bate, R. H., and Clure, V. S.. Geologic Society of London Special Publications, 153, 4153.Google Scholar
Gallagher, K., Hawkesworth, C. J., and Mantovani, M. S. M. (1994). The denudation history of the onshore continental margin of SE Brazil inferred from apatite fission track data. Journal of Geophysical Research, 99, 18117–45.CrossRefGoogle Scholar
Galloway, W. E. (1989a). Depositional framework and hydrocarbon resources of the early Miocene (Fleming) episode, Northwest Gulf Coast Basin. Marine Geology, 90, 1929.CrossRefGoogle Scholar
Galloway, W. E. (1989b). Genetic stratigraphic sequences in basin analysis; I, Architecture and genesis of flooding-surface bounded depositional units. American Association of Petroleum Geologists Bulletin, 73, 125–42.Google Scholar
Gamond, J. F. (1987). Bridge structures as sense of displacement criteria in brittle fault zones. Journal of Structural Geology, 9, 609–20.CrossRefGoogle Scholar
Gans, P. B. (1987). An open-system, two-layer crustal stretching model for the eastern Great Basin. Tectonics, 6, 112.CrossRefGoogle Scholar
Garcia, A. J. V., Morad, S., De Ros, L. F., and Al-Aasm, I. S. (1998). Palaeogeographical, palaeoclimatic and burial history controls on the diagenetic evolution of reservoir sandstones: evidence from the Lower Cretaceous Serraria sandstones in the Sergipe-Alagoas Basin, NE Brazil. Special Publication of the International Association of Sedimentologists, 26, 107–40.Google Scholar
Garden, I. R., Guscott, S. C., Burley, S. D., Foxford, K. A., Walsh, J. J., and Marshall, J. (2001). An exhumed palaeo-hydrocarbon migration fairway in a faulted carrier system, Entrada Sandstone of SE Utah, U.S.A. Geofluids, 1, 195213.CrossRefGoogle Scholar
Gardner, G. H. F., Gardner, L. W., and Gregory, A. R. (1974). Formation velocity and density – the diagnostic basics for stratigraphic traps. Geophysics, 39, 770–80.CrossRefGoogle Scholar
Garfunkel, Z. B. Y (1977). The tectonics of the Suez Rift. Bulletin – Geological Survey of Israel, 71, 44.Google Scholar
Garland, C. R. (1993). Miller Field: reservoir stratigraphy and its impact on development. In Petroleum Geology of Northwest Europe. Proceedings of the 4th Conference, ed. Parker, J. R.. Geological Society of London, 401–14.Google Scholar
Garven, G. (1989). A hydrogeologic model for the formation of the giant oil sand deposits of the western Canada sedimentary basin. American Journal of Science, 289, 105–66.CrossRefGoogle Scholar
Garven, G. and Freeze, R. A. (1984a). Theoretical analysis of the role of groundwater flow in the genesis of stratabound ore deposits: 1. Mathematical and numerical model. American Journal of Science, 284, 1085–124.Google Scholar
Garven, G. and Freeze, R. A. (1984b). Theoretical analysis of the role of groundwater flow in the genesis of stratabound ore deposits: 2. Quantitative results. American Journal of Science, 284, 1124–56.Google Scholar
Gaulier, J. M., Le Pichon, X., Lyberis, N., Avedik, F., Gely, I., Moretti, I., Deschamps, A., and Hafez, S. (1988). Seismic study of the crustal thickness, northern Red Sea and Gulf of Suez. Tectonophysics, 153, 5588.CrossRefGoogle Scholar
Gawthorpe, R. L., Jackson, C. A. L., Young, M. J., Sharp, I. R., Moustafa, A. R., and Leppard, C. W. (2003). Normal fault growth, displacement localisation and the evolution of normal fault populations: the Hammam Faraun fault block, Suez Rift, Egypt. Journal of Structural Geology, 25, 883–95.Google Scholar
Gawthorpe, R. L., Fraser, A. J., and Collier, R. E. L. (1994). Sequence stratigraphy in active extensional basins: implications for the interpretation of ancient basin-fills. Marine and Petroleum Geology, 11, 642–58.CrossRefGoogle Scholar
Gawthorpe, R. L. and Hurst, J. M. (1993). Transfer zones in extensional basins: their structural style and influence on drainage development and stratigraphy. Journal of the Geological Society of London, 150, 1137–52.CrossRefGoogle Scholar
Gawthorpe, R. L., Hurst, J. M., and Sladen, C. P. (1990). Evolution of Miocene footwall-derived coarse-grained deltas, Gulf of Suez, Egypt: implications for exploration. American Association of Petroleum Geologists Bulletin, 74, 1077–86.Google Scholar
Gawthorpe, R. L. and Leeder, M. R. (2000). Tectono- sedimentary evolution of active extensional basins. Basin Research, 12, 195218.CrossRefGoogle Scholar
Gawthorpe, R. L., Sharp, I. R., Underhill, J. R., and Gupta, S. (1997). Linked sequence stratigraphic and structural evolution of propagating normal faults. Geology, 25, 795–8.2.3.CO;2>CrossRefGoogle Scholar
Ge, S. and Garven, G. (1989). Tectonically induced transient groundwater flow in foreland basins. In Origin and Evolution of Sedimentary Basins and Their Energy and Mineral Resources, ed. Price, R. A.. American Geophysical Union, Geophysical Monograph, 48, 145–57.Google Scholar
Ge, S. and Garven, G. (1992). Hydromechanical modeling of tectonically driven groundwater flow with application to the Arkoma foreland basin. Journal of Geophysical Research, 97, 9119–44.CrossRefGoogle Scholar
Ge, T. and Chen, Y. (1993). Liaohe Oil Field: Petroleum Geology of China III (in Chinese). Petroleum Industry Press, Beijing, p. 39.Google Scholar
Gebauer, D. (1993). The pre-Alpine evolution of the continental crust of the Central Alps – an overview. In Pre Mesozoic Geology in the Alps, ed. Von Raumer, J. F. and Neubauer, F.. Springer, Berlin, pp. 93117.CrossRefGoogle Scholar
Gee, M., Watts, A. B., Masson, D. G., and Mitchell, N. C. (2001). Landslides and the evolution of the El Hierro in the Canary Island. Marine Geology, 177, 271–93.CrossRefGoogle Scholar
Geiser, P. (1974). Cleavage in some sedimentary rocks of the central Valley and Ridge province, Maryland. Geological Society of America Bulletin, 85, 1399–412.2.0.CO;2>CrossRefGoogle Scholar
Genik, G. J. (1993). Petroleum geology of the Cretaceous- Tertiary rift basins in Niger, Chad and Central African Republic. American Association of Petroleum Geologists Bulletin, 77, 1405–34.Google Scholar
Geoffroy, L. (1994). Contraintes et volcanisme: Le domaine Nord-Est Atlantique au Tertiaire inferieur. PhD thesis, University of Paris, p. 426.Google Scholar
Geoffroy, L. (2001). The structure of volcanic margins: some problematics from the North Atlantic/Baffin Bay system. Marine and Petroleum Geology, 18, 463–9.CrossRefGoogle Scholar
Geoffroy, L. (2005). Volcanic passive margins. Comptes Rendus Geosciences, 337, 1395–408.CrossRefGoogle Scholar
Geoffroy, L., Callot, J. P., Caillet, S., Skuce, A. S., Gelard, J. P., Ravilly, M., Angelier, J., Bonin, B., Cayet, C., Perrot-Galmiche, K., and Lepvrier, C. (2001). Southeast Baffin volcanic margin and the North American–Greenland plate separation. Tectonics, 20, 566–84.CrossRefGoogle Scholar
George, R., Rogers, N., and Kelley, S. (1998). Earliest magmatism in Ethiopia: evidence for two mantle plumes in one flood basalt providence: Geology, 26, 923–6.2.3.CO;2>CrossRefGoogle Scholar
Geotrack-International (1999a). (U-Th)/He dating of apatite. Improved resolution at low temperatures. Geotrack Information Sheet 99/8. Geotrack International Pty Ltd, Brunswick West.Google Scholar
Geotrack-International (1999b). Thermal history interpretation of AFTA data. Geotrack Information Sheet 99/2. Geotrack International Pty Ltd, Brunswick West.Google Scholar
Gernigon, L. (2002). Extension et magmatisme en context de marge passive volcanique: deformation et satructure crustale de la marge norvegienne externe. PhD thesis. University of Eastern Bretagne, p. 301.Google Scholar
Geyh, M. A. and Schleicher, H. (1990). Absolute Age Determination: Physical and Chemical Dating Methods and Their Application. Springer, Berlin.CrossRefGoogle Scholar
Ghebreab, W. and Talbot, C. J. (2000). Red Sea extension influenced by pan-African tectonic grain in eastern Eritrea. Journal of Structural Geology, 22, 931–46.CrossRefGoogle Scholar
Ghignone, J. I. (1979). Geologia dos sedimentos fanerozoicos do estado da Bahia. Geology of Phanerozoic sediments in Bahia. In Quantificacao dos recursos hidricos subterraneos do aquifero reconcavo na bacia do rio Capivara, ed. Cavalcanti, S. S.. Tese de Doutorado em Geofísica. Instituto de Geociencias, Universidade Federal da Bahia. Salvador – Bahia, p. 121.Google Scholar
Ghignone, J. I. (1986). Estratigrafia, estrutura e possibilidades de petróleo das bacias do Araripe, Iguatu e Rio do Peixe. Anais do XXXIV Congresso Brasileiro de Geologia, Goiânia, 1, 271–85.Google Scholar
Giambalvo, E. R., Steefel, C. I., Fisher, A. T., Rosenberg, N. D., and Wheat, C. G. (2002). Effects of fluid-sediment reaction on hydrothermal fluxes of major elements, eastern flank of the Juan de Fuca Ridge. Geochimica et Cosmochimica Acta, 66, 1739–57.CrossRefGoogle Scholar
Gibbs, A. D. (1984). Structural evolution of extensional basin margins. Journal of the Geological Society of London, 141, 609–20.CrossRefGoogle Scholar
Gibbson, I. L. and Love, D. (1989). A listric fault model for the formation of the dipping reflectors penetrated during the drilling of hole 642E, ODP Leg 104. In ODP Proceedings, ed. O. Eldholm, J. Thiede, E. Taylor and colleagues. Scientific Results, 104, 979–83.Google Scholar
Gilchrist, A. R. and Summerfield, M. A. (1990). Differential denudation and flexural isostasy in formation of rifted-margin upwarps. Nature, 346, 739–42.CrossRefGoogle Scholar
Gillot, P.-Y. and Cornette, Y. (1986). The Cassignol technique for potassium-argon dating, precision and accuracy: examples from the late Pleistocene to Recent volcanics from southern Italy. Chemical Geology: Isotope Geoscience Section, 59, 205–22.Google Scholar
Gilpin, B. and Lee, T. C. (1978). A microearthquake study in the Salton Sea geothermal area, California. Bulletin of the Seismological Society of America, 68, 441–50.Google Scholar
Girdler, R. W. (1970). A review of Red Sea heat flow. Philosophical Transactions of the Royal Society of London, 267, 191203.Google Scholar
Gladczenko, T., Hinz, K., Eldholm, O., Meyer, H., Neben, S., and Skogseid, J. (1997). South Atlantic volcanic margins. Journal of the Geological Society of London, 154, 465–70.CrossRefGoogle Scholar
Glasby, G. P., Stuben, D., Jeschke, G., Stoffers, P., and Garbe-Schonberg, C.-D. (1997). A model for the formation of the hydrothermal manganese crusts from the Pitcairn Island Hotspot. Geochimica et Cosmochimica Acta, 61, 4583–97.CrossRefGoogle Scholar
Glasmann, J. R., Lundegard, P. D., Clark, R. A., Penny, B. K., and Collins, I. D. (1989). Geochemical evidence for the history of diagenesis and fluid migration: Brent Sandstone, Heather Field, North Sea. Clay Minerals, 24, 255–84.CrossRefGoogle Scholar
Glasspool, I. J., Edwards, D., and Axe, L. (2004). Charcoal in the Silurian as evidence for the earliest wildfire. Geology, 32, 381–3.CrossRefGoogle Scholar
Gleadow, A. J. W. and Fitzgerald, P. G. (1987). Uplift history and structure of the Transantarctic Mountains: new evidence from fission track dating of basement apatites in the Dry Valleys area, southern Victoria Land. Earth and Planetary Science Letters, 82, 114.CrossRefGoogle Scholar
Gleason, G. C. and Tullis, J. (1995). A flow law for dislocation creep of quartz aggregates determined with the molten salt cell. Tectonophysics, 247, 123.CrossRefGoogle Scholar
Glennie, K. W. and Armstrong, L. A. (1991). The Kittiwake Field, Block 21/18, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 339–45.Google Scholar
Goddard, J. V. and Evans, J. P. (1995). Chemical changes and fluid-rock interaction in faults of crystalline thrust sheets, northwestern Wyoming, U.S.A. Journal of Structural Geology, 17, 533–47.CrossRefGoogle Scholar
Godfrey, N. J., Beaudoin, B. C., Klemperthe, S. L., Levander, A., Luetgert, J., Meltzer, A., Mooney, W., and Trehu, A. (1997). Ophiolitic basement to the Great Valley forearc basin, California, from seismic and gravity data: implications for crustal growth at the North American continental margin. Geological Society of America Bulletin, 109, 1536–62.2.3.CO;2>CrossRefGoogle Scholar
Goetze, C. and Evans, B. (1979). Stress and temperature in the bending lithosphere as constrained by experimental rock mechanics. Geophysical Journal of the Royal Astronomic Society, 59, 463–78.CrossRefGoogle Scholar
Goff, F. (1980). Geology of the Geysers-Clear Lake geothermal regime, Northern California. In Continental Scientific Drilling Program Thermal Regimes: Comparative Site Assessment Geology of Five Magma- Hydrothermal Systems, ed. Goff, F. and Waters, A. C.. Los Alamos Scientific Laboratory, Report LA-8550-OBES, pp. 728.CrossRefGoogle Scholar
Goff, J. C. (1983). Hydrocarbon generation and migration from Jurassic source rocks in the E Shetland Basin and Viking Graben of the northern North Sea. Journal of the Geological Society of London, 140, 445–74.CrossRefGoogle Scholar
Goldflam, P., Hinz, K., Weigel, W., and Wissmann, G. (1980). Some features of the northwest African margin and magnetic quiet zone. Philosophical Transactions of the Royal Society of London, Series A: Mathematical and Physical Sciences, 294, 8796.Google Scholar
Golonka, J. (2000). Cambrian-Neogene Plate Tectonic Maps. Wydawnictwo Uniwersytetu Jagiellonskiego, Krakow, p. 125.Google Scholar
Gomes, J. R. C., Gatto, C. M. P. P., Souza, G. M. C., Luz, D. S., Pires, J. L., and Teixeira, W. (1981). Projeto Radambrasil, Folhas SB 24–25, Jaguaride – Natal. Geologia, geomorfologia, pedologia, vegetação e uso potencial da terra, 27, 741.Google Scholar
Gomes, P. O., Gomes, B. S., Palma, J. J. C., Jinno, K., and de Souza, J. M. (2000). Ocean-continent transition and tectonic framework of the oceanic crust at the continental margin of NE Brazil: results of LEPLAC Project. Geophysical Monograph, 115, 261–91.Google Scholar
Goolsby, S. M., Druyff, L., and Fryt, M. S. (1988). Trapping mechanisms and petrophysical properties of the Permian Kaibab Formation, south-central Utah. In Properties of Carbonate Reservoirs in the Rocky Mountain Region Occurrence and Petrophysical, ed. Goolsby, S. M. and Longman, M. W.. Denver, Rocky Mountain Association of Geologists, 1988 Guidebook, pp. 193210.Google Scholar
Goth, K., de Leeuw, J. W., Puttmann, W., and Tegelaar, E. W. (1988). Origin of Messel oil shale kerogen. Nature, 336, 759–61.CrossRefGoogle Scholar
Govers, R. and Wortel, M. J. R. (1995). Extension of stable continental lithosphere and the initiation of lithospheric scale faults. Tectonics, 14, 1041–55.CrossRefGoogle Scholar
Gowers, M. B., Holtar, E., and Swensson, E. (1993). The structure of the Norwegian Central Trough (Central Graben area). In Petroleum Geology of Northwest Europe, ed. J. R. Parker. Proceedings of the 4th Conference, The Geological Society of London, 1245–54.CrossRefGoogle Scholar
Grace, J. D. and Hart, G. F. (1986). Giant fields of northern West Siberia. American Association of Petroleum Geologists Bulletin, 70, 830–52.Google Scholar
Gradijan, S. J. and Wiik, M. (1987). Statfjord Nord. In Geology of the Norwegian Oil and Gas Fields, ed. Spencer, A. M., Campbell, C. J., Hanslien, S. H., Holter, E., Nelson, P. H. H., Nysaether, E., and Ormaasen, G.. Graham and Trotman, London, pp. 341–50.Google Scholar
Gradstein, F. M., Ogg, J. G., Smith, A. G., Bleeker, W., and Lourens, L. J. (2004). A new geologic time scale, with special reference to Precambrian and Neogene. Episodes, 27(2), 83100.CrossRefGoogle Scholar
Graham, S. A. (1976). Tertiary sedimentary tectonics of the central Salinian Block of California. PhD thesis, Stanford University, California, p. 510.Google Scholar
Graham, S. A. and Williams, L. A. (1985). Tectonic, depositional, and diagenetic history of Monterey Formation (Miocene), central San Joaquin Basin, California. American Association of Petroleum Geologists Bulletin, 69, 385411.Google Scholar
Gras, R. and Thusu, B. (1998). Trap architecture of the Early Cretaceous Sarir sandstone in the eastern Sirt Basin, Libya. In Petroleum Geology of North Africa, ed. MacGregor, D. S., Moody, R. T. J., and Clark-Lowes, D. D.. Geological Society of London Special Publications, 132, 317–34.Google Scholar
Grasemann, B. and Mancktelow, N. S. (1993). Two dimensional thermal modeling of normal faulting: the Simplon Fault Zone, central Alps, Switzerland. Tectonophysics, 225, 155–65.CrossRefGoogle Scholar
Gratier, J. P. and Guiguet, R. (1986). Experimental pressure solution-deposition on quartz grains: the crucial effect of the nature of the fluid. Journal of Structural Geology, 8, 845–56.CrossRefGoogle Scholar
Graue, K. (2000). Mud volcanoes in deepwater Nigeria. Marine and Petroleum Geology, 17, 959–74.CrossRefGoogle Scholar
Grauls, D. J. and Baleix, J. M. (1994). Role of overpressures and in situ stresses in fault-controlled hydrocarbon migration: a case study. Marine and Petroleum Geology, 11, 734–42.CrossRefGoogle Scholar
Gray, D. I. (1987). Troll. In Geology of the Norwegian Oil and Gas Fields, ed. Spencer, A. M., Campbell, C. J., Hanslien, S. H., Holter, E., Nelson, P. H. H., Nysaether, E., and Ormaasen, G.. Graham and Trotman, London, pp. 389401.Google Scholar
Green, P. F. (1981). A new look at statistics in fission-track dating. Nuclear Tracks, 5, 7786.CrossRefGoogle Scholar
Green, P. F., Crowhurst, P. V., and Duddy, I. R. (2004). Integration of AFTA and (U-Th)/He thermochronology to enhance the resolution and precision of the thermal history reconstruction in the Anglesea-1 well, Otway Basin, SE Australia. PESA Eastern Australian Basins Symposium II. Adelaide, September 19–22, 2004, pp. 117–31.Google Scholar
Green, P. F., Duddy, I. R., Gleadow, A. J. W., Tingate, P. R., and Laslett, G. M. (1986a). Fission track annealing in apatite: track length measurements and the form of the Arrhenius plot. Nuclear Tracks, 10, 323–8.Google Scholar
Green, P. F., Duddy, I. R., Gleadow, A. J. W., Tingate, P. R., and Laslett, G. M. (1986b). Thermal annealing of fission tracks in apatite 1. A qualitative description. Chemical Geology, 59, 237–53.CrossRefGoogle Scholar
Green, P. F., Duddy, I. R., and Hegarty, K. A. (2002). Quantifying exhumation from apatite fission-track analysis and vitrinite reflectance data: precision, accuracy and latest results from the Atlantic margin of NW Europe. In Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration, ed. Dore, A. G., Cartwright, J., Stoker, M. S., Turner, J. P., and White, N.. Geological Society of London Special Publications, 196, 331–54.Google Scholar
Green, P. F., Duddy, I. R., Hegarty, K. A., Gleadow, A. J. W., and Lovering, J. F. (1989). Thermal annealing of fission tracks in apatite 4. Quantitative modeling techniques and extention to geological timescales. Chemical Geology, 79, 155–82.Google Scholar
Green, W. V., Achauer, U., and Meyer, R. P. (1991). A three-dimensional seismic image of the crust and upper mantle beneath the Kenya Rift. Nature, 354, 199203.CrossRefGoogle Scholar
Gregory, A. R. (1977). Aspects of rock physics from laboratory and log data that are important to seismic interpretation. In Seismic Stratigraphy, Applications to Hydrocarbon Exploration, ed. Payton, C. E.. American Association of Petroleum Geologists Memoir, 26, 1546.Google Scholar
Gretener, P. E. (1981). Geothermics: using temperature in hydrocarbon exploration. American Association of Petroleum Geologists Short Course Notes, 17, 1156.Google Scholar
Griffin, W. L., O’Reilly, S. Y., Alonso, J. C., and Begg, G. C. (2009). The composition and evolution of lithospheric mantle: a re-evaluation and its tectonic implications. Journal of Petrology, 50, 1185–204.CrossRefGoogle Scholar
Griffin, W. L., O’Reilly, S. Y., and Ryan, C. G. (1999a). The composition and origin of subcontinental lithospheric mantle. In Mantle Petrology: Field Observations and High Pressure Experimentation: A Tribute to Francis F. (Joe) Boyd, ed. Y. Fei, C. M. Bertka, and B. O. Mysen. The Geochemical Society, pp. 13–45.Google Scholar
Griffin, W. L., Shee, S. R., Ryan, C. G., Win, T. T., Wyatt, B. A. (1999b). Harzburgite to lherzolite and back again: metasomatic processes in ultramafic xenoliths from the Wesselton kimberlite, Kimberley, South Africa. Contributions to Mineralogy and Petrology, 134, 232–50.CrossRefGoogle Scholar
Griffiths, R. W. and Campbell, I. H. (1991). Interaction of mantle plume heads with the Earth’s surface and onset of small-scale convection. Journal of Geophysical Research, 96, 18295–310.CrossRefGoogle Scholar
Grist, A. and Zentilli, M. (2000). Apatite fission track constraints on a widespread Early Cretaceous heating episode in the Canadian Atlantic margin: added effect of a climatically perturbed gradient? http://earthsciences.dal.ca/people/zentilli/zenabs13.htm.Google Scholar
Groschel-Becker, H. M. (1996). Formational processes of oceanic crust at sedimented spreading centers; perspectives from the West African continental margin and Middle Valley, Juan de Fuca Ridge. PhD thesis, University of Miami (Florida), Coral Gables.Google Scholar
Gross, G. W. and Wilcox, R. (1983). Groundwater circulation in the Socorro geothermal area. Field Conference Guidebook, New Mexico Geological Society, 34, 311–18.CrossRefGoogle Scholar
Grove, T. L. and Parman, S. W. (2004). Thermal evolution of the Earth as recorded by komatiites. Earth and Planetary Science Letters, 219, 173–87.CrossRefGoogle Scholar
Grunau, H. R. (1981). Worldwide review of seals for major accumulation of natural gas. American Association of Petroleum Geologists Bulletin, 65, 933.Google Scholar
Guardado, L. R., Gamboa, L. A. P., and Lucchesi, C. F. (1990). Petroleum geology of the Campos Basin, Brazil: a model for a producing Atlantic-type basin. In Divergent/Passive Margins, ed. Edwards, J. D. and Santogrossi, P. A.. American Association of Petroleum Geologists Memoir, 48, 379.Google Scholar
Guillou, H., Carracedo, J. C., Perez Torrado, F., and Rodriguez Badiola, E. (1996a). K-Ar ages and magnetic stratigraphy of ahotspot-induced, fast grown oceanic island: El Hiero, Canary Islands. Journal of Volvanology and Geothermal Research, 73, 141–55.Google Scholar
Guillou, H., Turpin, L. and Garcia, M. O. (1996b). Unspiked K-Ar dating of young submarine volcanic rocks from Loihi and Mauna Loa. Eos, Transactions, American Geophysical Union, 77, 812.Google Scholar
Guillou, H., Turpin, L., Garnier, F., Charbit, S., and Thomas, D. M. (1997). Unspiked K-Ar dating of Pleistocene tholeiitic basalts from the deep core SOH-4, Kilauea, Hawaii. Chemical Geology, 140, 81–8.CrossRefGoogle Scholar
Guillou-Frottier, L., Burov, E., Nehlig, P., and Wyns, R. (2007). Deciphering plume–lithosphere interactions beneath Europe from topographic signatures, Global and Planetary Change, 58, 119–40.CrossRefGoogle Scholar
Guiraud, R. and Bosworth, W. (1997). Senonian basin inversion and rejuvenation of rifting in Africa and Arabia: synthesis and implications to plate-scale tectonics. Tectonophysics, 282, 3982.CrossRefGoogle Scholar
Guiraud, R. and Maurin, J. C. (1991). Le Rifting en Afrique au Cretace inferieur: synthese structurale, mise en evidence de deux etapes dans la genese des bassins, relations avec les ouvertures oceaniques peri-africaines. Lower Cretaceous rifting in Africa: structural synthesis, evidence for two-stage basin formation, relationship to the opening of the peri-African oceans. Bulletin de la Societe Geologique de France, Huitieme Serie, 162, 811–23.Google Scholar
Guiraud, R. and Maurin, J. C. (1992). Early Cretaceous rifts of Western and Central Africa: an overview. Tectonophysics, 213, 153–68.CrossRefGoogle Scholar
Gulbrandsen, A. (1987). Agat. In Habitat of Hydrocarbons on the Norwegian Continental Shelf, ed. Spencer, A. M. and 6 others. Graham and Trotman, London, pp. 363–70.Google Scholar
Guliyev, I. S., Feyzullayev, A. A., and Tagiyev, M. F. (2001a). Source potential of the Mesozoic-Cenozoic rocks in the South Caspian Basin and their role in forming the oil accumulations in the Lower Pliocene reservoirs. Petroleum Geoscience, 7, 409–17.Google Scholar
Guliyev, I. S., Tagiyev, M. F., and Feyzullayev, A. A. (2001b). Geochemical characteristics of organic matter from Maykop rocks of eastern Azerbaijan. Lithology and Mineral Resources, 36, 280–5.CrossRefGoogle Scholar
Gunnell, Y., Gallagher, K., Carter, A., Widdowson, M., and Hurford, A. J. (2003). Denudation history of the continental margin of western peninsular India since the early Mesozoic – reconciling apatite fission-track data with geomorphology. Earth and Planetary Science Letters, 215, 187201.CrossRefGoogle Scholar
Gunnell, Y. and Radhakrishna, B. P. (2001). Shyadri, the Great Escarpment of the Indian subcontinent. Patterns of landscape development in the Western Ghats. Geological Society of India Memoir, 47, 1054.Google Scholar
Gupta, S., Cowie, P. A., Dawers, N. H., and Underhill, J. R. (1998). A mechanism to explain rift-basin subsidence and stratigraphic patterns through fault-array evolution. Geology, 26, 595–8.2.3.CO;2>CrossRefGoogle Scholar
Gupta, S., Underhill, J. R., Sharp, I. R., and Gawthorpe, R. L. (1999). Role of fault interactions in controlling synrift sediment dispersal patterns: Miocene, Abu Alaqa Group, Suez Rift, Sinai, Egypt. Basin Research, 11, 167–89.CrossRefGoogle Scholar
Gussow, W. C. (1954). Differential entrapment of gas and oil: a fundamental principle. American Association of Petroleum Geologists Bulletin, 38, 816–53.Google Scholar
Gustavson Associates, Inc. (1992). Petroleum Geology and Exploration Potential in the Former Soviet Republics. Gustavson Associates, Inc., Boulder.Google Scholar
Guterch, A., Grad, M., Janik, T., Materzok, R., Luosto, U., Yliniemi, J., Lück, E., Schulze, A., and Förste, K. (1994). Crustal structure of the transition zone between Precambrian and Variscan Europe from new seismic data along LT-7 profile (NW Poland and eastern Germany). Comptes Rendus de l’Academie des Sciences, Serie II, Sciences de la Terre et des Planetes, 319, 1489–96.Google Scholar
Guterch, A., Kowalski, T., Materzok, R., and Toporkiewicz, S. (1976). Seismic refraction study of the Earth’s crust in the Teisseyre-Tornquist line zone in Poland along the regional profile LT-2. Publications of the Institute of Geophysics, Series A: Physics of the Earth Interior, 2, 1523.Google Scholar
Guzman, A. and Marquez-Dominguez, B. (2001). The Gulf of Mexico Basin south of the border: the petroleum province of the twenty-first century. In Petroleum Provinces of the Twenty-first Century, ed. Downey, M. W., Threet, J. C. and Morgan, W. A.. American Association of Petroleum Geologists Memoir, 74, 337–51.Google Scholar
Gvirtzman, H., Garven, G., and Gvirtzman, G. (1997). Thermal anomalies associated with forced and free ground-water convection in the Dead Sea rift valley. Geological Society of America Bulletin, 109, 1167–76.2.3.CO;2>CrossRefGoogle Scholar
Gvirtzman, H. and Stanislavsky, E. (2000). Paleohydrology of hydrocarbon maturation, migration and accumulation in the Dead Sea rift. Basin Research, 12, 7993.CrossRefGoogle Scholar
GX TECHNOLOGY (2005a). EquatorSPAN gravity and magnetic maps. Archive of GX Technology, Houston.Google Scholar
GX TECHNOLOGY (2005b). EquatorSPAN in Gabon. Archive of GX Technology, Houston.Google Scholar
GX TECHNOLOGY (2005c). CongoSPAN in Congo. Archive of GX Technology, Houston.Google Scholar
Haack, R. C., Sundararaman, P., Diedjomahor, J. O., Xiao, H., Gant, N. J., May, E. D., and Kelsch, K. (2000). Niger Delta petroleum systems, Nigeria. In Petroleum Systems of South Atlantic Margins, ed. Mello, M. R. and Katz, B. J.. American Association of Petroleum Geologists Memoir, 73, 213–31.Google Scholar
Haberland, C., Agnon, A., El-Kelani, R., Naercklin, N., Qabbani, I., Rumpker, G., Ryberg, T., Scherbaum, F., and Weber, M. (2003). Modeling of seismic guided waves at the Dead Sea Transform. Journal of Geophysical Research, 108.CrossRefGoogle Scholar
Habermehl, G. and Hundrieser, H. J. (1983). Fossile Relikte der “Wasserblute” im Messeler Olschiefer. Naturwissenschaften, 70, 566–7.CrossRefGoogle Scholar
Habib, D. (1982). Sediment supply origin of Cretaceous black shales. In Nature and Origin of Cretaceous Carbon-rich Facies, ed. Schlanger, S. O. and Cita, M. B.. London, Academic Press, pp. 113–27.Google Scholar
Haenel, R., Rybach, L., and Stegena, L. (1988). Handbook of Terrestrial Heat-Flow Density Determination. Kluwer Academic Publishers, Dordrecht.CrossRefGoogle Scholar
Hall, S. A. (1992). The Angus Field, a subtle trap. In Exploration Britain: Geological Insights into the Next Decade, ed. Hardman, R. F. P.. Special Publication of Geological Society of London, 67, 151–85.Google Scholar
Hallam, A. (1984). Pre-Quaternary sea-level changes. Annual Review of Earth and Planetary Science, 12, 205–43.CrossRefGoogle Scholar
Halliday, A. N., Dickin, A. P., Fallick, A. E., and Fitton, J. G. (1988). Mantle dynamics – a Nd, Sr, Pb and O isotopic study of the Cameroon Line volcanic chain. Journal of Petroleum, 29, 181211.Google Scholar
Hamblin, W. K. (1965). Origin of “reverse drag” on the downthrown side of normal faults. Geological Society of America Bulletin, 76, 1145–64.CrossRefGoogle Scholar
Hamblin, W. K. (1976). Patterns of displacement along the Wasatch Fault. Geology, 4, 619–22.2.0.CO;2>CrossRefGoogle Scholar
Hamdani, Y., Mareschal, J.-C., and Arkani-Hamed, J. (1994). Phase change and thermal subsidence of the Williston Basin. Geophysical Journal International, 116, 585–97.CrossRefGoogle Scholar
Hames, W. E., Renne, P. R., and Ruppel, C. (2000). New evidence for geologically-instantaneous emplacement of earliest Jurassic Central Atlantic magmatic province basalts on the North American margin. Geology, 28, 859–62.2.0.CO;2>CrossRefGoogle Scholar
Hamilton, R. M., Behrendt, J. C., and Ackermann, H. D. (1983). Land multichannel seismic-reflection evidence for tectonic features near Charleston, South Carolina. In Studies Related to the Charleston, South Carolina, Earthquake of 1886 – Tectonics and Seismicity, ed. Gohn, G. S.. Geological Survey Professional Paper, 1313, 118.Google Scholar
Hamilton, W. B. (1987). Crustal extension in the Basin and Range Province, Southwestern United States. Geological Society Special Publications, 28, 155–76.CrossRefGoogle Scholar
Hammond, D. E., Leslie, B. W., and Ku, T.-L. (1988). 222Rn concentrations in deep formation waters and the geohydrology of the Cajon Pass borehole. Geophysical Research Letters, 15, 1045–8.CrossRefGoogle Scholar
Hammond, W. C. and Thatcher, W. (2005). Nortwest Basin and Range tectonic deformation observed with the Global Positioning System, 1999–2003. Journal of Geophysical Research, 110 (B10405).CrossRefGoogle Scholar
Hampson, G. J. (2004). Facies architecture and stratigraphy of: stray “shelf” sandstones, Late Cretaceous Mancos Shale, northern Utah and Colorado. AAPG Annual Meeting Expanded Abstracts, 13, 57058.Google Scholar
Hampson, G. J., Rodriguez, A. B., Storms, J. E. A., Johnson, H. D., and Meyer, C. T. (2008). Geomorphology and high-resolution stratigraphy of progradational wave-dominated shoreline deposits: impact on reservoir-scale facies. In Recent Advances in Models of Siliciclastic Shallow Marine Stratigraphy, ed. Hampson, G. J., Steel, R. J., Burgess, P. M., and Dalrymple, R. W.. Special Publication of Society for Sedimentary Geology, 90, 117–42.CrossRefGoogle Scholar
Hampson, G. J., Sixsmith, P. J., and Johnson, H. D. (2004). A sedimentological approach to refining reservoir architecture in a mature hydrocarbon province: the Brent Province, UK North Sea. Marine and Petroleum Geology, 21, 457–84.CrossRefGoogle Scholar
Hancock, N. J. and Fisher, M. J. (1981). Middle Jurassic North Sea deltas with particular reference to Yorkshire. In Petroleum Geology of the Continental Shelf of North-West Europe, ed. Illing, L. V. and Hobson, G. D.. Heyden and Son, London, pp. 186–95.Google Scholar
Handy, M. R. and Stuenitz, H. (2002). Strain localization by fracturing and reaction weakening: a mechanism for initiating exhumation of subcontinental mantle beneath rifted margins. Geological Society Special Publications, 200, 387407.CrossRefGoogle Scholar
Handy, M. R. and Zingg, A. (1991). The tectonic and rheological evolution of an attenuated cross section of the continental crust: Ivrea crustal section, southern Alps, northwestern Italy and southern Switzerland. Geological Society of America Bulletin, 103, 236–53.2.3.CO;2>CrossRefGoogle Scholar
Hannington, M., Herzig, P., Stoffers, P., Scholten, J., Botz, R., Garbe-Schonberg, D., Jonasson, I. R., Roest, W., and Shipboard Scientific Party (2001). First observations of high-temperature submarine hydrothermal vents and massive anhydrite deposits off the north coast of Iceland. Marine Geology, 177, 199220.CrossRefGoogle Scholar
Hanor, J. S. (1979). Sedimentary genesis of hydrothermal fluids. In Geochemistry of Hydrothermal Ore Deposits, ed. Barnes, H. L.. John Wiley, New York, pp. 137–68.Google Scholar
Hanor, J. S. (1987). Kilometer-scale thermohaline overturn of pore fluid in the Louisiana Gulf Coast. Nature, 327, 501–3.CrossRefGoogle Scholar
Hansen, D. L. and Nielsen, S. B. (2002). Does thermal weakening explain basin inversion? Stochastic modelling of the thermal structure beneath sedimentary basins. Earth and Planetary Science Letters, 198, 113–27.CrossRefGoogle Scholar
Hao, F., Li, S. T., Sun, Y. C., and Zhang, Q. M. (1998). Geology, compositional heterogeneities and geochemical origin of the Yacheng gas field in the Qiongdongnan Basin, South China Sea. American Association of Petroleum Geologists Bulletin, 82, 1372–84.Google Scholar
Hao, F., Li, S. T., Gong, Z. S., and Yang, J. M. (2000). Thermal regime, inter-reservoir compositional heterogeneities, and reservoir-filling history of the Dongfang gas field, Yinggehai Basin, South China: Evidence for episodic fluid injections in overpressured basins? American Association of Petroleum Geologists Bulletin, 84, 607–26.Google Scholar
Hao, F., Zhou, X., Zhu, Y., Bao, X., and Yang, Y. (2009). Charging of the Neogene Penglai 19-3 field, Bohai Bay Basin, China: oil accumulation in a young trap in an active fault zone. American Association of Petroleum Geologists Bulletin, 93, 155–79.CrossRefGoogle Scholar
Haq, B. U., Hardenbol, J., and Vail, P. R (1987). Chronology of fluctuating sea levels since the Triassic. Science, 235, 1156–67.CrossRefGoogle ScholarPubMed
Hardarson, B. S., Fitton, J. G., Ellam, R. M., and Pringle, M. S. (1997). Rift relocation – a geochemical and geochronological investigation of a paleo-rift in northwest Iceland. Earth and Planetary Science Letters, 153, 181–96.CrossRefGoogle Scholar
Harding, T. P. (1984). Graben hydrocarbon occurrences and structural styles. American Association of Petroleum Geologists Bulletin, 68, 333–62.Google Scholar
Hardy, S. and Gawthorpe, R. L. (1998). Effects of variations in fault slip rate on sequence stratigraphy in fan deltas; insights from numerical modeling. Geology , 26, 911–14.2.3.CO;2>CrossRefGoogle Scholar
Harker, S. D. (1998). The palingenesy of the Piper oil field, UK North Sea. Petroleum Geoscience, 4, 271–86.CrossRefGoogle Scholar
Harker, S. D. and Chermak, A. (1992). Detection and prediction of Lower Cretaceous sandstone distribution in the Scapa field, North Sea. In Exploration Britain: Geological Insights into the Next Decade, ed. Hardman, R. F. P.. Geological Society of London Special Publications, 67, 221–46.Google Scholar
Harker, S. D., Green, S. C. H., and Romani, R. S. (1991). The Claymore field, Block 14/19, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 269–78.Google Scholar
Harp, E. L., Wells, W. G., and Sarmiento, J. G. (1990). Pore pressure response during failure in soils. Geological Society of America Bulletin, 102, 428–38.2.3.CO;2>CrossRefGoogle Scholar
Harris, N. B., Freeman, K. H., Pancost, R. D., White, T. S., and Mitchell, G. D. (2004). The character and origin of lacustrine source rocks in the Lower Cretaceous synrift section, Congo Basin, West Africa. American Association of Petroleum Geologists Bulletin, 88, 1163–84.CrossRefGoogle Scholar
Harris, R. N., Von Herzen, R. P., McNutt, M. K., Garven, G., and Jordahl, K. (2000). Submarine hydrogeology of the Hawaiian archipelagic apron 1. Heat flow patterns north of Oahu and Maro Reef. Journal of Geophysical Research, 105, 21353–69.Google Scholar
Harrison, T. M., Copeland, P., Hall, S. A., Quade, J., Burner, S., Ojha, T. P., and Kidd, W. S. F. (1993). Isotopic preservation of Himalayan/Tibetan uplift, denudation, and climatic histories of two molasse deposits. Journal of Geology, 101, 157–75.CrossRefGoogle Scholar
Harry, D. L. and Sawyer, D. S. (1992). Basaltic volcanism, mantle plumes, and the mechanics of rifting: the Paraná flood basalt province of South America. Geology, 20, 207–10.2.3.CO;2>CrossRefGoogle Scholar
Hasebe, N., Barbarand, J., Jarvis, K., Carter, A., and Hurford, A. J. (2004). Apatite fission-track chronometry using laser ablation ICP-MS. Chemical Geology, 207, 135–45.CrossRefGoogle Scholar
Haszeldine, R. S., Wilkinson, M., Darby, D., Macaulay, C. I., Couples, G. D., Fallick, A. E., Fleming, C. G., Stewart, R. N. T., and McAulay, G. (1999). Diagenetic porosity creation in an overpressured graben. The Geological Society of London, 5, 1339–50.Google Scholar
Hatcher, R. D., Osberg, P. H., and Drake, A. A. (1986). Tectonic Map of the U.S. Appalachians, DNAG Appalachians – Ouachitas Volume, Plate 1.Google Scholar
Hathaway, J. C. and Degens, E. T. (1969). Methane-derived marine carbonates of Pleistocene age. Science, 165, 609–92.CrossRefGoogle ScholarPubMed
Haughton, P. D. W., McCaffrey, W. D., Davis, C., and Barker, S. P. (2008). Sediment gravity flow deposits and bed-scale heterogeneity: lessons from North Sea fields. In Answering the Challenges of Production from Deep-water Reservoirs: Analogues and Case Histories to Aid a New Generation, ed. K. Schofield, N. C. Rosen, and D. Pfeiffer. Papers presented at the Gulf Coast Section, Society of Economic Paleontologists and Mineralogists Foundation Annual Bob F. Perkins Research Conference, 28, 407.Google Scholar
Hawkesworth, C. J., Gallagher, K., Kelley, S., Mantovani, M. S. M., Peate, D. W., Regelous, M., and Rogers, N. W. (1992). Parana magmatism and the opening of the South Atlantic. In Magmatism and the Causes of Continental Break-up, ed. B. Storey, A. Alabaster, and R. Pankhurst. Geological Society [London] Special Publication 68, pp. 221–40.Google Scholar
Haworth, R. T., Daniels, D. L., Williams, H., and Zietz, I. (1980). Bouguer gravity anomaly map of the Appalachian Orogen – Map No. 3. Memorial University of Newfoundland, St John’s, Newfoundland.Google Scholar
Haxby, W. F. (1985). Gravity Field of the World’s Oceans. Lamont-Doherty Geological Observatory, Columbia University, Palisades, New York. (http://www.ngdc.noaa.gov).Google Scholar
Hayward, N. J. and Ebinger, C. J. (1996). Variations in the along-axis segmentation of the Afar Rift system. Tectonics, 15, 244–57.CrossRefGoogle Scholar
He, B., Xu, Y.-G., Chung, S.-L., Xiao, L., and Wang, Y. (2003). Sedimentary evidence for a rapid doming prior to the eruption of the Emeishan flood basalts. Earth and Planetary Science Letters, 213, 389403.CrossRefGoogle Scholar
Hébert, H., Deplus, C., Huchon, P., Khanbari, K., and Audin, L. (2001). Lithospheric structure of a nascent spreading ridge inferred from gravity data: the western Gulf of Aden. Journal of Geophysical Research, 106, 26345–63.CrossRefGoogle Scholar
Hecker, S. (1993). Quaternary tectonics of Utah with emphasis on earthquake hazard characterization. Utah Geological Survey Bulletin, 127, 1157.Google Scholar
Hedenquist, J. W., Arribas, A. Jr., and Reynolds, T. J. (1998). Evolution of an intrusion – centered hydrothermal system: Far Southeast – Lepanto porphyry and epithermal Cu-Au deposits, Philippines. Economic Geology, 93, 373404.CrossRefGoogle Scholar
Heggland, R. (1997). Detection of gas migration from a deep source by the use of exploration 3D seismic data. Marine Geology, 137, 41–7.CrossRefGoogle Scholar
Heggland, R. (1998). Gas seepage as an indicator of deeper prospective reservoirs: a study based on exploration 3-D seismic data. Marine and Petroleum Geology, 15, 19.CrossRefGoogle Scholar
Hein, J. R., Koski, R. A., Embley, R. W., Reid, J., and Chang, S.-W. (1999). Diffuse-flow hydrothermal field in an oceanic fracture zone setting, northeast Pacific: deposit composition. Exploration and Mining Geology, 8, 299322.Google Scholar
Hekinian, R., Hoffert, M., Larque, P., Cheminee, J. L., Stoffers, P., and Bideau, D. (1993). Hydrothermal Fe and Si oxyhydroxide deposits from the South Pacific intraplate volcanoes and East Pacific Rise axial and off-axial regions. Economic Geology, 88, 2099–121.CrossRefGoogle Scholar
Helgeson, D. E. (1999). Structural development and trap formation in the central North Sea HP/HT play. The Geological Society of London, 5, 1029–34.Google Scholar
Heling, D. and Teichmuller, M. (1974). The transition zone between montmorillonite and mixed-layer minerals and its relation to coalification in the Graue Beds of the Oligocene in the Upper Rhine Graben. Fortschritte in der Geologie von Rheinland und Westfalen, Inkohlung und Erdoel, 24, 113–28.Google Scholar
Helland-Hansen, W., Ashton, M., Lomo, L., and Steel, R. (1992). Advance and retreat of the Brent delta: recent contributions to the depositional model. In Geology of the Brent Group, ed. Morton, A. C., Haszeldine, R. S., Giles, M. R., and Brown, S.. Geological Society of London Special Publications, 61, 109–28.Google Scholar
Hellinger, S. J. and Sclater, J. G. (1983). Some remarks on two-layer extensional models or the evolution of sedimentary basins. Journal of Geophysical Research, 88, 8251–69.CrossRefGoogle Scholar
Hempton, M. R. and Neher, K. (1986). Experimental fracture, strain and subsidence patterns over en echelon strike-slip faults: implications for the structural evolution of pull-apart basins. Journal of Structural Geology, 8, 597605.CrossRefGoogle Scholar
Hendrie, D. B., Kusznir, N. J., Morley, C. K., and Ebinger, C. J. (1994). Cenozoic extension in northern Kenya: a quantitative model of rift basin development in the Turkana region. Tectonophysics, 236, 409–38.CrossRefGoogle Scholar
Henk, A. (2006). Impact of crustal rheology on subsidence history and basin geometry- numerical experiments of half-graben formation. EGU General Assembly. University of Freiburg, Germany, Wien.Google Scholar
Henry, A. A. and Lewan, M. D. (2001). Comparison of kinetic-model prediction of deep gas generation. In Geologic Studies of Deep Natural Gas Resources, ed. Dyman, T. S. and Kuuskraa, V. A.. U.S. Geological Survey Digital Data Series, 67, 125.Google Scholar
Henry, S., Danforth, A., and Venkatraman, S. (2005). New insights into petroleum systems and plays in Angola, the Congo and Gabon from PDSM sub-salt imaging. In Petroleum Systems of Divergent Continental Margin Basin, ed. P. J. Post, N. Rosen, D. L. Olson, S. L. Palmes, K. T. Lyons, and G. B. Newton. GCSSEPM25th Annual Bob F. Perkins Research Conference, SEPM Foundation, Houston, 25, 49–55.Google Scholar
Hermanrud, C. (1986). On the importance to the petroleum generation of heating effects from compaction-derived water: an example from the northern North Sea. In 1st IFP Exploration Research Conference, ed. J. Burrus. Technip, Paris.Google Scholar
Herzig, C. T., Mehegan, J. M., and Stelting, C. E. (1988). Lithostratigraphy of the State 2–14 Borehole; Salton Sea Scientific Drilling Project. Journal of Geophysical Research, 93, 12969–80.Google Scholar
Hesthammer, J. and Fossen, H. (1999). Evolution and geometries of gravitational collapse structures with examples from the Statfjord Field, northern North Sea. Marine and Petroleum Geology, 16, 259–81.CrossRefGoogle Scholar
Hesthammer, J., Jourdan, C. A., Nielsen, P. E., Ekern, T. E., and Gibbons, K. A. (1999). A tectonostratigraphic framework for the Statfjord Field, northern North Sea. Petroleum Geoscience, 5, 241–56.CrossRefGoogle Scholar
Hickman, S., Sibson, R. H., and Bruhn, R. (1995). Introduction to special section: mechanical involvement of fluids in faulting. Journal of Geophysical Research, 100, 12831–40.CrossRefGoogle Scholar
Hill, D. P., Eaton, J. P., and Jones, L. M. (1990). Seismicity 1980–86. US Geological Survey.Google Scholar
Hillier, R. D. and Cosgrove, J. W. (2002). Core and seismic observations of overpressure-related deformation within Eocene sediments of the Outer Moray Firth, UKCS. Petroleum Geoscience, 8, 141–9.CrossRefGoogle Scholar
Hindle, A. D. (1997). Petroleum migration pathways and charge concentration: a three-dimensional model. American Association of Petroleum Geologists Bulletin, 81, 1451–81.Google Scholar
Hinz, K., Neben, S., Schreckenberger, B., Roeser, H. A., Block, M. et al. (1999). The Argentine continental margin north of 48 degrees S: sedimentary successions, volcanic activity during breakup. Marine and Petroleum Geology, 16, 125.CrossRefGoogle Scholar
Hinz, K. and Weber, J. (1976). Zum geologischen Augbau des Norwegischen Kontinental Randes und der Barents Sea nach Reflections seismichen Messungen. Erdol und Kohle, Erdgas, Petrochemie, 75–6, 329.Google Scholar
Hirth, G. and Kohlstedt, D. L. (1996). Water in the oceanic upper mantle: implications for rheology, melt extraction and the evolution of the lithosphere. Earth and Planetary Science Letters, 144, 93108.CrossRefGoogle Scholar
Hirth, G. and Kohlstedt, D. L. (2004). Rheology of the upper mantle and the mantle wedge: a view from the experimentalists. In Inside the Subduction Factory, ed. Eiler, J.. AGU Monograph, 138, 83105.Google Scholar
Hitchon, B. (1984). Geothermal gradients, hydrodynamics, and hydrocarbon occurrences, Alberta, Canada. American Association of Petroleum Geologists Bulletin, 68, 713–43.Google Scholar
Hite, R. J. and Anders, D. E. (1991). Petroleum and evaporates. In Evaporites, Petroleum and Mineral Resources, ed. Melvin, J. L.. Amsterdam: Elsevier, pp. 349411.CrossRefGoogle Scholar
Hochstein, M. P. and Browne, P. R. L. (2000). Surface manifestations of geothermal systems with volcanic heat sources. In Encyclopedia of Volcanoes, ed. Sigurdsson, H., Houghton, B. F., McNutt, S. R., Rymer, H., and Stix, J.. Academic Press, San Diego, pp. 835–55.Google Scholar
Hodgson, N. A., Farnsworth, J., and Fraser, A. J. (1992). Salt-related tectonics, sedimentation and hydrocarbon plays in the Central Graben, North Sea, UKCS. 31–63. In Exploration Britain: Geological Insights into the Next Decade, Hardman, R. F. P.. Special Publication of the Geological Society of London, 67.Google Scholar
Hodkinson, R. A., Stoffers, P., Scholten, J., Cronan, D. S., Jeschke, G., and Roger, T. D. S. (1994). Geochemistry of hydrothermal manganese deposits from the Pitcairn Island hotspot, southeastern Pacific. Geochimica et Cosmochimica Acta, 58, 5011–29.CrossRefGoogle Scholar
Hoefs, J. (1981). Isotopic composition of the ocean-atmospheric system in the geologic past. In Evolution of the Earth, ed. O’Connell, R. J. and Fyfe, W. S.. Geodynamic Series, 5, 110–19.Google Scholar
Hofmann, C., Courtillot, V., Feraud, F., Rochette, P., Yirgu, G., Ketefo, E., and Pik, R. (1997). Timing of the Ethiopian flood basalt event: implications for plume birth and global change. Nature, 389, 838–40.CrossRefGoogle Scholar
Hohndorf, A., Haenel, R., and Giesel, W. (1975). Geothermal models of the Ivrea zone. Journal of Geophysics, 41, 179–87.Google Scholar
Hoholick, J. D., Metarko, T., and Potter, P. E. (1984). Regional variation of porosity and cement: St. Peter and Mount Simon sandstones in Illinois Basin. American Association of Petroleum Geologists Bulletin, 68, 753–64.Google Scholar
Holbrook, W. S. and Kelemen, P. B. (1993). Large igneous province on the US Atlantic margin and implications for magmatism during continental break-up. Nature, 364, 433–6.CrossRefGoogle Scholar
Holbrook, W. S., Larsen, H. C., Korenaga, J., Dahl-Jensen, T., Reid, L. D., Kelemen, P. B., Hopper, J. R., Kent, G. M., Lizarralde, D., Bernstein, S., and Detrick, R. S. (2001). Mantle thermal structure and active upwelling during continental breakup in the North Atlantic. Earth and Planetary Science Letters, 190, 251–66.CrossRefGoogle Scholar
Holbrook, W. S., Mooney, W. D., and Christensen, N. I. (1992). The seismic velocity structure of the deep continental crust. In Continental Lower Crust, ed. Fountain, D. M., Arculus, R., and Kay, R. W.. Developments in Geotectonics, Elsevier, pp. 144.Google Scholar
Holcombe, T. L., Bryant, W. R., Bouma, A. H., Taylor, L. A., and Liu, J. Y. (2002). Northern Gulf of Mexico bathymetry and feature names. Gulf Coast Association of Geological Societies, Technical Papers and Abstracts, 52, 397405.Google Scholar
Hold, I. M., Brussee, N. J., Schouten, S., and Sinninghe Damste, J. S. (1998). Changes in the molecular structure of a Type II-S kerogen (Monterey Formation, U.S.A.) during sequential chemical degradation. Organic Geochemistry, 29, 1403–17.CrossRefGoogle Scholar
Holmes, M. A. (1998). Thermal diagenesis of Cretaceous sediment recovered at the Côte d’Ivoire Ghana transform margin. In Proceedings of the Ocean Drilling Program, ed. J. Mascle, G. P. Lohmann, and M. Moullade. Scientific Results, 159, 53–70.Google Scholar
Holt, W. E. and Stern, T. A. (1991). Sediment loading on the western platform of the New Zealand continent: implication for the strength of a continental margin. Earth and Planetary Science Letters, 107, 523–38.CrossRefGoogle Scholar
Homberg, C., Hu, J. C., Angelier, J., Bergerat, F., and Lacombe, O. (1997). Characterization of stress perturbations near major fault zones: insights from 2-D distinct-element numerical modelling and field studies (Jura Mountains). Journal of Structural Geology, 19, 703–18.CrossRefGoogle Scholar
Hooper, E. C. D. (1991). Fluid migration along growth faults in compacting sediments. Journal of Petroleum Geology, 14, 161–80.CrossRefGoogle Scholar
Hooper, R. J., Fitzsimmons, R. J., Grant, N., and Vendeville, B. C. (2002). The role of deformation in controlling depositional patterns in the south-central Niger Delta, West Africa. Journal of Structural Geology, 24, 847–59.CrossRefGoogle Scholar
Hopper, J. R. and Buck, W. R. (1996). The effect of lower crustal flow on continental extension and passive margin formation. Journal of Geophysical Research, 101, 20175–94.CrossRefGoogle Scholar
Hopper, J. R. and Buck, W. R. (1998). Styles of extensional decoupling. Geology, 26, 699702.2.3.CO;2>CrossRefGoogle Scholar
Hopper, J. R., Funck, T., Tucholke, B. E., Larsen, H. C., Holbrook, W. S., Louden, K. E., Shillington, D., and Lau, H. (2004). Continental breakup and the onset of ultraslow seafloor spreading off Flemish Cap on the Newfoundland rifted margin. Geology, 32, 93–6.CrossRefGoogle Scholar
Hopper, J. R., Mutter, J. C., Larson, R. L., and Mutter, C. Z. (1992). Magmatism and rift margin evolution: evidence from northwest Australia. Geology, 20, 853–7.2.3.CO;2>CrossRefGoogle Scholar
Horn, M. K. (2003). Giant fields 1868–2003 (CD-ROM). In Giant Oil and Gas Fields of the Decade 1990–1999, ed. M. K. Halbouty. American Association of Petroleum Geologists Bulletin, 78, 340.Google Scholar
Horn, M. K. (2004). Giant fields 1868–2004 (CD-ROM). AAPG/Datapages Miscellaneous Data Series, version 1.2, revision of Horn, 2003.Google Scholar
Horn, P., Muller-Sohnius, D., and Schult, A. (1988). Potassium-argon ages on a Mesozoic tholeiitic dike swarm in Rio Grande do Norte, Brazil. Rev. Bras. De Geociences, 18, 50–3.Google Scholar
Hornafius, J. S., Quigley, D., and Luyendyk, B. P. (1999). The world’s most spectacular marine hydrocarbon seeps (Coal Oil Point, Santa Barbara Channel, California): quantification of emissions. Journal of Geophysical Research, 104, 20703–11.CrossRefGoogle Scholar
Horsfield, B., Curry, D. J., Bohacz, K., Littke, R., Rullkötter, J., Schenk, H. J., Radke, M., Schaefer, R. G., Carroll, A. R., Isaksen, G., and Witte, E. G. (1994). Organic geochemistry of freshwater and alkaline lacustrine environments, Green River Formation, Wyoming. Organic Geochemistry, 22, 415–40.CrossRefGoogle Scholar
Horsfield, W. T. (1980). Contemporaneous movement along crossing conjugate normal faults. Journal of Structural Geology, 2, 305–10.CrossRefGoogle Scholar
Horváth, F. (1993). Towards a mechanical model for the formation of the Pannonian Basin. Tectonophysics, 226, 333–57.CrossRefGoogle Scholar
House, M. A., Kohn, B. P., Farley, K. A., and Raza, A. (2000). (U-Th)/He thermochronometry in southeastern Australia; confirmation of laboratory diffusion experiments and insights into the Cenozoic thermal history of the Otway Basin. Abstracts – Geological Society of Australia, 58, 167–71.Google Scholar
House, M. A., Wernicke, B. P., Farley, K. A., and Dumitru, T. A. (1997). Cenozoic thermal evolution of the central Sierra Nevada, California, from (U-Th)/He thermochronometry. Earth and Planetary Science Letters, 151, 167–79.CrossRefGoogle Scholar
Houseman, G. and England, P. (1986). A dynamical model of lithosphere extension and sedimentary basin formation. Journal of Geophysical Research, 91, 719–29.CrossRefGoogle Scholar
Houseman, G. A., McKenzie, D. P., and Molnar, P. (1982). Convective instability of a thickened boundary layer and its relevance for the thermal evolution of continental convergent belts. Journal of Geophysical Research, 86, 6115–32.Google Scholar
Hovland, M., Croker, P. F., and Martin, M. (1994). Fault-associated seabed mounds (carbonate knolls?) off western Ireland and north-west Australia. Marine and Petroleum Geology, 11, 232–46.CrossRefGoogle Scholar
Howell, D. G., Crouch, J. K., Greene, H. G., McCulloch, D. S., and Vedder, J. G. (1980). Basin development along the late Mesozoic and Cainozoic California margin: a plate tectonic margin of subduction, oblique subduction and transform tectonics. Special Publication of the International Association of Sedimentologists, 4, 4362.Google Scholar
Hoye, T., Damsleth, E., and Hollund, K. (1994). Stochastic modeling of Troll West, with special emphasis on the thin oil zone. In Stochastic Modeling and Geostatistics: Principles, Methods and Case Studies, ed. Yarus, J. M. and Chambers, R. L.. AAPG Computer Applications in Geology, 3, 217–39.Google Scholar
Huang, D., Li, J., Zhang, D., Huang, X., and Zhou, Z. (1991). Maturation sequence of Tertiary crude oils in the Qaidam basin and its significance in petroleum resource assessment. Journal of Southeast Asian Earth Sciences, 5, 359–66.Google Scholar
Huang, H. P. and Pearson, J. M. (1999). Source rock paleoenvironments and controls on the distribution of dibenzothiophenes in lacustrine crude oils, Bohai Bay Basin, eastern China. Organic Geochemistry, 30, 1455–70.CrossRefGoogle Scholar
Huang, Q. (1988). Geometry and tectonic significance of Albian sedimentary dykes in the Sisteron area, SE France. Journal of Structural Geology, 10, 453–62.CrossRefGoogle Scholar
Huang, Z., Williamson, M. A., Fowler, M. G., and McAlpine, K. D. (1994). Predicted and measured petrophysical and geochemical characteristics of the Egret Member oil source rock, Jeanne d’Arc Basin, offshore eastern Canada. Marine and Petroleum Geology, 11, 294306.CrossRefGoogle Scholar
Hubbert, M. K. (1940). The theory of ground-water motion. Journal of Geology, 48, 785944.CrossRefGoogle Scholar
Hubbert, M. K. (1953). Entrapment of petroleum under hydrodynamic conditions. American Association of Petroleum Geologists Bulletin, 37, 19542026.Google Scholar
Hubbert, M. K. (1969). The Theory of Ground-Water Motion and Related Papers. Hafner Publication, Co New York-London.Google Scholar
Hubbert, M. K. and Willis, D. G. (1955). Important fractured reservoirs in the United States. Proceedings of the World Petroleum Congress, Actes et Documents, Congres Mondial du Petrole, pp. 57–81.Google Scholar
Huc, A. Y. (1988). Sedimentology of organic matter. In Humic Substances and Their Role in the Environment: Report of the Dahlem Workshop, ed. F. H. Frimmel and R. F. Christman. Life Sciences Research Reports, 41, 215–43.Google Scholar
Hudec, M. R. and Jackson, M. P. A. (2006). Advance of allochthonous salt sheets in passive margins and orogens. American Association of Petroleum Geologists Bulletin, 90, 1535–64.CrossRefGoogle Scholar
Hughes, W. B., Holba, A. G., and Dzou, L. I. (1995). The ratios of dibnzothiophene to phenathrene and pristine to phytane as indicators of depositional environment and lithology of petroleum source rocks. Geochimica et Cosmochimica Acta, 59, 3581–98.CrossRefGoogle Scholar
Huismans, R. S. (1998). Dynamic Modelling of the Transition from Passive to Active Rifting. Vrije University, Amsterdam, p. 182.Google Scholar
Huismans, R. S. (1999). Dynamic modeling of the transition from passive to active rifting. Application to the Pannonian basin. PhD thesis, Institute of Earth Sciences, Vrije Universiteit Amsterdam, p. 196.Google Scholar
Huismans, R. S. and Beaumont, C. (2005). Effect of lithospheric stratification on extensional styles and rift basin geometry. In Petroleum Systems of Divergent Continental Margin Basins, ed. Post, P. J., Rosen, N. C., Olson, D. L., Palmes, S. L., Lyons, K. T., and Newton, G. B.. 25th Annual GCSSEPM Foundation Bob F. Perkins Research Conference, GCSSEPM, Houston, 25, 1255.CrossRefGoogle Scholar
Huismans, R. S. and Beaumont, C. (2007). Roles of lithospheric strain softening and heterogeneity in determining the geometry of rifts and continental margins. Geological Society of London Special Publications, 282, 111–38.CrossRefGoogle Scholar
Huismans, R. S. and Beaumont, C. (2008). Complex rifted continental margins explained by dynamical models of depth-dependent lithospheric extension. Geology, 36, 163–6.CrossRefGoogle Scholar
Huismans, R. S. and Beaumont, C. (2011). Depth-dependent extension, two-stage breakup and cratonic underplating at rifted margins. Nature, 473, 74–9.CrossRefGoogle ScholarPubMed
Huismans, R. S., Buiter, S. J. H., and Beaumont, C. (2005). Effect of plastic-viscous layering and strain softening on mode selection during lithospheric extension. Journal of Geophysical Research, 110, 17.CrossRefGoogle Scholar
Huismans, R. S., Podladchikov, Y. Y., and Cloetingh, S. A. P. L. (2001). Transition from passive to active rifting: relative importance of asthenospheric doming and passive extension of the lithosphere. Journal of Geophysical Research, 106, 11271–91.CrossRefGoogle Scholar
Hulen, J. B., Goff, F., Ross, J. R., Bortz, L. C., and Bereskin, S. R. (1994). Geology and geothermal origin of Grant Canyon and Bacon Flat oil fields, Railroad Valley, Nevada. American Association of Petroleum Geologists Bulletin, 78, 596623.Google Scholar
Hulen, J. B., Norton, D., Kaspereit, D., Murray, L., van de Putte, T., and Wright, M. (2003). Geology and working conceptual model of the Obsidian Butte (Unit 6) sector of the Salton Sea geothermal field, California. Geothermal Resources Council Transactions, 27, 227–40.Google Scholar
Hulen, J. B. and Pulka, F. S. (2001). Newly-discovered, ancient extrusive rhyolite in the Salton Sea geothermal field, Imperial Valley, California. Twenty-Sixth Workshop on Geothermal Reservoir Engineering, Stanford, California, January 29–31, 2001, pp. 1–8.Google Scholar
Hunstad, I., Selvaggi, G., D’Agostino, N., England, P., Clarke, P., and Pierozzi, M. (2003). Geodetic strain in peninsular Italy between 1875 and 2001. Geophysical Research Letters, 30, 1181.CrossRefGoogle Scholar
Hunt, C. B. and Mabey, D. R. (1966). General geology of Death Valley, California: stratigraphy and structure. U.S. Geological Survey Professional Paper, 0494-A.Google Scholar
Hunt, J. M. (1979). Petroleum Geochemistry and Geology. W. H. Freeman and Company, San Francisco, California.Google Scholar
Hunt, J. M. (1990). Generation and migration of petroleum from abnormally pressured fluid compartments. American Association of Petroleum Geologists Bulletin, 74, 112.Google Scholar
Hunt, J. M. (1991). Generation of gas and oil from coal and other terrestrial organic matter. Organic Geochemistry, 17, 673–80.CrossRefGoogle Scholar
Huntoon, P. W. (1982). The Meander Anticline, Canyonlands, Utah: an unloading structure resulting from horizontal gliding on salt. Geological Society of America Bulletin, 93, 941–50.2.0.CO;2>CrossRefGoogle Scholar
Hurford, A. J. and Green, P. F. (1983). The zeta age calibration of fission-track dating. Chemical Geology, 41, 285317.CrossRefGoogle Scholar
Hurlbut, C. S. and Klein, C. (1977). Manual of mineralogy (after James D. Dana). John Wiley and Sons, New York, p. 532.Google Scholar
Hurlow, H. A., Snoke, A. W., and Hodges, K. V. (1991). Temperature and pressure of mylonitization in a Tertiary extensional shear zone, Ruby Mountains–East Humboldt Range, Nevada: tectonic implications. Geology, 19, 82–6.2.3.CO;2>CrossRefGoogle Scholar
Husmo, T., Hamar, G., Hoiland, O., Johannessen, E. P., Romuld, A., Spencer, A., and Titterton, R. (2002). Chapter 10. Lower and Middle Jurassic. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. D. Evans, C. Graham, A. Armour, and P. Bathurst. Geological Society of London, pp. 129–55.Google Scholar
Hutchinson, I. (1985). The effects of sedimentation and compaction on oceanic heat flow. Geophysical Journal of the Royal Astronomic Society, 82, 439–59.Google Scholar
Hutton, A. C. and Cook, A. C. (1980). Influence of alginite on the reflectance of vitrinite from Joadia, NSW, and some other coals and oils shales containing alginite. Fuel, 59, 711–14.CrossRefGoogle Scholar
Hutton, A. C., Kantsler, A. J., Cook, A. C., and McKirdy, D. M. (1980). Organic matter in oil shales. The APEA Journal, 20, 4467.Google Scholar
Huuse, M. (2002). Cenozoic uplift and denudation of southern Norway: insights from the North Sea Basin. In Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration, ed. Dore, A. G., Cartwright, J. A., Stoker, M. S., Turner, J. P., and White, N.. Geological Society of London Special Publications, 196, 209–33.Google Scholar
Huuse, M., Duranti, D., Steinsland, N., Guargena, C. G., Prat, P., Holm, K., Cartwright, J. A., and Hurst, A. (2004). Seismic characteristics of large-scale sandstone intrusions in the Paleogene of the south Viking Graben, UK and Norwegian North Sea. In 3D Seismic Data: Application to the Exploration of Sedimentary Basins, ed. Davies, R. J., Cartwright, J. A., Stewart, S. A., Underhill, J. R., and Lappin, M.. Geological Society of London Memoir, 29, 257–71.Google Scholar
Ibrahim, A. E., Ebinger, C. J., and Fairhead, J. D. (1991). Interpretation of the Central African Rift system of Sudan based on new gravity and aeromagnetic data. EOS, 72, 462.Google Scholar
Ibrmajer, J., Suk, M., and 35 others (1989). Geophysical Picture of Czechoslovak Socialistic Republic (in Czech). ÚÚG Praha, p. 354.Google Scholar
IES (1993) In Poelchau, H. S., Baker, D. R., Hantschel, T., Horsfield, B. and Wygrala, B. (1997). Basin simulation and the design of the conceptual basin model. In Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling, ed. Welte, D. H., Horsfield, B., and Baker, D. R.. Springer, Berlin, pp. 370.Google Scholar
Illies, J. H. (1981). Mechanism of graben formation. Tectonophysics, 73, 249–66.CrossRefGoogle Scholar
Ingebritsen, S. E. and Manning, C. E. (1999). Geological implications of a permeability-depth curve for the continental crust. Geology, 27, 1107–10.2.3.CO;2>CrossRefGoogle Scholar
International Heat Flow Commission (1991). A New Global Heat Flow Compilation Assembled by Pollack, H. N., Hurter, S. J. and Johnson, J. R. Department of Geological Sciences, University of Michigan, Ann Arbor.Google Scholar
Inyang, M. I., Ekweozor, C. M., and Pratt, L. M. (1995). Mid-Cretaceous anoxic events in southeastern Nigeria sedimentary basins – geochemical signatures and petroleum potential implications. Nigerian Association of Petroleum Explorationists, Official Programme, p. 34.Google Scholar
ION/GX Technology 2007. IndiaSpan East, Phase 1, Interpretation Report. Archive of Reliance, Mumbai, 39.Google Scholar
Irwin, W. P. and Barnes, I. (1975). Effect of geologic structure and metamorphic fluids on seismic behavior of the San Andreas fault system in Central and Northern California. Geology, 3, 713–16.2.0.CO;2>CrossRefGoogle Scholar
Irwin, W. P. and Barnes, I. (1980). Tectonic relations of carbon dioxide discharges and earthquakes. Journal of Geophysical Research, 85, 3115–21.CrossRefGoogle Scholar
Isaacs, C. M. and Garrison, R. E. (1983). Petroleum Generation and Occurrence in the Miocene Monterey Formation, California. Guidebook, Los Angeles, SEPM, Pacific Section.Google Scholar
Issler, D. H., McQueen, H., and Beaumont, C. (1989). Thermal and isostatic consequences of simple shear extension of the continental lithosphere. Earth and Planetary Science Letters, 91, 341–58.CrossRefGoogle Scholar
Issler, D. R. and Beaumont, C. (1989). A finite-element model of the subsidence and thermal evolution of extensional basins: application to the Labrador continental margin. In Thermal History of Sedimentary Basins, Methods and Case Histories, ed. Naeser, N. D. and McCulloch, T. H.. Springer-Verlag, Berlin, pp. 239–67.Google Scholar
Ito, G., Lin, J., and Gable, C. W. (1996). Dynamics of mantle flow and melting at a ridge-centered hotspot: Iceland and the Mid-Atlantic Ridge. Earth and Planetary Science Letters, 144, 5374.CrossRefGoogle Scholar
Ito, G., Lin, J., and Gable, C. (1997). Interaction of mantle plumes and migrating mid-ocean ridge systems: implications for the Galapagos plume-ridge system. Journal of Geophysical Research, 102, 15403–17.CrossRefGoogle Scholar
Ivanov, A. V., Boven, A. A., Brandt, I. S., and Rasskazov, S. V. (2002). Achievements and Limitations of the K-Ar and 40Ar/39AR Methods: What’s in it for dating the Quaternary Sedimentary Deposits? International Symposium – Speciation of Ancient Lakes, Berliner Palaobiologische Abhandlungen, Irkutsk, pp. 65–75.Google Scholar
Izundu, U. (2007). Petrobras to start Tupi oil development by 2011. Oil and Gas Journal, p. 2.Google Scholar
Jackson, J. (1994). Active tectonics of the Aegean region. Annual Review of Earth and Planetary Sciences, 22, 239–71.CrossRefGoogle Scholar
Jackson, J. A., Gagnepain, J., Houseman, G., King, G. C. P., Papadimitriou, P., Soufleris, C., and Virieux, J. (1982). Seismicity, normal faulting, and the geomorphological development of the Gulf of Corinth (Greece): the Corinth earthquakes of February and March 1981. Earth and Planetary Science Letters, 57, 377–97.CrossRefGoogle Scholar
Jackson, J. A. and Leeder, M. (1994). Drainage systems and the development of normal faults: an example from Pleasant Valley, Nevada. Journal of Structural Geology, 16, 1041–59.CrossRefGoogle Scholar
Jackson, J. A. and McKenzie, D. (1983). The geometrical evolution of normal fault systems. Journal of Structural Geology, 5, 471–82.CrossRefGoogle Scholar
Jackson, J. A., White, N. J., Garfunkel, Z., and Anderson, H. (1988). Relations between normal-fault geometry, tilting and vertical motions in extensional terrains: an example from the southern Gulf of Suez. Journal of Structural Geology, 10, 155–70.CrossRefGoogle Scholar
Jackson, M. P. A., Cramez, C., and Fonck, J.-M. (2000). Role of subaerial volcanic rocks and mantle plumes in creation of South Atlantic margins: implications for salt tectonics and source rocks. Marine and Petroleum Geology, 17, 477–98.CrossRefGoogle Scholar
Jaeger, J. C. (1960). Shear failure of anisotropic rocks. Geological Magazine, 97, 6572.CrossRefGoogle Scholar
James, E. W. and Silver, L. T. (1988). Implications of zeolites and their zonation in the Cajon Pass deep drillhole. Geophysical Research Letters, 15, 973–6.CrossRefGoogle Scholar
Janecky, D. R. and Seyfried, W. E. Jr. (1986). Hydrothermal serpentinization of peridotite within the oceanic crust: experimental investigations of mineralogy and major element chemistry. Geochimica et Cosmochimica Acta, 50, 1357–78.CrossRefGoogle Scholar
Jansa, L. F. and Pe-Piper, G. (1985). Early Cretaceous volcanism on the northeastern American margin and implications for plate tectonics. GSA Bulletin, 96, 8391.2.0.CO;2>CrossRefGoogle Scholar
Jardim de Sá, E. F. (1984). A evolução Proterozoica da Província Borborema. Atas do 11th simposio de geologia do nordeste, Natal, Brazil, pp. 297–315.Google Scholar
Jarrige, J.-J., Ott d´Estevou, P., Burollet, P. F., Montenat, C., Prat, P., Richert, J.-P., and Thiriet, J. P. (1990). The multistage tectonic evolution of the Gulf of Suez and northern Red Sea continental rift from field observations. Tectonics, 9, 441–65.CrossRefGoogle Scholar
Jeffrey, A. W. A., Alimi, H. M., and Jenden, P. D. (1991). Geochemistry of Los Angeles basin oil and gas systems. In Active Margin Basins, ed. Biddle, K. T.. American Association of Petroleum Geologists Memoir, 52, 197–219.Google Scholar
Jensenius, J. and Munksgaard, N. C. (1989). Large scale hot water migration systems around salt diapirs in the Danish central trough and their impact on diagenesis of chalk reservoirs. Geochimica et Cosmochimica Acta, 53, 7988.CrossRefGoogle Scholar
Jenssen, A. I., Bergslien, D., Rye-Larsen, M., and Lindholm, R. M. (1993). Origin of complex mound geometry of Palaeocene submarine-fan sandstone reservoirs, Balder field, Norway. In Petroleum Geology of Northwest Europe, ed. J. R. Parker. Proceedings of the 4th Conference, Geological Society of London, pp. 135–43.Google Scholar
Jeremiah, J. M. and Nicholson, P. H. (1999). Middle Oxfordian to Volgian sequence stratigraphy of the Greater Shearwater area. In Petroleum Geology of Northwest Europe, ed. A. J. Fleet and S. A. R. Boldy. Proceedings of the 5th Conference, The Geological Society of London, pp. 153–70.Google Scholar
Jerram, D. A., Mountney, N., Holzforster, F., and Stollhofen, H. (1999). Internal stratigraphic relationships in the Etendeka Group in the Huab Basin, NW Namibia: understanding the onset of flood volcanism. Journal of Geodynamics, 28, 393418.CrossRefGoogle Scholar
Jessop, A. M. and Majorowicz, J. A. (1994). Fluid flow and heat transfer in sedimentary basins. In Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins, ed. Parnell, J.. Geological Society of London Special Publications, 78, 4354.Google Scholar
Jin, Z., Cao, J., Hu, W., Zhang, Y., Yao, S., Wang, X., Zhang, Y., Tang, Y., and Shi, X. (2008). Episodic petroleum fluid migration in fault zones of the northwestern Junggar Basin (northwest China): evidence from hydrocarbon-bearing zoned calcite cement. American Association of Petroleum Geologists Bulletin, 92, 1225–43.CrossRefGoogle Scholar
Jiracek, G. R. (1995). Geoelectromagnetics charges on. U.S. Quadrennial Report to IUGG. Reviews on Geophysics, 33, 169–76.Google Scholar
Johannessen, P. N. and Andsbjerg, J. (1993). Middle to Late Jurassic basin evolution and sandstone distribution in the Danish Central Trough. In Petroleum Geology of Northwest Europe, ed. J. R. Parker. Proceedings of the 4th Conference, Geological Society of London, pp. 271–83.CrossRefGoogle Scholar
Johannessen, P. N., Dybkjaer, K., and Rasmussen, E. S. (1996). Sequence stratigraphy of Upper Jurassic reservoir sandstones in the northern part of the Danish Central Trough, North Sea. Marine and Petroleum Geology, 13, 755–70.CrossRefGoogle Scholar
Johannssen, E. P., Mjøs, R., Renshaw, D., Dalland, A., and Jacobsen, T. (1995). Northern limit of the “Brent delta” at the Tampen Spur – a sequence stratigraphic approach for sandstone prediction. In Sequence Stratigraphy on the Northwest European Margin, ed. Steel, R. J., Felt, V. L., Johannssen, E. P., and Mathieu, C.. Special Publication of the Norwegian Petroleum Society, Amsterdam, Elsevier, 5, 213–56.Google Scholar
John, B. E. (1984). Primary corrugations in Tertiary low-angle normal faults, SE California: porpoising mullion structures? Geological Society of America Abstracts with Programs, 16, 291.Google Scholar
Johnson, A. and Eyssautier, M. (1987). Alwyn North Field and its regional geological context. In Petroleum Geology of North West Europe, ed. Brooks, J. and Glennie, K. W.. Graham and Trotman, London, pp. 963–77.Google Scholar
Johnson, H. D. and Fisher, M. J. (1998). North Sea plays: geological controls on hydrocarbon distribution. In Petroleum Geology of the North Sea, Basic Concepts and Recent Advances (fourth edition), ed. Glennie, K. W.. Blackwell Scientific Publications, Oxford, pp. 463547.CrossRefGoogle Scholar
Johnson, H. D., Mackay, T. A., and Stewart, D. J. (1986). The Fulmar oil field (Central north Sea): geological aspects of its discovery, appraisal and development. Marine and Petroleum Geology, 3, 99125.CrossRefGoogle Scholar
Johnson, N. M., Officer, C. B., Opdyke, N. D., Woodard, G. D., Zeitler, P. K, and Lindsay, H. (1983). Rates of late Cenozoic tectonism in the Vallecito-Fish Creek basin, western Imperial Valley, California. Geology, 11, 664–7.2.0.CO;2>CrossRefGoogle Scholar
Joint Chalk Research (1996). Geology, rock mechanics, rock properties and improved oil recovery in chalk of the Danish and Norwegian sectors of the North Sea. Joint Chalk Research, Phase IV Monograph.Google Scholar
Jolley, D. W. (1997). Palaeosurface palynofloras of the Skye lava field and the age of British Tertiary volcanic province. In Palaesurfaces: Recognition, Reconstruction and Palaeo-environmental Interpretation, ed. M. Widdowson. Geological Society [London] Special Publication 120, pp. 6794.Google Scholar
Jolly, R. J. H., Cosgrove, J. W., and Dewhurst, D. N. (1998). Thickness and spatial distributions of clastic dykes, northwest Sacramento Valley, California. Journal of Structural Geology, 20, 1663–72.CrossRefGoogle Scholar
Jolly, R. J. H. and Lonergan, L. (2002). Mechanisms and controls on the formation of sand intrusions. Journal of Geological Society, 159, 605–17.CrossRefGoogle Scholar
Jones, A. G. (1992). Electrical conductivity of the continental lower crust. In Continental Lower Crust, ed. Fountain, D. M., Arculus, R. J., and Kay, R. W.. Elsevier, Amsterdam, pp. 81143.Google Scholar
Jones, C. H., Unruh, J. R., and Sonder, L. J. (1996). The role of gravitational potential energy in active deformation in the Southwestern United States. Nature, 381, 3741.CrossRefGoogle Scholar
Jones, E., Jones, R., Ebdon, C., Ewen, D., Milner, P., Plunkert, J., Hudson, G., and Slater, P. (2003). Eocene. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. The Geological Society of London, pp. 261–77.Google Scholar
Jones, L. S., Garrett, S. W., Macleod, M., Guy, M., Condon, P. J., and Notman, L. (1999). Britannia field, UK Central North Sea: modeling heterogeneity in unusual deep-water deposits. In Petroleum Geology of Northwest Europe. Proceedings of the 5th Conference, ed. A. J. Fleet and S. A. R. Boldy. The Geological Society of London, pp. 1115–24.CrossRefGoogle Scholar
Jones, S. M., White, N., Clarke, B. J., Rowley, E., and Gallagher, K. (2002). Present and past influence of the Iceland Plume on sedimentation. In Exhumation of the North Atlantic Margin: Timing, Mechanisms, and Implications for Petroleum Exploration, ed. Dore, A. G., Cartwright, J. A., Stoker, M. S., Turner, J. P., and White, N.. Geological Society of London Special Publications, pp. 1325.Google Scholar
Jones, S. M., White, N., and Lovell, B. (2001). Cenozoic and Cretaceous transient uplift in the Porcupine Basin and its relationship to a mantle plume. In The Petroleum Exploration of Ireland’s Offshore Basins, ed. P. M. Shannon, P. D. W. Haughton, and D. Corcoran. Geological Society of London Special Publications, 188, 345–60.Google Scholar
Jongmans, D. and Malin, P. E. (1995). Microearthquake S-wave observations from 0 to 1 km in the Varian well at Parkfield, California. Bulletin of the Seismological Society of America, 85, 1805–21.CrossRefGoogle Scholar
Jonk, R., Duranti, D., Parnell, J., Hurst, A., and Fallick, A. E. (2003). The structural and diagenetic evolution of injected sandstones: examples from the Kimmeridgian of NE Scotland. Journal of the Geological Society of London, 160, 881–94.CrossRefGoogle Scholar
Jonk, R., Hurst, A., Duranti, D., Parnell, J., Mazzini, A., and Fallick, A. E. (2005). Origin and timing of sand injection, petroleum migration, and diagenesis in Tertiary reservoirs, south Viking Graben, North Sea. American Association of Petroleum Geologists Bulletin, 89, 329–57.CrossRefGoogle Scholar
Joppen, M. and White, R. S. (1990). The structure and subsidence of Rockall Trough from two-ship seismic experiments. Journal of Geophysical Research, 95, 19821–37.CrossRefGoogle Scholar
Jordan, T. E. and Flemings, P. B. (1991). Large-scale stratigraphic architecture, eustatic variation, and unsteady tectonism: a theoretical evaluation. Journal of Geophysical Research, 96, 6681–99.CrossRefGoogle Scholar
Jordan, T. H. (1978). Composition and development of the continental tectosphere. Nature, 274, 544–8.CrossRefGoogle Scholar
Jowett, E. C., Cathles, L. M., and Davis, B. W. (1993). Predicting depths of gypsum dehydration in evaporitic sedimentary basins. American Association of Petroleum Geologists Bulletin, 77, 402–13.Google Scholar
Juhlin, C. (1990). Interpretation of the reflections in the Siljan Ring area based on results from the Gravberg-1 borehole. Tectonophysics, 173, 345–60.CrossRefGoogle Scholar
Jurine, D., Jaupart, C., Brandeis, G., and Tackley, P. J. (2005). Penetration of mantle plumes through depleted lithosphere. Journal of Geophysical Research, 110, B10.CrossRefGoogle Scholar
Kadolsky, D., Johansen, S. J., and Duxbury, S. (1999). Sequence stratigraphy and sedimentary history of the Humber Group (Late Jurassic – Ryazanian) in the Outer Moray Firth (UKCS, North Sea). In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, Geological Society of London, pp. 839–60.Google Scholar
Kaila, K. L., Krishna, V. G., and Mall, D. M. (1981). Crustal structure along Mehmadabad-Billimora profile in Cambay basin, India from deep seismic soundings. Tectonophysics, 76, 99130.CrossRefGoogle Scholar
Kaila, K. L. and Sain, K. (1997). Variation of crustal velocity structure in India as determined from DSS studies and their implications as regional tectonics. Journal of the Geological Society of India, 49, 395407.Google Scholar
Kaila, K. L., Tewari, H. C., Krishna, V. G., Dixit, M. M., Sarkar, D., and Reddy, M. S. (1990). Deep Seismic Sounding studies in the north Cambay and Sanchor basins, India. Geophysical Journal International, 103, 621–37.CrossRefGoogle Scholar
Kappelmeyer, O. and Haenel, R. (1974). Geothermics with Special Reference to Application. Gebrueder Borntraeger, Berlin.Google Scholar
Karato, S.-I. and Wu, P. (1993). Rheology of the upper mantle: a synthesis. Science, 260, 771–8.CrossRefGoogle ScholarPubMed
Karig, D. E. (1986). Physical properties and mechanical state of accreted sediments in the Nankai Trough, Southwest Japan Arc. In Structural Fabric in Deep Sea Drilling Project Cores from Forearcs, ed. Moore, J. C.. Geological Society of America Memoir, p. 117–33.Google Scholar
Karim, A., Pe-Piper, G., Piper, D. J. W., and Hanley, J. J. (2010). Thermal history and the relationship between diagenesis and the reservoir connectivity: Venture field, offshore Nova Scotia, eastern Canada. Open File Report – Geological Survey of Canada, 6557, 1256.Google Scholar
Karl, D. M., McMurtry, G. M., Malahoff, A., and Garcia, M. O. (1988). Loihi Seamount, Hawaii: a mid-plate volcano with a distinctive hydrothermal system. Nature, 335, 532–5.CrossRefGoogle Scholar
Karlo, J. F. and Shoup, R. C. (2000). Classification of syndepositional: structural systems, northern Gulf of Mexico. The Bulletin of the Houston Geological Society, 42, 1112.Google Scholar
Karner, G. D. (2000). Rifts of the Campos and Santos basins, southeastern Brazil: distribution and timing. American Association of Petroleum Geologists Memoir, 73, 301–15.Google Scholar
Karner, G. D. (2008). Depth-dependent extension and mantle exhumation: an extreme passive margin endmember or a new paradigm? Central Atlantic Conjugate Margins Conference, August 13–15, 2008, Halifax, Program & Extended Abstracts, 10–16. http://www.conjugatemargins.com/abstracts/by_presenter/jp, accessed on January 27, 2012.Google Scholar
Karner, G. D. and Dewey, J. F. (1986). Rifting: lithospheric versus crustal extension as applied to the Ridge Basin of southern California. American Association of Petroleum Geologists Memoir, A131, 317–37.Google Scholar
Karner, G. D. and Driscoll, N. W. (1999). Style, timing and distribution of tectonic deformation across the Exmouth Plateau, northwest Australia, determined from stratal architecture and quantitative basin modelling. Geological Society of London Special Publications, 164, 271311.CrossRefGoogle Scholar
Karner, G. D. and Driscoll, N. W. (1999). Tectonic and stratigraphic development of the West African and eastern Brazilian margins: insights from quantitative basin modelling. Geological Society of London Special Publications, 153, 1140.CrossRefGoogle Scholar
Karner, G. D., Driscoll, N. W., and Barker, D. H. N. (2003). Syn-rift region subsidence across the West African continental margin: the role of lower plate ductile extension. In Petroleum Geology of Africa: New Themes and Developing Technologies, ed. Arthur, T. J., Macgregor, D. S., and Cameron, N.. Geological Society of London Special Publications, 207, 105–29.Google Scholar
Karner, G. D. and Gamboa, L. A. P. (2007). Timing and origin of the South Atlantic pre-salt basins and their capping evaporites. Geological Society of London Special Publications, 285, 1535.CrossRefGoogle Scholar
Karner, G. D., Neal, W., McGinnis, J. P., Brumbaugh, W. D., and Cameron, N. R. (1997). Tectonic significance of syn-rift sediment packages across the Gabon-Cabinda continental margin. Marine and Petroleum Geology, 14, 9731000.CrossRefGoogle Scholar
Karner, G. D. and Watts, A. B. (1982). On isostasy at Atlantic-type continental margins. Journal of Geophysical Research, 87, 2923–48.CrossRefGoogle Scholar
Karson, J. A. and Brooks, C. K. (1999). Structural and magmatic segmentation of the Tertiary east Greenland volcanic rifted margin. Geological Society of London Special Publications, 164, 313–18.CrossRefGoogle Scholar
Katili, J. A. and Sudradjat, A. (1984). Galunggung: the 1982–83 eruption. Volcanological Survey of Indonesia, Bandung, p. 102.Google Scholar
Katz, B. J. (1983). Limitations of “Rock-Eval” pyrolysis for typing organic matter. Organic Geochemistry, 4, 195–9.CrossRefGoogle Scholar
Katz, B. J. (1988). Clastic and carbonate lacustrine systems: an organic geochemical comparison (Green River Formation and East African lake sediments). Geological Society of London Special Publications, 40, 8190.CrossRefGoogle Scholar
Katz, B. J. (1990). Control son distribution of lacustrine source rocks through time and space. In Lacustrine Basin Exploration – Case Studies and Modern Analogues, ed. Katz, B. J.. American Association of Petroleum Geologists Memoir, 50, 6176.CrossRefGoogle Scholar
Katz, B. J. (1995a). Petroleum source rocks: an introductory overview. In Petroleum Source Rocks, ed. Katz, B. J.. Springer-Verlag, Berlin, pp. 18.CrossRefGoogle Scholar
Katz, B. J. (1995b). A survey of rift basin source rocks. In Hydrocarbon Habitat in Rift Basins, ed. Lambiase, J. J.. Geological Society of London Special Publications, 80, 213–42.Google Scholar
Katz, B. J., Kelley, P. A., Royle, R. A., and Jorjorian, T. (1991). Hydrocarbon products of coals as revealed by pyrolysis-gas chromatography. Organic Geochemistry, 17, 711–22.CrossRefGoogle Scholar
Katz, B. J. and Mello, M. R. (2000). Petroleum systems of South Atlantic marginal basins – an overview. In Petroleum Systems of South Atlantic Margins, ed. Mello, M. R. and Katz, B. J.. American Association of Petroleum Geologists Memoir, 73, 113.Google Scholar
Kaus, B. J. P., Connolly, J. A. D., Podladchikov, Y. Y., and Schmalholz, S. M. (2005). Effect of mineral phase transitions on sedimentary basin subsidence and uplift. Earth and Planetary Science Letters, 233, 213–28.CrossRefGoogle Scholar
Kawka, O. E. and Simoneit, B. R. T. (1987). Survey of hydrothermally-generated petroleums from the Guaymas Basin spreading center. Organic Geochemistry, 11, 311–28.CrossRefGoogle Scholar
Keaton, J. R., Currey, D. R., and Olig, S. S. (1993). Paleoseismicity and earthquake hazards evaluation of the West Valley fault zone, Salt Lake City urban area, Utah. Contract Report, 93–8, 155.Google Scholar
Keefer, D. K. (1984). Landslides caused by earthquakes: source characteristics. Bulletin of the Geological Society of America, 95, 406–21.2.0.CO;2>CrossRefGoogle Scholar
Keen, C. E. (1985a). The dynamics of rifting: deformation of the lithosphere by active and passive driving forces. Geophysical Journal of the Royal Astronomical Society, 80, 95120.CrossRefGoogle Scholar
Keen, C. E. (1985b). Evolution of rifted continental margins. Paper – Geological Survey of Canada, 85, 810.Google Scholar
Keen, C. E. and Boutilier, R. R. (1995). Lithosphere- asthenosphere interaction below rifts. Rifted Ocean-Continent Boundaries, 463, 1730.Google Scholar
Keen, C. E., Boutilier, R., de Voogd, B., Mudford, B., and Enachescu, M. E. (1987). Crustal geometry and extensional models for the Grand Banks, Eastern Canada: constraints from deep seismic reflection data. Atlantic Geoscience Society Special Publication, 5, 101–15.Google Scholar
Keen, C. E., Kay, W. A., and Roest, W. R. (1990). Crustal anatomy of a transform continental margin. Tectonophysics, 173, 527–44.CrossRefGoogle Scholar
Keen, C. E., Kay, W. A., Keppie, J. D., Marillier, F., Pe-Piper, G., and Waldron, J. W. F. (1991). Deep seismic reflection data from the Bay of Fundy and Gulf of Maine: tectonic implications for the Northern Appalachians. Canadian Journal of Earth Sciences, 28, 1096–111.CrossRefGoogle Scholar
Keen, C. E. and Potter, D. P. (1995). Formation and evolution of the Nova Scotia rifted margin: evidence from deep seismic reflection data. Tectonics, 14, 918–32.Google Scholar
Keith, J. D. and Shanks, W. C. (1988). Chemical evolution and volatile fugacities of the Pine Grove porphyry molybdenum and ash-flow tuff system, southwestern Utah. In Recent Advances in the Geology of Granite-Related Mineral Deposits, ed. Taylor, R. P. and Strong, D. F.. Canadian Institute of Mining and Metallurgy Special Volume, 39, 402–23.Google Scholar
Keith, J. F., Nemčok, M., Fleischmann, K. H., Nemčok, J., Gross, P., Rudinec, R., and Kanes, W. H. (1991). Sedimentary basins of Slovakia, Part III, Hydrocarbon potential of the Central Carpathian Paleogene Basin. University of South Carolina, Columbia, Earth Sciences and Resources Institute Technical Report, 90-07-306, pp. 1–302.Google Scholar
Kelemen, P. B. and Holbrook, W. S. (1995). Origin of thick, high-velocity igneous crust along the U.S. East Coast Margin. Journal of Geophysical Research, 100, 10077–94.CrossRefGoogle Scholar
Keller, G. R., Braile, L. W., and Morgan, P. (1979). Crustal structure, geophysical models and contemporary tectonism of the Colorado Plateau. Tectonophysics, 61, 131–47.CrossRefGoogle Scholar
Kelley, P. A., Mertani, B., and Williams, H. H. (1995). Brown Shale Formation: Paleocene lacustrine source rocks of central Sumatra. In Petroleum Source Rocks, ed. Katz, B. J.. Springer-Verlag, Berlin, pp. 283308.CrossRefGoogle Scholar
Kendall, C. (2008). Sequence stratigraphy provides a basic framework to conceptual models used to interpret depositional systems: the key to simplification of the complex terminology of sequence stratigraphy is to use simple depositional models. International Congress, Abstracts – Congres Geologique International, Resumes, 33, 1255505.Google Scholar
Kendall, C. and Haughton, P. (2008). Sequence stratigraphy of deepwater fans and mass transport debris. In SEPM Sequence Stratigraphy Web. http://strata.geol.sc.edu/Deepwater/DeepwaterCLasticSeqStrat.ht.Google Scholar
Kendall, J. M., Stuart, G. W., Ebinger, C. J., Bastow, I. D., and Keir, D. (2005). Magma-assisted rifting in Ethiopia. Nature, 433, 146–8.CrossRefGoogle ScholarPubMed
Kennedy, B. M., Kharaka, Y. K., Evans, W. C., Ellwood, A., DePaolo, D. J., Thordsen, J., Ambats, G., and Mariner, R. H. (1997). Mantle fluids in the San Andreas fault system, California. Science, 278, 1278–81.CrossRefGoogle Scholar
Kent, R. W., Saunders, A. D., Kempton, P. D., and Ghose, N. C. (1997). Rajmahal basalts, eastern India: mantle sources and melt distribution at a volcanic rifted margin. Large Igneous Provinces: Continental Oceanic and Planetary Flood Volcanism, ed. J. Mahoney and M. F. Coffin. American Geophysical Union Geophysical Monograph 100, 145–82.Google Scholar
Keppie, J. D. (1982). Tectonic map of Nova Scotia. Abstracts with Programs – Geological Society of America, 14, 30.Google Scholar
Kerr, A. C. (1994). Lithospheric thinning during the evolution of continental large igneous provinces: a case study from the North Atlantic Tertiary Province. Geology, 22, 1027–30.2.3.CO;2>CrossRefGoogle Scholar
Kerr, D. R., Pappajohn, S., and Bell, P. J. (1979). Neogene continental rifting and sedimentation in the western Salton Trough, California. Abstracts with Programs – Geological Society of America, 11, 457.Google Scholar
Khalil, S. M. (1998). Tectonic and structural development of the eastern Gulf of Suez margin. PhD thesis, Royal Holloway University of London, p. 349.Google Scholar
Kharaka, Y. K., Ambats, G., Evans, W. C., and White, A. F. (1988b). Geochemistry of water at Cajon Pass, California: preliminary results. Geophysical Research Letters, 15, 1037–40.Google Scholar
Kharaka, Y. K., Thordsen, J. J., Evans, W. C., and Kennedy, B. M. (1999). Geochemistry and hydromechanical interactions of fluids associated with the San Andreas fault system, California. In Faults and Subsurface Fluid Flow in the Shallow Crust, ed. Haneberg, W. C., Mozley, P. S., Moore, J. C., and Goodwin, L. B.. American Geophysical Union, pp. 129–48.Google Scholar
Kharaka, Y. K., White, L. D., Ambats, G., and White, A. F. (1988a). Origin of subsurface water at Cajon Pass, California. Geophysical Research Letters, 15, 1049–52.Google Scholar
Kim, E. M. (2002). Natural gas resource characterization of the Federal Offshore Gulf of Mexico: data analysis of key sand-reservoir attributes for future exploration and development opportunities. Gulf Coast Association of Geological Societies and Gulf Coast Transactions, 52, 527–36.Google Scholar
Kincaid, C., Sparks, D. W., and Detrick, R. S. (1996). The relative importance of plate-driven and buoyancy-driven flow at mid-ocean ridges. Journal of Geophysical Research, 101, 16177–93.CrossRefGoogle Scholar
King, S. D. (2005). Archean cratons and mantle dynamics. Earth Planetary Science Letters, 234, 114.CrossRefGoogle Scholar
King, S. D. and Anderson, D. L. (1995). An alternative mechanism of flood basalt formation. Earth and Planetary Science Letters, 136, 269–79.CrossRefGoogle Scholar
King, S. D. and Anderson, D. L. (1998). Edge-driven convection. Earth and Planetary Science Letters, 160, 289–96.CrossRefGoogle Scholar
Kincaid, C., Ito, G., and Gable, C. (1995a). Laboratory investigation of the interaction of off-axis mantle plumes and spreading centres. Nature, 376, 758–61.Google Scholar
Kincaid, C., Schilling, J.-G., and Gable, C. (1995b). The dynamics of off-axis plume-ridge interaction in the uppermost mantle. Earth and Planetary Science Letters, 137, 2943.CrossRefGoogle Scholar
Kirby, S. H. (1983). Rheology of the lithosphere. Reviews of Geophysics, 21, 1458–87.CrossRefGoogle Scholar
Kirby, S. H. (1985). Rock mechanics observations pertinent to the rheology of the continental lithosphere and the localization of strain along shear zones. Tectonophysics, 119, 127.CrossRefGoogle Scholar
Kirby, S. H. and Kronenberg, A. K. (1987). Rheology of the lithosphere: selected topics. Reviews of Geophysics, 25, 1219–44.Google Scholar
Kirby, S. H. and McCormick, J. W. (1984). Inelastic properties of rocks and minerals; strength and rheology. In Handbook of Physical Properties of Rocks, ed. Carmichael, R. S.. CRC Press, Boca Raton, 3, 140280.Google Scholar
Kiss, B. and Toth, J. (1985). Well log interpretation of metamorphic hydrocarbon-bearing formations. First Break, 3, 2431.CrossRefGoogle Scholar
Kissin, Y. V. (1987). Catagenesis and composition of petroleum: origin of n-alkanes and isoalkanes in petroleum crudes. Geochimica et Cosmochimica Acta, 51, 2445–57.CrossRefGoogle Scholar
Kissling, E., Deichmann, N., Husen, S., and Zappone, A. (2006). Lower crustal seismic velocities and seismicity in the northern Alpine foreland: II. Geodynamical interpretation. EGU General Assembly, 2–7 April, 2006, Vienna, p. 562.Google Scholar
Klausen, B. M. and Larsen, H. C. (2002). East Greenland coast-parallel dike swarm and its role in continental breakup. In Volcanic Rift Margins, ed. M. A. Menzies, S. L. Klemperer, C. Ebinger, and J. Baker. Geological Society of America Special Paper, 362, pp. 133–58.Google Scholar
Klemme, H. D. (1994). Petroleum systems in the world that involve Upper Jurassic source rocks. AAPG Memoir, 60, 5172.Google Scholar
Klemme, H. D. and Ulmishek, G. F. (1991). Effective petroleum source rocks of the world: stratigraphic distribution and controlling depositional factors. American Association of Petroleum Geologists Bulletin, 75, 1809–51.Google Scholar
Klemperer, S. and Hobbs, R. (1991). The BIRPS Atlas: Deep Seismic Reflection Profiles around the British Isles. Cambridge University Press.Google Scholar
Klitgord, K. D. and Schouten, H. (1986). Plate kinematics of the central Atlantic. In The Western North Atlantic Region: The Geology of North America, ed. P. R. Vogt and B. E. Tucholke. Geological Society of America, Boulder, pp. 351–78.CrossRefGoogle Scholar
Kluegel, A., Hansteen, T. H., and Schmincke, H. U. (1997). Rates of magma ascent and depths of magma reservoir beneath La Palma, Canary Islands. In International Workshop on Volcanism and Volcanic Hazards in Immature Intraplate Oceanic Islands. Estanción Volcanológica de Canarias, La Palma, Canary Islands, Spain, pp. 29–30.CrossRefGoogle Scholar
Knipe, R. J. (1986). Faulting mechanisms in slope sediments: examples from deep sea drilling project cores. In Structural Fabric in Deep Sea Drilling Project Cores from Forearcs, ed. Moore, J. C.. Geological Society of America Memoir, 166, 4554.CrossRefGoogle Scholar
Knipe, R. J. (1989). Deformation mechanisms: recognition from natural tectonites. Journal of Structural Geology, 11, 127–46.CrossRefGoogle Scholar
Knipe, R. J. (1992). Faulting processes and fault seal. In Structural and Tectonic Modeling and Its Application to Petroleum Geology, ed. Larsen, R. M., Brekke, H., Larsen, B. T., and Talleraas, E.. Proceedings of the Norwegian Petroleum Society Special Publication, 1, 325–42.Google Scholar
Knipe, R. J. (1993). The influence of fault zone processes and diagenesis on fluid flow. In Diagenesis and Basin Development, ed. Horbury, A. D. and Robinson, A.. American Association of Petroleum Geologists Studies in Geology, 36, 135–51.Google Scholar
Knipe, R. J. (1997). Juxtaposition and seal diagrams to help analyze fault seals in hydrocarbon reservoirs. American Association of Petroleum Geologists Bulletin, 81, 187–95.Google Scholar
Knipe, R. J., Agar, S. M., and Prior, D. J. (1991). The microstructural evolution of fluid flow paths in semi-lithified sediments from subduction complex. In The Behaviour and Influence of Fluids in Subduction Zones, ed. Tarney, J., Pickering, K. T., Knipe, R. J., and Dewey, J. F.. Philosophical Transactions of the Royal Society of London, 335, 261–73.Google Scholar
Knox, G. J. and Omatsola, E. M. (1989). Development of the Cenozoic Niger delta in terms of the “Escalator Regression” model and impact on hydrocarbon distribution. Proceedings KNGMG Symposium “Coastal Lowlands, Geology and Geotechnology,” Dordrecht, pp. 181–202.CrossRefGoogle Scholar
Knox, R. W. O. B. (1996). Correlation of the early Paleogene in Northwest Europe: an overview. In Correlation of the Early Paleogene in Northwest Europe, ed. Knox, R. W. O. B., Corfield, R. M., and Dunay, R. E.. Geological Society Special Publications, pp. 111.Google Scholar
Knutson, C. A. and Munro, I. C. (1991). The Beryl Field, block 9/ 13, UK North Sea. 33–42 in United Kingdom Oil and Gas Fields. Memoir of the Geological Society of London, 14, 3342.CrossRefGoogle Scholar
Kohl, T., Bachler, D., and Rybach, L. (2000). Step towards a comprehensive thermo-hydraulic analysis of the HDR test site Soultz-Sous-Forets. Proceedings of the World Geothermal Congress 2000, Kyushu-Tohoku, May 28–June 10, 2671–6.Google Scholar
Kohl, T., Signorelli, S., and Rybach, L. (2001). Three-dimensional (3-D) thermal investigation below high Alpine topography. Physics of the Earth and Planetary Interiors, 126, 195210.CrossRefGoogle Scholar
Koike, L., Kiang, C. H., and Reis, F. A. M. (2007). Organic geochemistry of Santos basin oil, Brazil. Abstracts of Reports – International Congress on Organic Geochemistry, The 23rd International Meeting on Organic Geochemistry, Torquay, UK, 23, 921–2.Google Scholar
Koning, T. and Karsani, A. (1985). Petroleum geology of the Ombilin intermontane basin, West Sumatra. Proceedings of the Annual Convention – Indonesian Oetroleum Association, 14, 117–37.Google Scholar
Kooi, H., Cloetingh, S. A. P. L., and Burrus, J. (1992). Lithospheric necking and regional isostasy at extensional basins 1. Subsidence and gravity modeling with an application to the Gulf of Lions margin (SE France). Journal of Geophysical Research, 97, 17553–71.CrossRefGoogle Scholar
Kopietz, J. and Jung, R. (1978). Geothermal in situ Experiments in the Asse Salt-mine. Seminar on in situ Heating Experiments in Geological Formations. OECD Publications, Paris, France.Google Scholar
Kopylova, M. G., Russell, J. K., and Cookenboo, H. (1999). Petrology of peridotite and pyroxenite xenoliths from Jericho Kimberlite: implications for the thermal state of the mantle beneath the Slave Craton, northern Canada. Journal of Petrology, 40, 79104.CrossRefGoogle Scholar
Korenaga, J., Holbrook, S., Kent, G., Kelemen, P., Detrick, R., Larsen, H.-C., Hopper, J., and Dahl-Jensen, T. (2000). Crustal structure of the southeast Greenland margin from joint refraction and reflection seismic tomography. Journal of Geophysical Research, 105, 21591–614.CrossRefGoogle Scholar
Kotarba, M. J. and Koltun, Y. V. (2006). Origin and habitat of hydrocarbons of the Polish and Ukrainian parts of the Carpathian province. In The Carpathians: Geology and Hydrocarbon Resources, ed. Golonka, J. and Picha, F.. American Association of Petroleum Geologists Memoir, 84, 395443.Google Scholar
Koukovelas, I., Mpresiakas, A., Sokos, E., and Doutsus, T. (1996). The tectonic setting and earthquake ground hazard of the 1993 Pyrogos earthquake, Peloponnese, Greece. Journal of the Geological Society of London, 153, 1137–49.Google Scholar
Kováč, M., Baráth, I., and Nemčok, M. (1992). Bericht 1991 ueber geologische Aufnahmen im Quartaer und Tertiaer im Suedoestlichen Teil des Wiener Beckens auf Blatt 77 Eisenstadt. Jahrbuch der Geologischen Bundensanstalt, 135, 701–3.Google Scholar
Koyi, H., Jenyon, M. K., and Petersen, K. (1993). The effect of basement faulting on diapirism. Journal of Petroleum Geology, 16, 285311.CrossRefGoogle Scholar
Kranz, R. L. (1983). Microcracks in rocks: a review. In Continental Tectonics: Structure, Kinematics and Dynamics, ed. Friedman, M. and Toksoez, M. N.. Tectonophysics, 100, 449–80.Google Scholar
Kranz, R. L. and Scholz, C. H. (1977). Critical dilatant volume of rocks at the onset of tertiary creep. Journal of Geophysical Research, 82, 4893–8.CrossRefGoogle Scholar
Krawczyk, C. M., Reston, T. J., Beslier, M.-O., and Boillot, G. (1996). Evidence for detachment tectonics on the Iberia Abyssal Plain rifted margin. Proceedings of Ocean Drill Program Scientific Results, 149, 603–15.Google Scholar
Krebs, W. N., Wescott, W. A., Nummedal, D., Gaafar, I., Azazi, G., and Karamat, S. (1997). Graphic correlation and sequence stratigraphy of Neogene rocks in the Gulf of Suez. Bulletin de la Societe Geologique de France, 168, 6371.Google Scholar
Kreemer, C., Holt, W. E., and Haines, A. J. (2003). An integrated global model of present-day plate motions and plate boundary deformation. Geophysics Journal International, 154, 834.CrossRefGoogle Scholar
Krishna, K. S., Gopala Rao, D., and Sar, D. (2006). Nature of the crust in the Laxmi Basin (14° – 20°N), western continental margin of India. Tectonics, 25, 198.CrossRefGoogle Scholar
Krooss, B. M., Brothers, L., and Engel, M. H. (1991). Geochromatography in petroleum migration: a review. In Petroleum Migration, ed. England, W. A. and Fleet, A. J.. Geological Society of London Special Publications, 59, 149–63.Google Scholar
Kruijs, E. and Barron, E. (1990). Climatic model prediction of paleoproductivity and potential source-rock distribution. In Deposition of Organic Facies, ed. Huc, A. Y.. American Association of Petroleum Geologists Studies in Geology, 30, 195216.Google Scholar
Krumbein, W. C. and Monk, G. D. (1943). Permeability as a function of the size parameters of unconsolidated sand. Trans. AIME, 151, 153.CrossRefGoogle Scholar
Krus, S. and Šutora, A. (1986). Geophysical-geological Atlas of the Alpine-Carpathian Mountain System. Report, Archív Geofyziky, Brno.Google Scholar
Ksiazkiewicz, M. (1962). Geological Atlas of Poland, Stratigraphic and Facial. Instytut Geologiczny, Warszawa.Google Scholar
Kubala, M., Bastow, M., Thompson, S., Scotchman, I., and Kjell, O. (2002).Chapter 17. Geothermal regime, petroleum generation and migration. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. Geological Society of London, 298315.Google Scholar
Kuhn, T., Herzig, P. M., Hannington, M. D., Garbe-Schonberg, D., and Stoffers, P. (2003). Origin of fluids and anhydrite precipitation in the sediment-hosted Grimsey hydrothermal field north of Iceland. Chemical Geology, 202, 521.CrossRefGoogle Scholar
Kuhnt, W. (1992). An early Campanian paleoceanographic event in the Northern Atlantic and Western Tethys? Fourth International Conference on Paleoceanography, ICP IV, 57, 171.Google Scholar
Kuhnt, W. and Wiedmann, J. (1995). Cenomanian-Turonian source rocks: paleobiogeographic and paleoenvironmental aspects. American Association of Petroleum Geologists Studies in Geology, 40, 213–31.Google Scholar
Kukal, Z. (1990). The rate of geological processes. Earth-Science Reviews, 28, 1284.Google Scholar
Kusznir, N. J. (1991). The distribution of stress with depth in the lithosphere. Thermorheological and geodynamic constraints. Philosophical Transactions of the Royal Society of London, 337, 95110.Google Scholar
Kusznir, N. J. (2009). South Australia – Antarctica conjugate rifted margins: mapping crustal thickness and lithosphere thinning using satellite gravity inversion. A report for Geoscience Australia. https://www.ga.au/products/servlet/controller?event=GEOCAT_DETAILS&catno=68655 (accessed on August 5, 2010).Google Scholar
Kusznir, N. J., Davis, M. J., Roberts, A. M., and Baxter, K. (2000). Depth-dependent lithosphere stretching, mantle exhumation and the upper plate paradox at rifted continental margins. Eos, Transactions, American Geophysical Union, 81, 1112.Google Scholar
Kusznir, N. J. and Egan, S. S. (1989). Simple-shear and pure-shear models of extensional sedimentary basin formation: application to the Jeanne d’Arc basin, Grand Banks of Newfoundland. In Extensional Tectonics and Stratigraphy of the North Atlantic Margins, ed. Tankard, A. J. and Balkwill, H. R.. American Association of Petroleum Geologists Memoir, 46, 305–22.Google Scholar
Kusznir, N. J. and Karner, G. D., 2007. Continental lithospheric thinning and breakup in response to upwelling divergent mantle flow: application to the Woodlark, Newfoundland and Iberia margins. Geological Society of London Special Publications, 282, 389419.Google Scholar
Kusznir, N. J. and Park, R. G. (1987). The extensional strength of the continental lithosphere: its dependence on geothermal gradient, and crustal composition and thickness. Geological Society Special Publications, 28, 3552.CrossRefGoogle Scholar
Kusznir, N. J. and Ziegler, P. A. (1992). The mechanics of continental extension and sedimentary basin formation: a simple-shear/pure-shear flexural cantilever model. Tectonophysics, 215, 117–31.CrossRefGoogle Scholar
Kvenvolden, K. A. and Simoneit, B. R. T. (1990). Hydrothermally derived petroleum: examples from Guaymas Basin, Gulf of California, and Escanaba Trough, Northeast Pacific Ocean. American Association of Petroleum Geologists Bulletin, 74, 223–37.Google Scholar
Lachenbruch, A. H. (1968). Preliminary geothermal model of the Sierra Nevada. Journal of Geophysical Research, 73, 6977–89.CrossRefGoogle Scholar
Lachenbruch, A. H. (1970). Crustal temperature and heat production: implications of the linear heat-flow relation. Journal of Geophysical Research, 75, 3291–300.CrossRefGoogle Scholar
Lachenbruch, A. H. and Sass, J. H. (1977). Heat flow in the United States and the thermal regime of the crust. In The Earth’s Crust: Its Nature and Physical Properties, ed. Heacock, J. G.. American Geophysical Union, Geophysical Monograph, 20, 626–75.Google Scholar
Lachenbruch, A. H. and Sass, J. H. (1988). The stress heat-flow paradox and thermal results from Cajon Pass. Geophysical Research Letters, 15, 981–4.CrossRefGoogle Scholar
Lafargue, E. and Barker, C. (1988). Effect of water washing on crude oil compositions. American Association of Petroleum Geologists Bulletin, 72, 263–76.Google Scholar
LakeNet (2005). Global Lake Database. http://worldlakes.org/searchlakes.asp.Google Scholar
Lama, R. D. and Vutukuri, V. S. (1978). Handbook on Mechanical Properties of Rocks. Transtech Publications, Bay Village.Google Scholar
Lambe, T. W. and Whitman, R. V. (1969). Soil Mechanics. John Wiley and Sons, New York.Google Scholar
Lampe, C., Noth, S., and Ricken, W. (1999). Burial and thermal history of Cenozoic sediments in the northern Rhinegraben: a 1D simulation based on well Nordheim-1. Zentralblatt fur Geologie und Palaontologie Teil 1, 10–12, 1375–89.Google Scholar
Lampe, C. and Person, M. (2002). Advective cooling within sedimentary rift basins – application to the Upper Rhinegraben (Germany). Marine and Petroleum Geology, 19, 361–75.CrossRefGoogle Scholar
Lan, C. Y., Lee, T., and Lee, C. W. (1990). The Rb-Sr isotopic record in Taiwan gneisses and its tectonic implication. Tectonophysics, 183, 129–43.Google Scholar
Lana, M. C. (1993). Potencial petrolífero e exploração em agues profundas na Bacia de Sergipe- Alagoas. Relatório Interno Petrobrás, DEPEX / DINORD / SESEA, p. 36.Google Scholar
Landes, K. K. (1959). Petroleum Geology. John Wiley and Sons, New York, p. 443.Google Scholar
Landes, K. K., Amoruso, J. J., Charlesworth, L. J., Heany, F. Jr., and Lesperho ance, P.-J. (1960). Petroleum resources in basement rocks. American Association of Petroleum Geologists Bulletin, 44, 1682–91.Google Scholar
Lankreijer, A. C. (1998). Rheology and basement control on extensional basin evolution in Central and Eastern Europe: Variscan and Alpine-Carpathian-Pannonian tectonics. PhD thesis, Vrije University, Amsterdam, p. 158.Google Scholar
Lankreijer, A., Mocanu, V., and Cloetingh, S. (1997). Lateral variations in lithosphere strength in the Romanian Carpathians: constraints on basin evolution. Tectonophysics, 272, 269–90.CrossRefGoogle Scholar
Larsen, H.-C. and Saunders, A. (1998). Scientific Results, Ocean Drilling Program, Leg 152: Tectonism and Volcanism at the Southeast Greenland Rifted Margin: A Record of Plume Impact and Later Continental Rupture. College Station, Texas, Ocean Drilling program, pp. 503–33.Google Scholar
Larsen, P.-H. (1988). Relay structures in a Lower Permian basement-involved extension system, East Greenland. Journal of Structural Geology, 10, 38.CrossRefGoogle Scholar
Larsen, R. M. and Jarvik, L. J. (1980). The geology of the Sleipner field complex. Paper 15 in The sedimentation of the North Sea reservoir rocks, Geilo, May 11–14.Google Scholar
Larsen, S. C. and Reilinger, R. (1991). Age constraints for the present fault configuration in the Imperial Valley, California: evidence for northwestward propagation of the Gulf of California rift system. Journal of Geophysical Research, 96, 10339–46.CrossRefGoogle Scholar
Larter, S. R. and Aplin, A. C. (1995). Reservoir geochemistry: methods, applications and opportunities. In The Geochemistry of Reservoirs, ed. Cubitt, J. M. and England, W. A.. Geological Society Special Publication, 86, 532.Google Scholar
Laslett, G. M., Green, P. F., Duddy, I. R., and Gleadow, A. J. W. (1987). Thermal annealing of fission tracks in apatite, 2: a quantitative analysis. Chemical Geology, 65, 113.CrossRefGoogle Scholar
Laslett, G. M., Kendall, W. S., Gleadow, A. J. W., and Duddy, I. R. (1982). Bias in measurement of fission-track length distributions. Nuclear Tracks and Radiation Measurements, 6, 7985.CrossRefGoogle Scholar
Latter, J. H. (1987). Volcanoes and volcanic risks in the circum-Pacific. Pacific Rim Congress, 87, 745–52.Google Scholar
Laubach, S. E. (2003). Practical approaches to identifying sealed and open fractures. American Association of Petroleum Geologists Bulletin, 87, 561–79.CrossRefGoogle Scholar
Laubach, S. E., Marrett, R., and Olson, J. (2000). New directions in fracture characterization. The Leading Edge, 19, 704–11.CrossRefGoogle Scholar
Lavier, L. and Manatschal, G. (2006). A mechanism to thin the continental lithosphere at magma-poor margins. Nature, 440, 324–8.CrossRefGoogle ScholarPubMed
Lavier, L., Buck, W. R., and Poliakov, A. N. B. (1999). Self-consistent rolling-hinge model for the evolution of large-offset lowangle normal faults. Geology, 27, 1127–30.2.3.CO;2>CrossRefGoogle Scholar
Laville, E. and Petit, J.-P. (1984). Role of synsedimentary strike-slip faults in the formation of Moroccan Triassic basins. Geology, 12, 424–7.2.0.CO;2>CrossRefGoogle Scholar
Law, A., Raymond, A., White, G., Atkinson, A., Clifton, M., Atherton, T., Dawes, I. A. N., Robertson, E., Melvin, A., and Brayley, S. (2000). The Kopevik fairway, Moray Firth, UK. Petroleum Geoscience, 6, 256–74.CrossRefGoogle Scholar
Lawrence, S. R. (1990). Aspects of the petroleum geology of the Junggar basin, Northwest China. In Classic Petroleum Provinces, ed. Brooks, J.. Geological Society of London Special Publications, 50, 545–57.Google Scholar
Lay, T. and Wallace, T. C. (1995). Modern Global Seismology. Academic Press, San Diego.Google Scholar
Le Pichon, X. and Alvarez, F. (1984). From stretching to subduction in back-arc regions: dynamic considerations. Tectonophysics, 102, 343–57.CrossRefGoogle Scholar
Le Roy, P. and Piqué, A. (2001). Triassic-Liassic Western Moroccan synrift basins in relation to the central Atlantic opening. Marine Geology, 172, 359–81.CrossRefGoogle Scholar
Lear, C. H., Elderfield, H., and Wilson, P. A. (2000). Cenozoic deep-sea temperatures and global ice volumes from Mg/Ca in benthic foraminiferal calcite. Science, 287, 269–72.CrossRefGoogle ScholarPubMed
Leckie, D. A. and Rumpel, T. (2003). Tide-influenced sedimentation in a rift basin – Cretaceous Qishn Formation, Masila, Yemen – a billion barrel oil field. American Association of Petroleum Geologists Bulletin, 87, 9871013.CrossRefGoogle Scholar
Lee, C. H. and Farmer, I. (1993). Fluid Flow in Discontinuous Rocks. Chapman & Hall, London.Google Scholar
Lee, C. T. A., Lenardic, A., Cooper, C. M., Niu, F., and Levander, A. (2005). The role of chemical boundary layers in regulating the thickness of continental and oceanic thermal boundary layers. Earth and Planetary Science Letters, 230, 379–95.CrossRefGoogle Scholar
Lee, C. T. and Rudnick, R. L. (1999). Compositionally stratified cratonic lithosphere petrology and geochemistry of peridotite xenoliths from the Labait tuff cone, Tanzania. In Proceedings of the VIIth International Kimberlite Conference, ed. J. J. Gurney, M. D. Gurney, S. Pascoe, and S. R. Richardson. Red Roof Design cc, Cape Town, pp. 503–21.Google Scholar
Lee, M. J. and Hwang, Y. J. (1993). Tectonic evolution and structural styles of the East Shetland Basin. In Petroleum geology of Northwest Europe; 4th conference, ed. J. R. Parker. Shell UK Exploration and Production, London, pp. 1137–49.CrossRefGoogle Scholar
Leeder, M. R. and Gawthorpe, R. L. (1987). Sedimentary models for extensional tilt-block/half-graben basins, extensional tectonics: regional-scale processes. Geological Society of London Special Publications, 28, 139–52.CrossRefGoogle Scholar
Leeder, M. R. and Jackson, J. A. (1993). The interaction between normal faulting and drainage in active extensional basins, with examples from the western United States and central Greece. Basin Research, 5, 79102.CrossRefGoogle Scholar
Leeder, M. R., Ord, D. M., and Collier, R. (1988). Development of alluvial fans and fan deltas in neotectonic extensional settings: implications for the interpretation of basin-fills. In Fan Deltas: Sedimentology and Tectonic Settings, ed. Nemec, W. and Steel, R. J.. Blackie and Son, Glasgow, pp. 17385.Google Scholar
Lees, C. H. (1910). On the shape of the isotherms under mountain ranges in radio-active districts. Proceedings of the Royal Society of London, 83, 339–46.Google Scholar
Lemoine, M., Bas, T., Arnaud Vanneau, A., Arnaud, H., Dumont, T., Gidon, T., Bourbon, M., de Graciansky, P. C., Rudkiewicz, J. L, Mégard-Galli, J., and Tricart, P. (1986). The continental margin of the Mesozoic Tethys in the Western Alps. Marine and Petroleum Geology, 3, 179–99.CrossRefGoogle Scholar
Lenkey, L. (1999). Geothermics of the Pannonian Basin and Its Bearing on the Tectonics of Basin Evolution. Vrije University, Amsterdam, p. 215.Google Scholar
Leroueil, S. (1988). Tenth Canadian Geotechnical Colloquium: Recent developments in consolidation of natural clays. Canadian Geotechnical Journal, 25, 85107.CrossRefGoogle Scholar
Leroueil, S., Bouclin, G., Tavenas, F., Bergeron, I., and La Rochelle, P. (1990). Permeability anisotropy of natural clays as a function of strain. Canadian Geotechnical Journal, 27, 568–79.CrossRefGoogle Scholar
Leslie, S. C., Moore, G. F., Morgan, J. K., and Hills, D. J. (2002). Seismic stratigraphy of the Frontal Moat: implications for sedimentary processes at the leading edge of an oceanic hotspot trace. Marine Geology, 184, 143–62.CrossRefGoogle Scholar
Lespinasse, M. C., Leroy, J. L., Pironon, J., and Boiron, M.-C. (1998). The paleofluids from the marginal ridge of the Côte d’Ivoire-Ghana transform margin (Hole 960A) as thermal indicators. In Proceedings of Ocean Drilling Programme, ed. Mascle, J., Lohmann, G. P., and Moullade, M.. Scientific Results, 159, 4952.Google Scholar
Levitte, D., Maurath, G., and Eckstein, Y. (1984). Terrestrial heat flow in a 3.5 km deep borehole in the Jordan-Dead Sea rift valley. Geological Society of America Abstracts with Programs, 16, 575.Google Scholar
Lewan, M. D. (1993). Laboratory simulation of petroleum formation: hydrous pyrolysis. In Organic Geochemistry, ed. Engel, M. H. and Macko, S. A.. Plenum Press, pp. 419–42.Google Scholar
Lewan, M. D. (1994). Assessing natural oil expulsion from source rocks by laboratory pyrolysis. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 201–10.Google Scholar
Lewan, M. D. (1997). Experiments on the role of water in petroleum formation. Geochimica et Cosmochimica Acta, 61, 3691–723.CrossRefGoogle Scholar
Lewan, M. D. and Henry, A. A. (2001). Gas: oil ratios for source rocks containing type-I, -II, -IIS, and -III kerogens as determined by hydrous pyrolysis. In Geologic Studies of Deep Natural Gas Resources, ed. Dyman, T. S. and Kuuskraa, V. A.. U.S. Geological Survey Digital Data Series, 67, 19.Google Scholar
Lewan, M. D., Kotarba, M. J., Curtis, J. B., Wieclaw, D., and Kosakowski, P. (2006). Oil-generation kinetics for organic facies with Type-II and -IIS kerogen in the Menilite Shales of the Polish Carpathians. Geochimica et cosmochimica acta, 70, 3351–68.CrossRefGoogle Scholar
Lewis, C. R. and Rose, S. C. (1970). A theory relating high temperatures and overpressures. Journal of Petroleum Technology, 22, 1116.CrossRefGoogle Scholar
Lewis, R. Q. Sr and Campbell, R. H. (1965). Geology and uranium deposits of Elk Ridge and vicinity, San Juan County, Utah. Publications of the U. S. Geological Survey, Reston, P 0474-B.Google Scholar
Leythaeuser, D., Little, R., Radke, M., and Schaefer, R. G. (1987). Geochemical effects of petroleum migration and expulsion from Toarcian source rocks in the Hils syncline area, NW-Germany. Organic Geochemistry, 13, 489502.CrossRefGoogle Scholar
Leythaeuser, D. and Poelchau, H. S. (1991). Expulsion of petroleum from type III kerogen source rocks in gaseous solution: modeling of solubility fractionation. In Petroleum Migration, ed. England, W. A. and Fleet, A. J.. Geological Society of London Special Publications, 59, 3346.Google Scholar
Leythaeuser, D. and Ruckheim, J. (1989). Heterogeneity of oil composition within a reservoir as a reflection of accumulation history. Geochimica et Cosmochimica Acta, 53, 2119–23.CrossRefGoogle Scholar
Leythaeuser, D., Schaefer, R. G., and Radke, M. (1988). Geochemical effects of primary migration of petroleum in Kimmeridge source rocks from Brae Field area, North Sea; I, Gross composition of C15-soluble organic matter and molecular composition of C15-saturated hydrocarbons. Geochimica et Cosmochimica Acta, 52, 701–13.Google Scholar
Lezzar, K. E., Tiercelin, J.-J., Turdu, C. L., Cohen, A. S., Reynolds, D. J., Le Gall, B., and Scholz, C. A. (2002). Control of normal fault interaction on the distribution of major Neogene sedimentary depocenters, Lake Tanganyika, East African Rift. American Association of Petroleum Geologists Bulletin, 86, 1027–59.Google Scholar
Li, R., Wang, Z., Lin, D., and Xin, M. (1988). Organic geochemical analysis of environments of Dongpu basin. Chinese Journal of Geochemistry, 7, 193206.Google Scholar
Li, R., Wu, H., Lin, D., Wang, Z., Chen, R., Tian, X., and Zhang, R. (1992). Biomarker features of the Chinese Meso-Cenozoic saline lake deposits. Advances in Geoscience, 2, 275–92.Google Scholar
Li, Y.-G., Chen, P., Cochran, E. S., Vidale, J. E., and Burdette, T. (2006). Seismic evidence for rock damage and healing on the San Andreas Fault associated with the 2004 M 6.0 Parkfield earthquake. Bulletin of the Seismological Society of America, 96, 349–63.CrossRefGoogle Scholar
Li, Y.-G., Ellsworth, W. L., Thurber, C. H., Malin, P. E., and Aki, K. (1997). Fault-zone guided waves from explosions in the San Andreas Fault at Parkfield and Cienga Valley, California. Bulletin of the Seismological Society of America, 87, 210–21.CrossRefGoogle Scholar
Li, Y.-G., Vidale, J. E., and Cochran, E. S. (2004). Low-velocity damaged structure of the San Andreas Fault at Parkfield from fault trapped waves. Geophysical Research Letters, 31, L12S06.CrossRefGoogle Scholar
Li, Y. H. (1976). Denudation of Taiwan island since the Pliocene epoch. Geology, 4, 105–7.2.0.CO;2>CrossRefGoogle Scholar
Ligi, M., Bonatti, E., Bortoluzzi, G., Brunelli, D., Caratori Tontini, F., Cipriani, A., Cocchi, I., Cuffaro, M., Ferrante, V., Khalil, S., Mitschell, N. C., Rasul, N., and Schetino, A. (2008). Sea-floor spreading initiation: constraints from geophysical data of the Thetis Deep, northern Red Sea. Abstract T33F-08, EOS Transactions, 89, 53.Google Scholar
Ligi, M., Bonatti, E., Tontini, F. C., Cipriani, A., Cocchi, L., Schettino, A., Bortoluzzi, G., Ferrante, V., Khalil, S., Mitchell, N. C., and Rasul, N. (2011). Initial burst of oceanic crust accretion in the Red Sea due to edge-driven mantle convection. Geology, 39, 1019–22.CrossRefGoogle Scholar
Ligtenberg, J. H. and Thomsen, R. O. (2003). Fluid migration path detection and its application to basin modeling. EAGE 65th Conference, Stavanger, Norway, Extended abstracts, C27.Google Scholar
Lillie, R. J. (1999). Whole Earth Geophysics. An Introductory Textbook for Geologists and Geophysicists. Prentice Hall, Upper Saddle River, New Jersey, p. 361.Google Scholar
Lin, G., Nunn, J. A., and Deming, D. (2000). Thermal buffering of sedimentary basins by basement rocks: implications arising from numerical simulations. Petroleum Geoscience, 6, 299307.CrossRefGoogle Scholar
Lindquist, S. J. (1983). Nugget formation reservoir characteristics affecting production in the overthrust belt of southwestern Wyoming. Journal of Petroleum Technology, 35, 1355–65.CrossRefGoogle Scholar
Liotta, D. and Ranalli, G. (1999). Correlation between seismic reflectivity and rheology in extended lithosphere: southern Tuscany, inner Northern Apennines, Italy. Tectonophysics, 315, 109–22.CrossRefGoogle Scholar
Lippman, P. W., Logatchev, N. A., and Zorin, Y. A. (1989). Intra-continental rift comparisons. EOS Transactions, 70, 578–88.Google Scholar
Lippolt, H. J., Leitz, M., Wernicke, R. S., and Hagedorn, B. (1994). (Uranium+thorium)/helium dating of apatite: experience with samples from different geochemical environments. Chemical Geology, 112, 179–91.CrossRefGoogle Scholar
Liro, L. M. and Coen, R. (1995). Salt deformation history and postsalt structural trends, offshore southern Gabon, West Africa. American Association of Petroleum Geologists Memoir, 65, 323–31.Google Scholar
Liro, L. M. and Pardus, Y. C. (1990). Seismic facies analysis of fluvial-deltaic lacustrine systems – upper Fort Union Formation (Paleocene), Wind River basin, Wyoming. In Lacustrine Basin Exploration – Case Studies and Modern Analogs, ed. Katz, B. J.. American Association of Petroleum Geologists Memoir, 50, 225–42.Google Scholar
Lisker, F. and Fachmann, S. (2001). Phanerozoic history of the Mahanadi region, India. Journal of Geophysical Research, 106, 22027–50.CrossRefGoogle Scholar
Lister, G. S. and Davis, G. A. (1989). The origin of metamorphic core complexes and detachment faults formed during Tertiary continental extension in the northern Colorado River region, U.S.A. Journal of Structural Geology, 11, 6594.CrossRefGoogle Scholar
Lister, G. S., Etheridge, M. A., and Symonds, P. A. (1986). Detachment faulting and the evolution of passive continental margins. Geology, 14, 246–50.Google Scholar
Lister, J.-R. and Kerr, R.-C. (1991). Fluid-mechanical models of crack propagation and their application to magma transport in dykes. Journal of Geophysical Research, 96, 10049–77.CrossRefGoogle Scholar
Littke, R. (1994). Deposition, Diagenesis, and Weathering of Organic Matter-Rich Sediments. Springer-Verlag, Heidelberg.Google Scholar
Littke, R., Baker, D. R., and Leythaeuser, D. (1988). Microscopic and sedimentologic evidence for the generation and migration of hydrocarbons in Toarcian source rocks of different maturities. Organic Geochemistry, 13, 549–59.CrossRefGoogle Scholar
Littke, R., Baker, D. R., and Rullkötter, J. (1997). Deposition of petroleum source rocks. In Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling, ed. Welte, D. H., Horsfield, B., and Baker, D. R.. Springer-Verlag, Berlin, pp. 273333.Google Scholar
Littke, R., Kroos, B., Idiz, E., and Frielingsdorf, J. (1995). Molecular nitrogen in natural gas accumulations: generation from sedimentary organic matter at high temperatures. American Association of Petroleum Geologists Bulletin, 9, 410–30.Google Scholar
Liu, L., Sun, X. M., Dong, F. X., Ma, F., and Ma, Y. P. (2004). Geochemical characteristics and fluid inclusion in calcite veins of lower part of member 1 of Shahejie Formation, offshore area, Dagang oil field: a case study of well Gangshen 67 (in Chinese). Journal of Jilin University, Earth Science Edition, 34, 4954.Google Scholar
Livera, S. E. and Gdula, J. E. (1990). Brent oil field. AAPG Treatise of Petroleum Geology, Atlas of Oil and Gas Fields, A-017, 2163.Google Scholar
Lizarralde, D. and Holbrook, W. S. (1997). U.S. mid-Atlantic Margin structure and early thermal evolution. Journal of Geophysical Research, 102, 22855–75.CrossRefGoogle Scholar
Lockner, D. A., Moore, D. E., and Reches, Z. (1992). Microcrack interaction leading to shear fracture. 33rd Symposium on Rock Mechanics, pp. 807–16.Google Scholar
Logan, J. M., Friedman, M., Higgs, M., Dengo, C., and Shimamoto, T. (1979). Experimental studies of simulated gouge and their application to studies of natural fault zones. Proceedings of VIII Conference on Analysis of Actual Fault Zones in Bedrock, US Geological Survey, Menlo Park, pp. 305–43.Google Scholar
Lomando, A. J. and Engelder, T. (1984). Strain indicated by calcite twinning: implications for deformation of the early Mesozoic northern Newark basin, New York. Northeastern Geology, 6, 192–5.Google Scholar
Lombardo, B., Rubatto, D., and Castelli, D. (2002). Ion microprobe U-Pb dating of zircon from a Monviso metaplagiogranite: implications for the evolution of the Piedmont-Liguria Tethys in the Western Alps. Ofioliti, 27, 108–17.Google Scholar
Lonergan, L. and Cartwright, J. A. (1999). Polygonal faults and their influence on deep-water sandstone reservoir geometries, Alba field, United Kingdom central North Sea. American Association of Petroleum Geologists Bulletin, 83, 410–32.Google Scholar
Lonergan, L., Lee, N., Johnson, H. D., Cartwright, J. A., and Jolly, R. J. H. (2000). Remobilization and injection in deepwater depositional systems: implications for reservoir architecture and prediction. In Deep Water Reservoirs of the World, ed. Weimer, P., Slatt, R. M., Coleman, J., Rosen, N. C., Nelson, H., Bouma, A. H., Styzen, M. J., and Lawrence, D. T.. Gulf Coast Section SEPM Foundation, 20th annual conference, Houston, pp. 515–32.Google Scholar
Lonsdale, P. (1989). Geology and tectonic history of the Gulf of California. In The Eastern Pacific Ocean and Hawaii, ed. Winterer, E. L., Hussong, D. M., and Decker, R. W.. Geological Society of America, Denver, pp. 499522.Google Scholar
Lonsdale, P. (1991). Structural patterns of the Pacific floor offshore of peninsular California. American Association of Petroleum Geologists Memoir, 47, 87125.Google Scholar
Lopatin, N. V. (1971). Temperature and geologic time as factors in coalifications (in Russian). Akademiya Nauk SSR Izvestiya Seriya Geologicheskaya, 3, 95106.Google Scholar
Losh, S. (1998). Oil migration in a major growth fault: structural analysis of the Pathfinder core, South Eugene Island Block 330, offshore Louisiana. American Association of Petroleum Geologists Bulletin, 82, 1694–710.Google Scholar
Losh, S., Eglinton, L., Shoell, M., and Wood, J. (1999). Vertical and lateral fluid flow related to a large growth fault, South Eugene Island Block 330 field, offshore Louisiana. American Association of Petroleum Geologists Bulletin, 83, 244–76.Google Scholar
Losh, S., Walter, L., Meulbroek, P., Martini, A., Cathles, L., and Whelan, J. (2002). Reservoir fluids and their migration into the South Eugene Island Block 330 reservoirs, offshore Louisiana. American Association of Petroleum Geologists Bulletin, 86, 1463–88.Google Scholar
Louvel, V., Dyment, J., and Sibuet, J.-C. (1997). Thinning of the Goban Spur continental margin and formation of early oceanic crust: constraints from forward modelling and inversion of marine magnetic anomalies. Geophysical Journal International, 128, 188–96.CrossRefGoogle Scholar
Lowell, J. D. (1995). Mechanics of basin inversion from worldwide examples. In Basin Inversion, ed. Buchanan, J. G. and Buchanan, P. G.. Geological Society of London Special Publications, 88, 3957.Google Scholar
Lowell, J. D. and Genik, G. J. (1972). Sea-floor spreading and structural evolution of the southern Red Sea. American Association of Petroleum Geologists Bulletin, 56, 247–59.Google Scholar
Lowell, R. P. (1975). Circulation in fractures, hot springs, and convective heat transport on mid-ocean ridge crests. Geophysical Journal of the Royal Astronomical Society, 40, 351–65.Google Scholar
Lowell, R. P. and Germanovich, L. N. (1994). On the thermal evolution of high-temperature hydrothermal systems at ocean ridge crests. Journal of Geophysical Research, 99, 565–75.CrossRefGoogle Scholar
Lowell, R. P. and Rona, P. A. (1985). Hydrothermal models of the generation of massive sulfide ore deposits. Journal of Geophysical Research, 90, 8769–83.CrossRefGoogle Scholar
Lucas, M., Hull, J., and Manspeizer, W. (1988). A foreland-type fold and related structures in the Newark rift basin. In Triassic–Jurassic Rifting, Continental Breakup and the Origin of the Atlantic Ocean Passive Margins, part A, ed. Manspeizer, W.. New York, Elsevier, pp. 307–32.Google Scholar
Ludwig, W. J., Nafe, J. E., and Drake, C. L. (1970). Seismic refraction. In The Sea, ed. Maxwell, A. E.. Willey-Interscience, New York, pp. 5384.Google Scholar
Lundegard, P. D. (1992). Sandstone porosity loss – a “big picture” view of the importance of compaction. Journal of Sedimentary Petrology, 62, 250–60.CrossRefGoogle Scholar
Lüning, S., Craig, J., Loydell, D. K., Storch, P., and Fitches, B. (2000). Lower Silurian “hot shales” in North Africa and Arabia: regional distribution and depositional model. Earth-Science Reviews, 49, 121200.CrossRefGoogle Scholar
Lupini, J. F., Skinner, A. E., and Vaugham, P. R. (1981). The drained residual strength of cohesive soils. Geotechnique, 31, 181213.CrossRefGoogle Scholar
Lyberis, N. (1988). Tectonic evolution of the Gulf of Suez and the Gulf of Aqaba. Tectonophysics, 153, 209–20.CrossRefGoogle Scholar
Macaulay, C. I., Fallick, A. E., Haszeldine, R. S., and Macaulay, G. E. (2000). Oil migration makes the difference: regional distribution of carbonate cement δ13C in northern North Sea Tertiary sandstones. Clay Minerals, 35, 6976.CrossRefGoogle Scholar
MacCaig, A. M. (1988). Deep fluid circulation in fault zones. Geology, 16, 867–70.Google Scholar
MacCready, T., Snoke, A. W., Wright, J. E., and Howard, K. A. (1997). Mid-crustal flow during Tertiary extension in the Ruby Mountains core complex, Nevada. Geological Society of America Bulletin, 109, 1576–94.2.3.CO;2>CrossRefGoogle Scholar
MacDonald, I. R., Reilly, J. F., Best, S. E., Venkataramaiah, R., Sassen, R., Guinasso, N. L. Jr., and Amos, J. (1996). Remote sensing inventory of active oil seeps and chemosynthetic communities in the northern Gulf of Mexico. In Hydrocarbon Migration and Its Near-Surface Expression, ed. Schumacher, D. and Abrams, M. A.. American Association of Petroleum Geologists Memoir, 66, 2737.Google Scholar
Macdonald, K. C. (1982). Mid-ocean ridges; fine scale tectonic, volcanic and hydrothermal processes within the plate boundary zone. Annual Review of Earth and Planetary Sciences, 10, 155–90.CrossRefGoogle Scholar
Macedo, J. and Marshak, S. (1999). Controls on the geometry of fold-thrust belt salients. Geological Society of America Bulletin, 111, 1808–22.2.3.CO;2>CrossRefGoogle Scholar
Macgregor, D. S. (1993). Relationships between seepage, tectonics and subsurface petroleum reserves. Marine and Petroleum Geology, 10, 606–19.CrossRefGoogle Scholar
Macgregor, D. S. (1995). Hydrocarbon habitat and classification of inverted rift basins. In Basin Inversion, ed. Buchanan, J. G. and Buchanan, P. G.. Geological Society of London Special Publications, 88, 8393.Google Scholar
Machette, M. N., Personius, S. F., and Nelson, A. R. (1991a). Paleoseismology of the Wasatch fault zone: a summary of recent investigations, interpretations, and conclusions. U.S. Geological Survey Professional Paper, 1500.Google Scholar
Machette, M. N., Personius, S. F., Nelson, A. R., Schwartz, D. P., and Lund, W. R. (1991b). The Wasatch fault zone, Utah: segmentation and history of Holocene earthquakes. Journal of Structural Geology, 13, 137–49.CrossRefGoogle Scholar
Mackenzie, A. S., Leythaeuser, D., Altebaeumer, F.-J., Disko, U., and Rullkoetter, J. (1988). Molecular measurements of maturity for Lias delta shales in N.W. Germany. Geochimica et Cosmochimica Acta, 52, 1145–54.CrossRefGoogle Scholar
Mackenzie, A. S. and Quigley, T. M. (1988). Principles of geochemical prospect appraisal. American Association of Petroleum Geologists Bulletin, 72, 399415.Google Scholar
MacLeod, C. J., Escartín, J., Banerji, D., Banks, G. J., Gleeson, M., Irving, D. H. B., Lilly, R. M., McCaig, A. M., Niu, Y., Allerton, S., and Smith, D. K. (2002). Direct geological evidence for oceanic detachment faulting: the Mid-Atlantic Ridge, 15 degrees 45’N. Geology, 30, 879–82.2.0.CO;2>CrossRefGoogle Scholar
Magara, K. (1976). Water expulsion from elastic sediments during compaction: directions and volumes. American Association of Petroleum Geologists Bulletin, 60, 543–53.Google Scholar
Maggi, A., Jackson, J. A., McKenzie, D., and Priestley, K. (2000a). Earthquake focal depths, effective elastic thickness, and the strength of the continental lithosphere. Geology, 28, 495–8.2.0.CO;2>CrossRefGoogle Scholar
Maggi, A., Jackson, J. A., Priestley, K., and Baker, C. (2000b). A re-assessment of focal depth distributions in southern Iran, the Tien Shan and northern India: Do earthquakes really occur in the continental mantle? Geophysical Journal International, 143, 629–61.CrossRefGoogle Scholar
Magnavita, L. P. (1992). Geometry and kinematics of the Reconcavo-Tucano-Jatoba rift, NE Brazil. PhD thesis, University of Oxford, England.Google Scholar
Magnavita, L. P. (2000). Deformation mechanisms in porous sandstones: implications for the development of fault seal and migration paths in the Reconcavo basin, Brazil. In Petroleum Systems of South Atlantic Margins, ed. Mello, M. P. and Katz, B. J.. American Association of Petroleum Geologists Memoir, 73, 195212.Google Scholar
Magoon, L. B. and Dow, W. G. (1994). The petroleum system. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 324.CrossRefGoogle Scholar
Magoon, L. B., Hudson, T. L., and Peters, K. E. (2005). Egret-Hibernia(!), a significant petroleum system, northern Grand Banks area, offshore eastern Canada. American Association of Petroleum Geologists Bulletin, 89, 1203–37.CrossRefGoogle Scholar
Maher, C. E. (1980). The Piper Oilfield. In Giant Oil and Gas Fields of the Decade: 1968–1978. American Association of Petroleum Geologists Memoir, 30, 131–72.Google Scholar
Maher, C. E. (1981). The Piper Oilfield. In Petroleum Geology of the Continental Shelf of North-West Europe, ed. Illing, L. V. and Hobson, G. D.. Heyden and Son, London, pp. 358–70.Google Scholar
Maher, C. E. and Harker, S. D. (1987). The Claymore oil field. In Petroleum Geology of North West Europe, ed. Brooks, J. and Glennie, K. W.. Graham and Trotman, London, pp. 835–45.Google Scholar
Mahoney, J. and Coffin, M. F. (1997). Large igneous provinces: continental, oceanic and planetary flood volcanism. AGU Geophysical Monograph, 100, 438.Google Scholar
Mailloux, B. J., Person, M., Kelley, S., Dunbar, N., Cather, S., Strayer, L., and Hudleston, P. (1999). Tectonic controls on the hydrogeology of the Rio Grande rift, New Mexico. Water Resources Research, 35, 2641–59.CrossRefGoogle Scholar
Majer, E. L. and McEvilly, T. V. (1979). Seismological investigations at The Geysers geothermal field. Geophysics, 44, 246–69.CrossRefGoogle Scholar
Makhous, M. and Galushkin, Y. I. (2005). Basin Analysis and Modeling of the Burial, Thermal and Maturation Histories in Sedimentary Basins. Technip, Paris, p. 380.Google Scholar
Malahoff, A., McMurty, G. M., Wiltshire, J. C., and Yeh, H.-W. (1982). Geology and geochemistry of hydrothermal deposits from active submarine volcano Loihi, Hawaii. Nature, 298, 234–9.CrossRefGoogle Scholar
Maltman, A. J. (1994). Prelithification deformation. In Continental Deformation, ed. Hancock, P. L.. Pergamon Press, Tarrytown, pp. 143–58.Google Scholar
Mammerickx, J. and Sandwell, D. T. (1986). Rifting of old oceanic lithosphere. Journal of Geophysical Research, 91, 1975–88.CrossRefGoogle Scholar
Manatschal, G. (2004). New models for evolution of magma-poor rifted margins based on a review of data and concepts from West Iberia and the Alps. International Journal of Earth Sciences, 93, 432–66.CrossRefGoogle Scholar
Manatschal, G. and Bernoulli, D. (1998). Rifting and early evolution of ancient ocean basins: the record of the Mesozoic Tethys and the Galicia-Newfoundland margins. Marine Geophysical Researches, 20, 371–81.CrossRefGoogle Scholar
Manatschal, G. and Bernoulli, D. (1999). Architecture and tectonic evolution of nonvolcanic margins; present-day Galicia and ancient Adria. Tectonics, 18, 1099–119.CrossRefGoogle Scholar
Manatschal, G., Froitzheim, N., Rubenach, M., and Turrin, B. D. (2001). The role of detachment faulting in the formation of an ocean-continent transition: insights from the Iberia Abyssal Plain. In Non-volcanic Rifting of Continental Margins: A Comparison of Evidence from Land and Sea, ed. Wilson, R. C. L., Whitmarsh, R. B, Taylor, B., and Froitzheim, N.. Geological Society of London Special Publications, 187, 405–28.Google Scholar
Manatschal, G. and Niegervelt, P. (1997). A continent-ocean transition recorded in the Err and Platta nappes (eastern Switzerland). Eclogae Geologicae Helvetiae, 90, 327.Google Scholar
Mancktelow, N. S. and Grasemann, B. (1997). Time-dependent effects of heat advection and topography on cooling histories during erosion. Tectonophysics, 270, 167–95.CrossRefGoogle Scholar
Mandl, G. (1987). Tectonic deformation by rotating parallel faults: the “bookshelf” mechanism. Tectonophysics, 141, 277316.CrossRefGoogle Scholar
Mandl, G. (1988). Mechanics of Tectonic Faulting: Models and Basic Concepts. Developments in Structural Geology. Elsevier Science Publications, Amsterdam, Netherlands (NLD).Google Scholar
Mango, F. D. (1997). The light hydrocarbons in petroleum: a critical review. Organic Geochemistry, 26, 417–40.CrossRefGoogle Scholar
Mann, D. M. and Mackenzie, A. S. (1990). Prediction of pore fluid pressures in sedimentary basins. Marine and Petroleum Geology, 7, 5565.CrossRefGoogle Scholar
Mann, P., Gahagan, L., and Gordon, M. B. (2004). Tectonic setting of the world’s giant oil and gas fields. In Giant Oil and Gas Fields of the Decade 1990–1999, ed. Halbouty, T.. American Association of Petroleum Geologists Memoir, 78, 15105.Google Scholar
Mann, U. (1994). An integrated approach to the study of primary petroleum migration. In Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins, ed. Parnell, J.. Geological Society of London Special Publications, 78, 233–60.Google Scholar
Mann, U., Hantschel, T., Schaefer, R. G., Krooss, B., Leythaeuser, D., Littke, R., and Sachsenhofer, R. F. (1997). Petroleum migration: mechanisms, pathways, efficiencies and numerical simulations. In Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling, ed. Welte, D. H., Horsfield, B., and Baker, D. R.. Springer, Berlin, pp. 405520.Google Scholar
Marcano, M., Lohman, K. C., and Pickett, E. A. (1998). Geochemistry of pore-filling and fracture-vein carbonates: Cote d’Ivoire–Ghana Marginal Ridge. Proceedings of the Ocean Drilling Program, Scientific Results, 159, 7180.Google Scholar
Marks, K. M. and Stock, J. M. (1994). Variations in ridge morphology and depth-age relationships on the Pacific-Antarctic Ridge. Journal of Geophysical Research, 99.B1, 531541.CrossRefGoogle Scholar
Mart, Y. and Dauteuil, O. (2000). Analogue experiments of propagation of oblique rifts. Tectonophysics, 316, 121–32.CrossRefGoogle Scholar
Martel, S. J., Pollard, D. D., and Segall, P. (1988). Development of simple strike-slip fault zones, Mount Abbot Quadrangle, Sierra Nevada, California. Geological Society of America Bulletin, 100, 1451–65.2.3.CO;2>CrossRefGoogle Scholar
Martin, M. A. and Pollard, J. E. (1996). The role of trace fossil (ichnofabric) analysis in the development of depositional models for the Upper Jurassic Fulmar Formation of the Kittiwake Field (Quadrant 21, UKCS). In Geology of the Humber Group: Central Graben and Moray Firth, UKCS, ed. Hurst, A., Johnson, H. D., Burley, S. D., Canham, A. C., and Mackertich, D. S.. Special Publication of the Geological Society of London, 114, 163–83.Google Scholar
Martin, R. G. (1978). Northern and eastern Gulf of Mexico continental margin; stratigraphic and structural framework. In Framework, Facies, and Oil-Trapping Characteristics of the Upper Continental Margin, ed. Bouma, A. H., Moore, G. T., and Coleman, J. M.. American Association of Petroleum Geologists Studies in Geology, 7, 2142.Google Scholar
Martínez-Martínez, J. M., Soto, J. I., and Balanyá, J. C. (2004). Domes in extended orogens: a mode of mountain rift in the Betics, southeast Spain. Special Paper – Geological Society of America, 380, 243–65.Google Scholar
Martini, A. M., Walter, L. M., Ku, T. C. W., Budai, J. M., McIntosh, J. C., and Schoell, M. (2003). Microbial production and modification of gases in sedimentary basins: a geochemical case study from a Devonian shale gas play, Michigan basin. American Association of Petroleum Geologists Bulletin, 87, 1355–75.CrossRefGoogle Scholar
Marton, L. G., Tari, G. C., and Lehmann, C. T. (2000). Evolution of the Angolan passive margin, West Africa, with emphasis on post-salt structural styles. Geophysical Monograph, 115, 129–49.Google Scholar
Marzoli, A., Bertrand, H., Knight, K. B., Cirilli, S., Buratti, N., Vérati, C., Nomade, S., Renne, P. R., Youbi, N., Martini, R., Allenbach, K., Neuwerth, R., Rapaille, C., Zaninetti, L., and Bellieni, G. (2004). Synchrony of the Central Atlantic Magmatic Province and the Triassic-Jurassic boundary climatic and biotic crisis. Geology, 32, 973–6.CrossRefGoogle Scholar
Mascle, J., Lohmann, G.P., Clift, P.D. et al. (1996). Proc. ODP Init. Repts., 159, College Station, TX (Ocean Drilling Program).CrossRefGoogle Scholar
Mascle, J., Lohmann, G.P., and Clift, P. (1997). Development of a passive transform margin: Cote d’Ivoire-Ghana transform margin – ODP Leg 159 preliminary results. Geo-Marine Letters, 17, 4–11.CrossRefGoogle Scholar
Massari, F. and Colella, A. (1988). Evolution and types of fan-delta systems in some major tectonic settings. In Fan Deltas: Sedimentology and Tectonic Settings, ed. Nemec, W. and Steel, R. J.. Blackie and Son, Glasgow, pp. 103–22.Google Scholar
Masson, D. G. (1996). Catastrophic collapse of the volcanic island of Hierro 15 ka ago and the history of landslides in the Canary Islands. Geology, 24, 231–4.2.3.CO;2>CrossRefGoogle Scholar
Mattavelli, L. and Novelli, L. (1988). Geochemistry and habitat of natural gases in Italy. Organic Geochemistry, 13, 113.CrossRefGoogle Scholar
Matthews, M. D. (1999). Migration of petroleum. In Exploring for Oil and Gas Traps, ed. Beaumont, E. A. and Foster, N. H.. American Association of Petroleum Geologists Treatise of Petroleum Geology, 7, 31–8.Google Scholar
Matyas, J. (1996). How detrical mica affects compactional porosity loss. Terra Nova, 8, 425–9.CrossRefGoogle Scholar
Mavko, G. M. and Nur, A. (1979). Wave attenuation in partially saturated rocks. Geophysics, 44, 161–78.CrossRefGoogle Scholar
May, S. R., Ehman, K. D., Gray, G. G., and Crowell, J. C. (1993). A new angle on the tectonic evolution of the Ridge Basin, a “strike-slip” basin in Southern California. Geological Society of America Bulletin, 105, 1357–72.2.3.CO;2>CrossRefGoogle Scholar
Mazzini, A., Duranti, D., Jonk, R., Parnell, J., Cronin, B., Hurst, A., and Quine, M. (2003). Palaeo-carbonate seep structures above an oil reservoir, Gryphon field, Tertiary, North Sea. Geomarine Letters, 23, 323–39.Google Scholar
McAfee, A. (2005). Hinge zone control on reservoir quality in the West African syn-rift succession. The 4th HGS/PESGB International Conference on African E7P, Houston Program, 9.Google Scholar
McAuliffe, C. D. (1979). Oil and gas migration: chemical and physical constraints. American Association of Petroleum Geologists Bulletin, 73, 1455–71.Google Scholar
McBirney, A. R. (1995). Mechanisms of differentiation in the Skaergaard intrusion. Journal of Geological Society of London, 152, 421–35.CrossRefGoogle Scholar
McBride, B. C. (1997). The geometry and evolution of allochthonous salt and its impact on petroleum systems, northern Gulf of Mexico basin: studies in three and four-dimensional analysis. PhD thesis, University of Colorado, Boulder, p. 276.Google Scholar
McBride, B. C., Rowan, M. G., and Weimer, P. (1998a). The evolution of allochthonous salt systems, Northern Green Canyon and Ewing Bank (Offshore Louisiana), Northern Gulf of Mexico. American Association of Petroleum Geologists Bulletin, 82, 1013–36.Google Scholar
McBride, B. C., Weimer, P., and Rowan, M. G. (1998b). The effect of allochthonous salt on petroleum systems of Northern Green Canyon and Ewing Bank (Offshore Louisiana), Northern Gulf of Mexico. American Association of Petroleum Geologists Bulletin, 82, 1083–112.Google Scholar
McBride, J. H. (1991). Constraints on the structure and tectonic development of the Early Mesozoic South Georgia rift, southeastern United States: seismic reflection data processing and interpretation. Tectonics, 10, 1065–83.CrossRefGoogle Scholar
McBride, J. H. and Brown, L. D. (1986). Reanalysis of the COCORP deep seismic reflection profile across the San Andreas fault, Parkfield, California. Bulletin of the Seismological Society of America, 76, 1668–86.Google Scholar
McCabe, P. J. and Parrish, J. T. (1992). Tectonic and climatic controls on the distribution and quality of Cretaceous coals. In Controls on the Distribution and Quality of Cretaceous Coals, ed. McCabe, P. J. and Parrish, J. T.. Geological Society of America Special Paper, 267, 115.Google Scholar
McCaffrey, M. A., Dahl, J. E., Sundararaman, P., Moldowan, J. M., and Schoell, M. (1994). Source quality determination from oil biomarkers II – a case study using Tertiary-reservoired Beaufort Sea oils. American Association of Petroleum Geologists Bulletin, 10, 1527–40.Google Scholar
McClay, K. R. (1995). The geometries and kinematics of inverted fault systems: a review of analogue model studies. In Basin Inversion, ed. Buchanan, J. G. and Buchanan, P. G.. Geological Society of London Special Publications, 88, 97118.Google Scholar
McClay, K. R., Dooley, T., Whitehouse, P., and Mills, M. (2002). 4-D evolution of rift systems: insights from scaled physical models. American Association of Petroleum Geologists Bulletin, 86, 935–59.Google Scholar
McClay, K. R. and Khalil, S. (1998). Extensional hard linkages, eastern Gulf of Suez, Egypt. Geology, 26, 563–6.2.3.CO;2>CrossRefGoogle Scholar
McClay, K. R., Nichols, G. J., Khalil, S. M., Darwish, M., and Bosworth, W. (1998). Extensional tectonics and sedimentation, eastern Gulf of Suez, Egypt. In Sedimentation and Tectonics of Rift Basins: Red Sea Gulf of Aden, ed. Purser, B. H. and Bosence, D. W. J.. Chapman and Hall, London, pp. 223–38.Google Scholar
McClure, N. M. and Brown, A. A. (1992). Miller Field: a subtle Upper Jurassic submarine fan trap in the South Viking Graben, U.K. Sector, North Sea. In Giant Oil and Gas Fields of the Decade: 1978–1988, ed. Halbouty, M. T.. American Association of Petroleum Geologists Memoir, 54, 1519.Google Scholar
McDougall, I. and Harrison, T. M. (1988). Geochronology and thermochronology by the (super 40) Ar/ (super 39) Ar method. Oxford Monographs on Geology and Geophysics, 9, 212.Google Scholar
McGann, G. J., Green, S. C. H., Harker, S. D., and Romani, R. S. (1991). The Scapa Field, Block 14/19, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 39376.Google Scholar
McGee, K. A. and Gerlach, T. M. (1998). Annual cycle of magmatic CO2 in a tree-kill soil at Mammoth Mountain, California: implications for soil acidification. Geology, 26, 463–6.2.3.CO;2>CrossRefGoogle Scholar
McGill, G. E. and Stromquist, A. W. (1975). Origin of graben in the Needles District, Canyonlands National Park, Utah. Field Symposium – Guidebook of the Four Corners Geological Society, 8, 235–43.Google Scholar
McGill, G. E. and Stromquist, A. W. (1979). The grabens of Canyonlands National Park, Utah: geometry, mechanics, and kinematics. Journal of Geophysical Research, 84, 4547–63.CrossRefGoogle Scholar
McHone, J. G. (1996). Broad-terrane Jurassic flood basalts across northeastern North America. Geology, 24, 319–22.Google Scholar
McHone, J. G. (2000). Non-plume magmatism and rifting during the opening of the central Atlantic Ocean. Tectonophysics, 316, 287–96.CrossRefGoogle Scholar
McHone, J. G. and Puffer, J. H. (2003). Flood Basalt Provinces of the Pangean Atlantic Rift: Regional Extent and Environmental Significance. Columbia University Press, New York.Google Scholar
McInnes, B. I. A., Farley, K. A., Sillitoe, R. H., and Kohn, B. P. (1999). Application of apatite (U-Th)/He thermochronometry to the determination of the sense and amount of vertical fault displacement at the Chuquicamata porphyry copper deposit, Chile. Economic Geology and the Bulletin of the Society of Economic Geologists, 94, 937–47.CrossRefGoogle Scholar
McKenzie, D. (1978). Some remarks on the development of sedimentary basins. Earth and Planetary Science Letters, 40, 2532.CrossRefGoogle Scholar
McKenzie, D. and Bickle, M. J. (1988). The volume and composition of melt generated by extension of the lithosphere. Journal of Petrology, 29, 625–79.CrossRefGoogle Scholar
McKenzie, D. and Jackson, J. (1986). A block model of distributed deformation by faulting. Journal of the Geological Society of London, 143, 349–53.CrossRefGoogle Scholar
McKenzie, D. and Priestley, K. (2008). The influence of lithospheric thickness variations on continental evolution. Lithos, 102, 111.CrossRefGoogle Scholar
McKenzie, D. P. (1984). A possible mechanism for epeirogenetic uplift. Nature, 307, 616–19.CrossRefGoogle Scholar
McKibben, M. A. and Hardie, L. A. (1997). Ore-forming brines in active continental rifts. In Geochemistry of Hydrothermal Ore Deposits, ed. Barnes, H. L.. John Wiley and Sons, New York, pp. 877935.Google Scholar
McLeod, A. E., Dawers, N. H., and Underhill, J. R. (2000). The propagation and linkage of normal faults: insights from the Strathspey-Brent-Statfjord fault array, northern North Sea. Basin Research, 12, 263–84.CrossRefGoogle Scholar
McLeod, A. E. and Underhill, J. R. (1999). Processes and products of footwall degradation, northern Brent Field, northern North Sea. In Petroleum Geology of Northwest Europe: Proceedings of the 5th Conference, ed. A. J. Fleet and S. A. R. Boldy. American Geological Institute, London, pp. 91–106.CrossRefGoogle Scholar
McLeod, A. E., Underhill, J. R., Davies, S. J., and Dawers, N. H. (2002). The influence of fault array evolution on synrift sedimentation patterns: controls on deposition in the Strathspey-Brent-Statfjord half graben, northern North Sea. American Association of Petroleum Geologists Bulletin, 86, 1061–93.Google Scholar
McNutt, M. (1987). Lithospheric stress and deformation. Reviews of Geophysics, 25, 1245–53.CrossRefGoogle Scholar
Mechie, J. and Brooks, M. (1984). A seismic study of deep geological structure in the Bristol Channel area, SW Britain. Geophysical Journal of the Royal Astronomical Society, 78, 661–89.Google Scholar
Mechie, J., Keller, G. R., Prodehl, C., Gaciri, S., Braile, L. W., Mooney, W. D., Gajewski, D., and Sandmeler, K. J. (1994). Crustal structure beneath the Kenya Rift from axial profile data. Tectonophysics, 236, 179–99.CrossRefGoogle Scholar
Meesters, A. G. C. A. and Dunai, T. J., (2002a). Solving the production-diffusion equation for finite diffusion domains of various shapes: Part I, Implications for low-temperature (U-Th)/He thermochronology. Chemical Geology, 186, 333–44.CrossRefGoogle Scholar
Meesters, A. G. C. A. and Dunai, T. J. (2002b). Erratum: solving the production-diffusion equation for finite diffusion domains of various shapes: Part II, Application to cases with alpha -ejection and nonhomogeneous distribution of the source [modified]. Chemical Geology, 186, 347–63.CrossRefGoogle Scholar
Megson, J. B. (1992). The North Sea chalk play: examples from the Danish Central Graben. In Exploration in Britain: Geological Insights into the Next Decade, ed. Hardman, R. F. P.. Special Publication of the Geological Society of London, 67, 247–82.Google Scholar
Meisling, K. E., Cobbold, P. R., and Mount, V. S. (2001). Segmentation of an obliquely rifted margin, Campos and Santos basins, southeastern Brazil. American Association of Petroleum Geologists Bulletin, 85, 1903–24.Google Scholar
Meissner, R. (1986). The Continental Crust: A Geophysical Approach. Academic Press, San Diego.Google Scholar
Meissner, R. and Strehlau, J. (1982). Limits of stresses in continental crusts and their relation to the depth-frequency distribution of shallow earthquakes. Tectonics, 1, 7389.CrossRefGoogle Scholar
Melinte, M. C. (2005). Oligocene palaeoenvironmental changes in the Romanian Carpathians, revealed by calcareous nannofossils. Studia Geologica Polonica, 124, 341–52.Google Scholar
Mello, M. R. and Maxwell, J. R. (1990). Organic geochemical and biological marker characterization of source rocks and oils derived from lacustrine environments in the Brazilian continental margin. In Lacustrine Basin Exploration – Case Studies and Modern Analogs, ed. Katz, B. J.. American Association of Petroleum Geologists Memoir, 50, 7797.Google Scholar
Mello, M. R., Koutsoukos, E. A. M., and Santos Neto, E. V. (1991). Stratigraphy and depositional settings of Cretaceous organic-rich sequences in the Brazilian marginal basins. American Association of Petroleum Geologists Bulletin, 75, 633–4.Google Scholar
Mello, M. R., Telnaes, N., Gaglianone, P. C., Chicarelli, M. I., Brassell, S. C., and Maxwell, J. R. (1988). Organic geochemical characterisation of depositional palaeoenvironments of source rocks and oils in Brazilian marginal basins. Organic Geochemistry, 43, 3145.CrossRefGoogle Scholar
Mello, U. T. and Bender, A. A. (1988). On isostasy at the equatorial margin of Brazil. Revista Brasileira de Geociencias, 18, 237–46.Google Scholar
Menard, H. W. (1956). Archipelagic aprons (Pacific Ocean). American Association of Petroleum Geologists Bulletin, 40, 2195–210.Google Scholar
Meneses-Rocha, J. J. (2001). Tectonic evolution of the Ixtapa Graben, an example of a strike-slip basin of a southeastern Mexico: implications for regional petroleum systems. In The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems, ed. Bartolini, C., Buffler, R. T., and Cantu-Chapa, A.. American Association of Petroleum Geologists Memoir, 75, 183216.Google Scholar
Menzies, M. A., Baker, J., Chazot, G., and Al’Kadasi, M. (1997b). Evolution of the Red Sea volcanic margin, western Yemen. In Large Igneous Provinces: Continental Oceanic and Planetary Flood Colcanism, ed. Mahoney, J. and Coffin, M. F.. American Geophysical Union Geophysical Monograph 100, pp. 2943.Google Scholar
Menzies, M. A., Gallagher, K., Hurford, A., and Yelland, A. (1997a). Red Sea volcanic and the Gulf of Aden non-volcanic margins, Yemen: denudational histories and margin evolution: Geochimica et Cosmochimica Acta, 61, 2511–28.CrossRefGoogle Scholar
Menzies, M. A., Klemperer, S. L., Ebinger, C. J., and Baker, J. (2002). Characteristics of volcanic rifted margins. In Volcanic Rifted Margins, ed. Menzies, M. A., Klemperer, S. L., Ebinger, C. J., and Baker, J.. GSA Special Paper, 362, 114.CrossRefGoogle Scholar
Meredith, D. J. and Egan, S. S. (2002). The geological and geodynamic evolution of the eastern Black Sea basin: insights from 2-D and 3-D tectonic modelling. Tectonophysics, 350, 157–79.CrossRefGoogle Scholar
Mero, W. E., Thurston, S. P., and Kropschot, R. E. (1992). The Point Arguello field. In Giant Oil and Gas Fields of the Decade 1978–1988, ed. Halbouty, M. T.. American Association of Petroleum Geologists Memoir, 5, 325.Google Scholar
Merrihue, C. and Turner, G. (1966). Potassium-argon dating by activation with fast neutrons. Journal of Geophysical Research, 71, 2852–9.CrossRefGoogle Scholar
Meyerhoff, A. A. (1980a). Geology and petroleum fields in Proterozoic and Lower Cambrian strata, Lena-Tunguska petroleum province, Eastern Siberia, U.S.S.R. In Giant Oil and Gas Fields of the Decade 1968–1978, ed. Halbouty, M. T.. American Association of Petroleum Geologists Memoir, 30, 225–52.Google Scholar
Meyerhoff, A. A. (1980b). Petroleum basins of the Soviet Arctic. Geological Magazine, 117, 101–86.CrossRefGoogle Scholar
Meyerhoff, A. A. (1982a). Hydrocarbon resources in Arctic and sub-Arctic regions. In Arctic Geology and Geophysics, ed. Embry, A. F. and Balkwill, H. R.. Canadian Society of Petroleum Geologists Memoir, 8, 451552.Google Scholar
Meyerhoff, A. A. (1982b). Petroleum basins of the Union of Socialist Soviet Republics and the People’s Republic of China and the politics of petroleum. Petroleum Exploration Society of Australia, Distinguished Lecture Series, Course Notes, p. 341.Google Scholar
Meyers, J. B. (1995). Rifted continental margin architecture off West Africa, as revealed by deep-penetrating multi-channel seismic reflection and potential field data. PhD thesis, University of Miami (Florida), Coral Gables.Google Scholar
Meyers, J. B. and Rosendahl, B. R. (1991). Seismic reflection character of the Cameroon volcanic line: evidence for uplifted oceanic crust. Geology, 19, 1072–6.2.3.CO;2>CrossRefGoogle Scholar
Meyers, J. B., Rosendahl, B. R., Groschel-Becker, H., Austin, J. A. Jr., and Rona, P. A. (1996). Deep penetrating MCS imaging of the rift-to-drift transition, offshore Douala and North Gabon basins, West Africa. Marine and Petroleum Geology, 13, 791835.CrossRefGoogle Scholar
Meyers, J. B., Rosendahl, B. R., Harrison, C. G. A., and Ding, Z.-D. (1998). Deep-imaging seismic and gravity results from the offshore Cameroon volcanic line, and speculation of African hotlines. Tectonophysics, 284, 3163.CrossRefGoogle Scholar
Miall, A. D. (1985). Stratigraphic and structural predictions from a plate-tectonic model of an oblique-slip orogen: the Eureka Sound Formation (Campanian-Oligocene), Northeast Canadian Arctic Islands. Special Publication – Society of Economic Paleontologists and Mineralogists, 37, 361–74.Google Scholar
Miall, A. D. (1996). The Geology of Fluvial Deposits: Sedimentary Facies, Basin Analysis, and Petroleum Geology. Springer-Verlag, Berlin, p. 582.Google Scholar
Michon, L. and Merle, O. (2003). Mode of lithospheric extension: conceptual models from analogue modeling. Tectonics, 22, 1028.CrossRefGoogle Scholar
Middleton, M. F. (1980). A model of intracratonic basin formation, entailing deep crustal metamorphism. Geophysical Journal of the Royal Astronomical Society, 62, 114.CrossRefGoogle Scholar
Middleton, M. F. (1982). Tectonic history from vitrinite reflectance. Geophysical Journal of the Royal Astronomical Society, 68, 121–32.Google Scholar
Miles, P. R., Munschy, M., and Segoufin, J. (1998). Structure and evolution of the Arabian Sea and the eastern Somali basin. Geophysical Journal International, 134, 876–88.CrossRefGoogle Scholar
Miller, D. J. and Christensen, N. I. (1997). Seismic velocities of lower crustal and upper mantle rocks from the slow-spreading Mid-Atlantic Ridge, south of the Kane transform zone (MARK). Proceedings of the Ocean Drilling Program, Scientific Results, 153, 437–54.Google Scholar
Miller, D. S. and Duddy, I. R. (1989). Early Cretaceous uplift and erosion of the northern Appalachian Basin, New York, based on apatite fission track analysis. Earth and Planetary Science Letters, 93, 3549.CrossRefGoogle Scholar
Miller, R. G. (1990). A paleoceanographic approach to Kimmeridge Clay Formation. In Deposition of Organic Facies, ed. Huc, A. Y.. American Association of Petroleum Geologists Studies in Geology, 30, 1326.Google Scholar
Miller, R. G. (1992). The global oil system: the relationship between oil generation, loss, half-life, and the world crude oil resource. American Association of Petroleum Geologists Bulletin, 76, 489500.Google Scholar
Mills, R. A., Teagle, D. A. H., and Tivey, M. K. (1998). Fluid mixing and anhydrite precipitation within the TAG mound. In Proceedings from the Oceanic Drilling Program Scientific Results, ed. P. M. Herzig, S. E. Humphris, D. J. Miller, and R. A. Zierenberg. College Station, Texas, 158, 119–27.Google Scholar
Mills, R. A. and Tivey, M. K. (1999). Sea water entrainment and fluid evolution within the TAG hydrothermal mound: evidence from analyses of anhydrite. In Dynamics of Processes Associated with New Ocean Crust, ed. Cann, J. H. and Laughton, A.. Cambridge University Press, Cambridge, pp. 225–63.Google Scholar
Milner, P. S. and Olsen, T. (1998). Predicted distribution of the Hugin Formation reservoir interval in the Sleipner Ostfield, South Viking Graben: the testing of a three-dimensional sequence stratigraphy model. In Sequence Stratigraphy: Concepts and Applications, ed. Gradstein, F. M., Sandvik, K. O., and Milton, N. J.. Special Publication of the Norwegian Petroleum Society, 8, 337–54.Google Scholar
Milner, S. C., Duncan, A. R., Whittigham, A. M., and Ewart, A. (1995). Trans-Atlantic correlation of eruptive sequences and individual silicic volcanic units within the Parana-Etendeka igneous province. Journal of Volcanology and Geothermal Research, 69(3–4), 137–57.CrossRefGoogle Scholar
Milton, N. J. and Bertram, G. T. (1992). Trap styles: a new classification based on sealing surfaces. American Association of Petroleum Geologists Bulletin, 76, 983–99.Google Scholar
Milton, N. J., Bertram, G. T., and Vann, I. R. (1990). Early Paleogene tectonics and sedimentation in the central North Sea. In Tectonic Events Responsible for Britain’s Oil and Gas Reserves, ed. Hardman, R. F. P. and Brooks, J.. Geological Society of London Special Publications, pp. 339–51.Google Scholar
Min, D.-H., Bullister, J. L., and Weiss, R. F. (2000). Constant ventilation age of thermocline water in the eastern subtropical North Pacific Ocean from chlorofluorocarbon measurements over a 12-year period. Geophysical Research Letters, 27, 3909–12.CrossRefGoogle Scholar
Min, K., Farley, K. A., Renne, P. R., and Marti, K. (2003). Single grain (U-Th)/He ages from phosphates in Acapulco Meteorite and implications for thermal history. Earth and Planetary Science Letters, 209, 323–36.CrossRefGoogle Scholar
Minshull, T. A., Dean, S. M., White, R. S., and Whitmarsh, R. B. (2001). Anomalous melt production after continental breakup in the southern Iberia Abyssal Plain. In Non-Volcanic Rifting of Continental Margins: Evidence from Land and Sea, ed. Wilson, R. C. L., Whitmarsh, R. B., Taylor, B., and Froitzheim, N.. Geological Society of London Special Publications, 187, 537–50.Google Scholar
Minshull, T. A., White, R. S., Mutter, J. C., Buhl, P., Detrick, R. S., Williams, C. A., and Morris, E. (1991). Crustal structure at the Blake Spur fracture zone from expanding spread profiles. Journal of Geophysical Research, 96, 9955–84.CrossRefGoogle Scholar
Mitra, S. (1987). Regional variations in deformation mechanisms and structural styles in the central Appalachian orogenic belt. Geological Society of America Bulletin, 98, 569–90.2.0.CO;2>CrossRefGoogle Scholar
Mittelstaedt, E. and Ito, G. (2005). Plume–ridge interaction, lithospheric stresses, and the origin of near-ridge volcanic lineaments. Geochemistry Geophysics Geosystems, 6.CrossRefGoogle Scholar
Mittelstaedt, E., Ito, G., and Behn, M. D. (2008). Mid-ocean ridge jumps associated with hotspot magmatism. Earth and Planetary Science Letters, 266, 256–70.CrossRefGoogle Scholar
MLWG (1997). MONA LISA: deep seismic investigations of the lithosphere in the southeastern North Sea. Tectonophysics, 269, 119.CrossRefGoogle Scholar
Mocanu, V. I. and Rădulescu, F. (1994). Geophysical features of the Romanian territory. Romanian Journal of Tectonics and Regional Geology, 75, 1736.Google Scholar
Mohriak, W. U., Bassetto, M., and Vieira, I. S. (1998). Crustal architecture and tectonic evolution of the Sergipe-Alagoas and Jacuipe basins, offshore northeastern Brazil. Tectonophysics, 288, 199220.CrossRefGoogle Scholar
Mohriak, W. U., Nemčok, M., and Enciso, G. (2008). South Atlantic divergent margin evolution: rift-border uplift and salt tectonics in the basins of SE Brazil. In West Gondwana: Pre-Cenozoic Correlations across the South Atlantic Region, ed. Pankhurst, R. J., Trouw, R. A. J., de Brito Neves, B. B., and De Witt, M. J.. Geological Society of London Special Publications, 294, 365–98.Google Scholar
Mohriak, W. U., Rosendahl, B. R., Turner, J. P., and Valente, S. C. (2002). Crustal architecture of South Atlantic volcanic margins. In Volcanic Rift Margins, ed. Menzies, M. A., Klemperer, S. L., Ebinger, C., and Baker, J.. Geological Society of America Special Paper, 362, pp. 159202.Google Scholar
Moldowan, J. M., Albrecht, P., and Philp, R. P. (1993). Biological Markers in Sediments and Petroleum. Prentice Hall, Englewood Cliffs, New Jersey.Google Scholar
Molina, P. (1988). Correlation geologique Afrique-Amerique du Sud et provinces uraniferes. African-South American geologic correlations and uraniferous provinces. Journal of African Earth Sciences, 7, 489–97.Google Scholar
Mollema, P. N. and Antonellini, M. A. (1996). Compaction bands: a structural analog for anti-mode I cracks in aeolian sandstone. Tectonophysics, 267, 209–28.CrossRefGoogle Scholar
Molnar, et al. (2005). In Paleogeography and maturation from Congo Span PSDM data, ed. S. G. Henry, A. Danforth, A. G. Requejo, G. F. Schiefelbein, and S. Venkatraman. The 4th HGS/PESGB International Conference on African E7P, Houston Program, 3.Google Scholar
Momper, J. A. (1978). Oil migration limitations suggested by geological and geochemical considerations. American Association of Petroleum Geologists Bulletin Continuing Education Course Notes, 8, 160.Google Scholar
Mongelli, F. (1985). Heat flow along the Italian segment of the EGT. Terra cognita, 5, 385.Google Scholar
Mongelli, F., Loddo, M., and Tramacere, A. (1982). Thermal conductivity, diffusivity and specific heat variation of some Travale field (Tuscany) rocks versus temperature. Tectonophysics, 83, 3343.CrossRefGoogle Scholar
Montadert, L., de Charpel, O., Roberts, D., Guennoc, P., and Sibuet, J. C. (1979). Northeast Atlantic passive continental margins; rifting and subsidence processes. In Deep Drilling Results in the Atlantic Ocean: Continental Margins and Paleoenvironment, ed. Talwani, M. H. and Ryan, W. B. F.. Maurice Ewing Series, Harriman, New York, pp. 154–86.Google Scholar
Montañez, I. P. (1994). Late diagenetic dolomitization of Lower Ordovician, Upper Knox carbonates: a record of the hydrodynamic evolution of the Southern Appalachian Basin. American Association of Petroleum Geologists Bulletin, 78, 1210–39.Google Scholar
Montenat, C., Guery, F., and Jamet, M. (1988). Mesozoic evolution of the Lusitanian Basin: comparison with the adjacent margin. Proceedings of the Ocean Drilling Program, Scientific Results, 103, 757–75.Google Scholar
Moore, C. H. (1989). Carbonate Diagenesis and Porosity. Elsevier, Amsterdam.Google Scholar
Moore, D. G. and Curray, J. R. (1982). Geologic and tectonic history of the Gulf of California. Initial Reports of the Deep Sea Drilling Project, 64, 1279–94.CrossRefGoogle Scholar
Moore, D. M. and Reynolds, R. C. Jr. (1997). X-ray Diffraction and the Identification and Analysis of Clay Minerals. Oxford University Press, Oxford, United Kingdom.Google Scholar
Moore, J. C., Orange, D. L., and Kulm, L. D. (1990). Interrelationship of fluid venting and structural evolution, Oregon margin. Journal of Geophysical Research, 95, 8795–808.CrossRefGoogle Scholar
Moore, J. C. and Vrolijk, P. (1992). Fluids in accretionary prisms. Reviews of Geophysics, 30, 113–35.Google Scholar
Moore, J. G., Normark, W. R., and Holcomb, R. T. (1994). Giant Hawaiian landslides. Annual Review of Earth and Planetary Sciences, 22, 119–44.CrossRefGoogle Scholar
Moore, J. N. and Adams, M. C. (1988). Evolution of the thermal cap in two wells from the Salton Sea geothermal system, California. Geothermics, 17, 695710.CrossRefGoogle Scholar
Moore, M. E., Gleadow, A. J. W., and Lowering, J. F. (1986). Thermal evolution of rifted continental margins: new evidence from fission tracks in basement apatites from southeastern Australia. Earth and Planetary Science Letters, 78, 255–70.CrossRefGoogle Scholar
Mordecai, M. and Morris, L. H. (1971). An investigation into the changes of permeability occurring in a sandstone when failed under triaxial stress condition. In Dynamic Rock Mechanics, ed. D. B. Clark. Proceedings of the 12th Symposium of Rock Mechanics, American Institute of Mining, Metallurgical and Petroleum Engineers, pp. 221–40.Google Scholar
Morency, C., Doin, M.-P., and Dumoulin, C. (2002). Convective destabilization of thickened continental lithosphere. Earth and Planetary Science Letters, 202, 303–20.CrossRefGoogle Scholar
Morgan, J. K. and Clague, D. A. (2003). Volcanic spreading on Mauna Loa volcano, Hawaii: evidence from accretion, alteration, and exhumation of volcanoclastic sediments. Geology, 31, 411–14.2.0.CO;2>CrossRefGoogle Scholar
Morgan, P. (1984). The thermal structure and thermal evolution of the continental lithosphere. In Structure and Evolution of the Continental Lithosphere, ed. Pollack, H. N. and Murthy, V. R.. Physics and Chemistry of the Earth, 15, 107–93.Google Scholar
Morgan, P., Boulos, F. K., Hennin, S. F., El-Sherif, A. A., El-Sayed, A. A., Basta, N. Z., and Melek, Y. S. (1985). Heat flow in eastern Egypt: the thermal signature of a continental breakup. Journal of Geodynamics, 4, 107–31.CrossRefGoogle Scholar
Morgan, P., Harder, V., Swanberg, C. A., and Daggett, P. H. (1981). A groundwater convection model for the Rio Grande rift geothermal resources. Geothermal Research Council Transactions, 5, 193–6.Google Scholar
Morgan, W. J. (1978). Rodriguez, Darwin, Amsterdam, …, a second type of hotspot island. Journal of Geophysical Research, 83, 5355–60.CrossRefGoogle Scholar
Morin, R. and Silva, A. J. (1984). The effects of high pressure and high temperature on physical properties of ocean sediments. Journal of Geophysical Research, 89, 511–26.CrossRefGoogle Scholar
Morley, C. K. (1999). Marked along-strike variations in dip of normal faults: the Lokichar Fault, N. Kenya Rift: a possible cause for metamorphic core complexes. Journal of Structural Geology, 21(5), 479–92.CrossRefGoogle Scholar
Morley, C. K., Nelson, R. A., Patton, T. L., and Munn, S. G. (1990). Transfer zones in the East African Rift system and their relevance to hydrocarbon exploration in rifts. American Association of Petroleum Geologists Bulletin, 74, 1234–53.Google Scholar
Morley, C. K. and Wescott, W. A. (1999). Sedimentary environments and geometry of sedimentary bodies determined from subsurface studies in East Africa. University of Brunei, Department of Petroleum Geoscience, Darusalam, Brunei.CrossRefGoogle Scholar
Morley, C. K., Wescott, W. A., Stone, D. M., Harper, R. M., Wigger, S. T., and Karanja, F. M. (1992). Tectonic evolution of the northern Kenyan Rift. Journal of the Geological Society of London, 149, 333–48.CrossRefGoogle Scholar
Morley, C. K., Wonganan, N., Sankumarn, N., Hoon, T. B., Alief, A., and Simmons, M. (2001). Late Oligocene–recent stress evolution in rift basins of northern and central Thailand: implications for escape tectonics. Tectonophysics, 334, 115–50.CrossRefGoogle Scholar
Morris, J. E., Hampson, G. J., and Johnson, H. D. (2006). A sequence stratigraphic model for an intensely bioturbated shallow-marine sandstone: the Bridport Sand Formation, Wessex Basin, UK. Sedimentology, 53, 1229–63.CrossRefGoogle Scholar
Morrow, C. and Byerlee, J. (1988). Permeability of rock samples from Cajon Pass, California. Geophysical Research Letters, 15, 1033–6.CrossRefGoogle Scholar
Morse, S. A. (1986). Thermal structure of crystallizing magma with two-phase convection. Geological Magazine, 123, 205–14.CrossRefGoogle Scholar
Morton, A. C. and Taylor, P. N. (1987). Lead isotope evidence for the structure of the Rockall dipping-reflector passive margin. Nature, 326, 381–3.CrossRefGoogle Scholar
Moser, F. and Frisch, W. (1996). Tertiary deformation in the Southern Carpathians – structural analysis of a brittle deformation. TSK 6 Symposium Erweitere Kurzfassungen, Vienna, Facultas-Universitatsverlag, pp. 280–2.Google Scholar
Moss, B., Barson, D., Rakhit, K., Dennis, H., and Swarbrick, R., 2002. Formation pore pressures and formation waters. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. Geological Society of London, pp. 317–29.Google Scholar
Mottana, A., Nicoletti, M., Petrucciani, C., Liborio, G., De Capitani, L., and Bocchio, R. (1985). Pre-Alpine and Alpine evolution of the South-alpine basement of the Southern Alps. Geologische Rundschau, 74, 353–66.CrossRefGoogle Scholar
Moulin, M., Aslanian, D., Olivet, J.-L., Contrucci, I., Matias, L., Geli, L., Klingelhoefer, F., Nouze, H., Rehault, J.-P., and Unternehr, P. (2005). Geological constraints on the evolution of the Angolan margin based on reflection and refraction seismic data (ZaiAngo Project). Geophysical Journal International, 162, 793810.CrossRefGoogle Scholar
Moustafa, A. R. (1992). Structural setting of the Sidri-Feiran area, eastern side of the Suez Rift. Earth Science Series, 6, 4454.Google Scholar
Moustafa, A. R. (1996). Internal structure and deformation of an accommodation zone in the northern part of the Suez Rift. Journal of Structural Geology, 18, 93107.CrossRefGoogle Scholar
Moustafa, A. R. (1997). Controls on the development and evolution of transfer zones: the influence of basement structure and sedimentary thickness in the Suez Rift and Red Sea. Journal of Structural Geology, 19, 755–68.CrossRefGoogle Scholar
Moustafa, A. R. (2002). Controls on the geometry of transfer zones in the Suez Rift and Northwest Red Sea: implications for the structural geometry of rift systems. American Association of Petroleum Geologists Bulletin, 86, 9791002.Google Scholar
Mueller, S. (1983). The EGT project. EOS Transactions, 64, 458.CrossRefGoogle Scholar
Muentener, O., Desmurs, L., Pettke, T., and Schaltegger, U. (2002). Melting and melt/rock reaction in extending mantle lithosphere; trace element and isotopic constraints from passive margin peridotites. Geochimica et Cosmochimica Acta, 66, 526.Google Scholar
Muentener, O. and Hermann, J. (2001). The role of lower crust and continental upper mantle during formation of non-volcanic passive margins: evidence from the Alps. Geological Society of London Special Publications, 187, 267–88.Google Scholar
Muentener, O. and Piccardo, G. B. (2003). Melt migration in ophiolitic peridotites: the message from Alpine-Apennine peridotites and implications for embryonic ocean basins. Geological Society of Special Publications, 218, 6989.CrossRefGoogle Scholar
Muffler, L. J. P. and White, D. E. (1968). Origin of CO2 in the Salton Sea geothermal system, southeastern California, U.S.A. Proceedings of International Geological Congress, 97, 185–94.Google Scholar
Mukhopadhyay, P. K. (1994).Vitrinite reflectance as a maturity parameter: petrographic and molecular characterization and its applications to basin modeling. In Vitrinite Reflectance as a Maturity Parameter: Applications and Limitations, ed. Mukhopadhyay, P. K. and Dow, W. G.. American Chemical Society Symposium Series, Washington, DC, 570, 124.CrossRefGoogle Scholar
Mukhopadhyay, P. K. (2002). Evaluation of petroleum systems for five dummy wells from various seismic sections of the Scotian slope, offshore Nova Scotia, based on one-dimensional numerical modelling and using geochemical concepts. Confidential Internal Report, Canada-Nova Scotia Offshore Petroleum Board.Google Scholar
Muller, M. R., Robinson, C. J., Minshull, T. A., White, R. S., and Bickle, M. J. (1997). Thin crust beneath Ocean Drilling Program Borehole 735B at the Southwest Indian Ridge? Earth and Planetary Science Letters, 148, 93107.CrossRefGoogle Scholar
Müller, R. D., Roest, W. R., Royer, J.-Y., Gahagan, L. M., and Sclater, J. G. (1993). A digital age map of the ocean floor. SIO Reference Series 93–30. Scripps Institution of Oceanography, San Diego. http://www.ngdc.noaa.gov/mgg/fliers/96mgg04.html.Google Scholar
Müller, R. D., Roest, W. R., Royer, J. Y., Gahagan, L. M., and Sclater, J. G. (1997). Digital isochrons of the world’s ocean floor. Journal of Geophysical Research, 102, 3211–14.CrossRefGoogle Scholar
Mutter, J. C., Buck, W. R., and Zehnder, C. M. (1988). Convective partial melting. 1. A model for the formation of thick basaltic sequences during the initiation of spreading. Journal of Geophysical Research, 93, 1031–48.CrossRefGoogle Scholar
Mutter, J., Talwani, M., and Stofa, P. L. (1982). Origin of seaward-dipping reflectors in oceanic crust off the Norwegian margin by “subaerial sea-floor spreading.” Geology, 10, 353–7.2.0.CO;2>CrossRefGoogle Scholar
Nadeau, R. M. and McEvilly, T. V. (1997). Seismological studies at Parkfield: V, Characteristic microearthquake sequences as fault-zone drilling targets. Bulletin of the Seismological Society of America, 87, 1463–72.CrossRefGoogle Scholar
Nadin, P. A. and Kusznir, N. J. (1995). Palaeocene uplift and Eocene subsidence in the northern North Sea basin from 2D forward and reverse stratigraphic modelling. Journal of the Geological Society of London, 152, 833–48.CrossRefGoogle Scholar
Nadin, P. A., Kusznir, N. J., and Cheadle, M. J. (1997). Early Tertiary plume uplift of the North Sea and Faeroe-Shetland basins. Earth and Planetary Science Letters, 148, 109–27.CrossRefGoogle Scholar
Nadin, P. A., Kusznir, N. J., and Toth, J. (1995). Transient regional uplift in the early Tertiary of the northern North Sea and the development of the Iceland Plume. Journal of the Geological Society of London, 152, 953–8.CrossRefGoogle Scholar
Naeser, N. D., Naeser, C. W., and McCulloh, T. H. (1989). The application of fission-track dating to depositional and thermal history of rocks in sedimentary basins. In Thermal History of Sedimentary Basins, ed. Naeser, N. D. and McCulloh, T. H.. Springer-Verlag, New York, pp. 157–80.CrossRefGoogle Scholar
Naylor, M. A. and Spring, L. Y. (2002). Exploration strategy development and performance management: a portfolio-based approach. The Leading Edge, 21, 159–67.CrossRefGoogle Scholar
Naylor, P. H., Bell, B. R., Jolley, D. W., Durnall, P., and Fredsted, R. (1999). Palaeogene magmatism in the Faeroe-Shetland Basin: influences on uplift history and sedimentation. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, Geological Society of London, pp. 545–58.Google Scholar
Neculae, P. and Stanescu, V. (2001). Oil Fields in the External Carpathian Flysch. Vergiliu, Bucharest, p. 209.Google Scholar
Nederlof, M. H. and Mohler, H. P. (1981). Quantitative investigation of trapping effect of unfaulted caprock. American Association of Petroleum Geologists Bulletin, 65, 964.Google Scholar
Nelson, R. A. (2001). Geologic Analysis of Naturally Fractured Reservoirs. Gulf Professional Publishing, Boston.Google Scholar
Nemčok, M. (2001). Preliminary Petroleum Systems and Source Rock Risk Evaluation: Santos Basin, Brazil – Phase II. EGI Report, 01-00059-5000-50500773.Google Scholar
Nemčok, M. (2004). Niko-1 well project for Frontera Resources. EGI Report, 01-00059-5000-50501264.Google Scholar
Nemčok, M. (2005). Combined structural/sedimentological research in Block 12 for Frontera Resources during year 2005. EGI Report, 01-00059-5000-50501492.Google Scholar
Nemčok, M. (2012). Structural aspects of tight shales. In Evolution of the Mental Picture of Tight Shales, ed. McLennan, J.. Energy and Geoscience Institute, Salt Lake City, pp. 4158.Google Scholar
Nemčok, M. and Allen, R. (2001). Fault propagation timing in the northern portion of the Gulf of Guinea, Africa. EGI report, 50500854, p. 51.Google Scholar
Nemčok, M. and Gayer, R. A. (1996). Modelling palaeostress magnitude and age in extensional basins: a case study from the Mesozoic Bristol Channel Basin, UK. Journal of Structural Geology, 18, 1301–14.CrossRefGoogle Scholar
Nemčok, M., Gayer, R. A., and Miliorizos, M. (1995). Structural analysis of the inverted Bristol Channel Basin: implications for the geometry and timing of the fracture porosity. Basin inversion. Geological Society of London Special Publications, 82, 355–92.Google Scholar
Nemčok, M., Henk, A., Allen, R., Sikora, P. J., and Stuart, C. (2012b). Continental break-up along strike-slip fault zones: observations from Equatorial Atlantic. In Conjugate Divergent Margins, ed. Mohriak, W. U., Danforth, A., Post, P. J., Brown, D. E., Tari, G. M., Nemčok, M., and Sinha, S. T.. Geological Society of London Special Publications, 369.Google Scholar
Nemčok, M., Henk, A., Gayer, R. A., Vandycke, S., and Hathaway, T. M. (2002). Strike-slip fault bridge fluid pumping mechanism: Insights from field-based palaeostress analysis and numerical modeling. Journal of Structural Geology, 24, 1885–901.CrossRefGoogle Scholar
Nemčok, M. and Kantor, J. (1990). Movement study in the selected area of the Velký Bok Unit (in Slovak). Regionálna Geológia Západných Karpát, 7583.Google Scholar
Nemčok, M., Konecny, P., and Lexa, A. (2000). Calculations of tectonic, magmatic and residual stress in the Stiavnica stratovolcano, Western Carpathians: implications for mineral precipitation paths. Geologica Carpathica, 51, 1936.Google Scholar
Nemčok, M. and Lexa, J. (1990). Evolution of the basin and range structure around Ziar mountain range. Geologicky Sbornik, 41, 229–58.Google Scholar
Nemčok, M., Moore, J. N., Allis, R., and McCulloch, J. (2004a). Fracture development within a stratovolcano: the Karaha-Telaga Bodas geothermal field, Java volcanic arc. In The Initiation, Propagation, and Arrest of Joints and other Fractures: A Field Workshop Dedicated to the Memory of Paul Hancock, ed. Cosgrove, J. and Engelder, T.. Geological Society of London Special Publications, 231, 223–42.Google Scholar
Nemčok, M., Moore, J. N., Christensen, C., Allis, R., Powell, T., Murray, B., and Nash, G. (2006a). Controls on the Karaha-Telaga Bodas geothermal reservoir, Indonesia. Geothermics, 36, 946.CrossRefGoogle Scholar
Nemčok, M., Pogacsás, G., and Pospíšil, L. (2006b). Activity timing of the main tectonic systems in the Carpathian-Pannonian region in relation to the roll-back destruction of the lithosphere. The Carpathians and their foreland: geology and hydrocarbon resources. American Association of Petroleum Geologists Memoir, 84, 743–66.Google Scholar
Nemčok, M., Pospíšil, L., Lexa, J., and Donelick, R. A. (1998). Tertiary subduction and slab break-off model of the Carpathian-Pannonian region. Tectonophysics, 295, 307–40.CrossRefGoogle Scholar
Nemčok, M. and Rosendahl, B. R. (2006). CameroonSpan and GabonSpan interpretation. EGI Report, 01-00059-5000-50501399.Google Scholar
Nemčok, M., Rosendahl, B. R., Segall, M., Silva, C., Stuart, C., Allen, R., Sikora, P., Christensen, C., Francu, J., and Abrams, M. (2004b). Equatorial Atlantic margins basins project. A joint EGI-Industry evaluation of the evolution, development and prospectivity of basins along the eastern and western Equatorial Atlantic continental margins. EGI Report, 01-0059-5000-50500936.Google Scholar
Nemčok, M., Schamel, S., and Gayer, R. (2005a). Thrustbelts: Structural Architecture, Thermal Regimes and Petroleum Systems. Cambridge University Press, Cambridge, United Kingdom, p. 541.CrossRefGoogle Scholar
Nemčok, M., Sheya, C., Welker, C., Smith, S., and Rybár, S. (2010). Crustal types, structural architecture, and plate configurations study of the Reliance West Coast region, Phase 1 project. EGI Report, 01-00059-5000-50501853 for Reliance (confidential).Google Scholar
Nemčok, M., Sinha, S. T., Stuart, C., Welker, C., Choudhuri, M., Sharma, S. P., Misra, A. A., Sinha, N., and Venkatraman, S. (2012c). East Indian margin evolution and crustal architecture: integration of deep reflection seismic interpretation and gravity modeling. In Conjugate Divergent Margins, ed. Mohriak, W. U., Danforth, A., Post, P. J., Brown, D. E., Tari, G. M., Nemčok, M., and Sinha, S. T.. Geological Society of London Special Publications, 369.Google Scholar
Nemčok, M., Stuart, C., Odegard, M., Sheya, C., Welker, C., Dvorakova, V., Meissner, A., Bubniak, I., Bubniak, A., Vangelov, D., and Baraboshkin, E. (2009). Tectonic development of the Black Sea region. Phase 2 of the circum-Black Sea project. EGI Report, 01-00059-5000-50501563.Google Scholar
Nemčok, M., Stuart, C., Moore, J. N., Smith, S., Welker, C., and Sheya, C. (2008). Crustal types, structural architecture, and plate configurations study of the Reliance East Coast region. Phase 1-extension study. EGI report, 01-00059-5000-50501784-2, p. 37.Google Scholar
Nemčok, M., Stuart, C. J., Rosendahl, B. R., Welker, C., Smith, S., Sheya, C., Sinha, S. T., Choudhuri, M., Allen, R., Reeves, C., Sharma, S. P., Venkatraman, S., and Sinha, N. (2012a). Continental break-up mechanism: lessons from intermediate- and fast-extension settings. In Conjugate Divergent Margins, ed. Mohriak, W. U., Danforth, A., Post, P. J., Brown, D. E., Tari, G. M., Nemčok, M., and Sinha, S. T.. Geological Society of London Special Publications, 369.Google Scholar
Nemčok, M., Stuart, C., Segall, M. P., Allen, R. B., Christensen, C., Hermeston, S. A., and Davison, I. (2005b). Structural development of South Morocco: interaction of tectonics and deposition. In Petroleum Systems of Divergent Continental Margin Basins, ed. Post, P. J., Rosen, N., Olson, D. L., Palmes, S. L., Lyons, K. T., and Newton, G. B.. 25th Annual GCSSEPM Foundation Bob F. Perkins Research Conference, Session I, Crustal architecture of divergent margins, Houston, pp. 151–202.Google Scholar
Nemčok, M., Stuart, C., Welker, C., Smith, S., Yalamanchili, R., Srivastava, D. C., Sinha, S., and Choudhuri, M. (2007). Crustal types, structural architecture, and plate configurations study of the Reliance East Coast region, Phase 1 project. EGI Report, 01-00059-5000-50501546 for Reliance (confidential).Google Scholar
Nemčok, M., Vangelov, D., Stuart, C., Francu, J., Christensen, C., Abrams, M., Moore, J., Jones, C., Clausen, S., Sahm, E., and Jones, P. (2004c). Bulgaria Bourgas Block Project Phase 2. EGI Report, 50501139.Google Scholar
Nemec, W. (1990). Aspects of sediment movement on steep delta slopes. In Coarse Grained Deltas, ed. Colella, A. and Prior, D. B.. Special Publication of the International Association of Sedimentologists, pp. 2973.CrossRefGoogle Scholar
Netto, A. S. T. and de Oliveira, J. J. (1985). O preenchimento do rift-valley na Bacia do Reconcavo. The filling of the rift valley in the Reconcavo Basin. Revista Brasileira de Geociencias, 15, 97102.Google Scholar
Neuzil, C. E. (1986). Groundwater flow in low-permeability environments. Water Resources Research, 22, 1163–95.CrossRefGoogle Scholar
Newmark, R. L., Kasameyer, P. W., and Younker, L. W. (1988). Shallow drilling in the Salton Sea region: the thermal anomaly. Journal of Geophysical Research, 93, 13005–23.CrossRefGoogle Scholar
Nilsen, T. H. and McLaughlin, R. J. (1985). Comparison of tectonic framework and depositional patterns of the Hornelen strike-slip basin of Norway and the Ridge and Little Sulphur Creek strike-slip basins of California. Special Publication – Society of Economic Paleontologists and Mineralogists, 37, 79103.Google Scholar
Noguti, I. and Santos, J. F. D. (1972). Zoneamento preliminar por foraminiferos planctonicos do aptiano ao mioceano na plataforma continental do Brasil. Preliminary zoning of planktonic foraminifera of the Aptian to Miocene rocks of the continental shelf of Brazil. Boletim Tecnico da PETROBRAS, 15, 265–83.Google Scholar
Noll, C. A. and Hall, M. (2006). Normal fault growth and its function on the control of sedimentation during basin formation: a case study from field exposures of the Upper Cambrian Owen Conglomerate, West Coast Range, western Tasmania, Australia. American Association of Petroleum Geologists Bulletin, 90, 1609–30.CrossRefGoogle Scholar
Norris, R. J. and Carter, R. M. (1980). Offshore sedimentary basins at the southern end of the Alpine Fault, New Zealand. Special Publication of the International Association of Sedimentologists, 4, 237–65.Google Scholar
Norton, D. and Knight, J. (1977). Transport phenomena in hydrothermal systems: cooling plutons. American Journal of Science, 277, 937–81.Google Scholar
Nuckolls, H. M. and McCulley, B. L. (1987). Origin of saline springs in Cataract Canyon, Utah. Field Symposium – Guidebook of the Four Corners Geological Society, 10, 193200.Google Scholar
Nunn, J. A. (1985). State of stress in the northern Gulf Coast. Geology, 13, 429–32.2.0.CO;2>CrossRefGoogle Scholar
Nunn, J. A. and Deming, D. (1991). Thermal constraints on basin-scale flow systems. In Crustal-Scale Fluid Transport – Magnitude and Mechanisms, ed. Torgensen, T.. Geophysical Research Letters, 18, 967–70.Google Scholar
Nunn, J. A. and Meulbroek, P. (2002). Kilometer-scale upward migration of hydrocarbons in geopressured sediments by buoyancy-driven propagation of methane-filled fractures. American Association of Petroleum Geologists Bulletin, 86, 907–18.Google Scholar
Nunn, J. A. and Sassen, R. (1986). The framework of hydrocarbon generation and migration, Gulf of Mexico continental slope. Gulf Coast Association of Geological Societies Transactions, 35, 257–62.Google Scholar
Obermeier, S. (1996). Use of liquefaction-induced features for paleoseismic analysis – an overview of how seismic liquefaction features can be distinguished from other features and how their regional response distribution and properties of source sediment can be used to infer the location and strength of Holocene paleo-earthquakes. Engineering Geology, 44, 176.CrossRefGoogle Scholar
O’Connell, R. J. and Budiansky, B. (1974). Seismic velocities in dry and saturated cracked solids. Journal of Geophysical Research, 79, 5412–26.Google Scholar
O’Connell, R. J. and Wasserburg, G. J. (1972). Dynamics of submergence and uplift of a sedimentary basin underlain by a phase-change boundary. Reviews of Geophysics and Space Physics, 10, 335–68.CrossRefGoogle Scholar
Offshore Energy Technical Research (OETR) Association (2011). Play Fairway Analysis. http://www.offshoreenergyresearch.ca/OETR/OETRPlayFairwayProgramMainPage/tabid/402/Default.aspx.Google Scholar
Oh, J., Austin, J. A., Phillips, J. D., Coffin, M. F., and Stoffa, P. L. (1995). Seaward-dipping reflectors offshore the southeastern United States: seismic evidence for extensive volcanism accompanying sequential formation of the Carolina Trough and Blake Plateau basin. Geology, 23, 912.2.3.CO;2>CrossRefGoogle Scholar
O’Hara, I., Hower, J. G., and Rimmer, S. M. (1990). Constraints on the emplacement and uplift history of the Pine Mountain thrust sheet, eastern Kentucky: evidence from coal rank trends. Journal of Geology, 98, 4351.CrossRefGoogle Scholar
Ohlmacher, G. C. and Aydin, A. (1997). Mechanics of vein, fault and solution surface formation in the Appalachian Valley and Ridge, northeastern Tennessee, U.S.A.: implications for fault friction, state of stress and fluid pressure. Journal of Structural Geology, 19, 927–44.CrossRefGoogle Scholar
Ojeda, H. A. O. (1982). Structural framework, stratigraphy, and evolution of Brazilian marginal basins. American Association of Petroleum Geologists Bulletin, 66, 732–49.Google Scholar
Oliver, J. (1986). Fluids expelled tectonically from orogenic belts: their role in hydrocarbon migration and other geologic phenomena. Geology, 14, 99102.2.0.CO;2>CrossRefGoogle Scholar
Olsen, K. H. and Morgan, P. (1995). Introduction: progress in understanding continental rifts. Developments in Geotectonics, 25, 326.CrossRefGoogle Scholar
Olsen, P. E. (1990). Tectonic, climatic and biotic modulation of lacustrine ecosystems – examples from Newark Supergroup of eastern North Americas. In Lacustrine Basin Exploration – Case Studies and Modern Analogs, ed. Katz, B. J.. American Association of Petroleum Geologists Memoir, 50, 209–24.Google Scholar
Olsen, P. E. and Schlische, R. W. (1990). Transtensional arm of the early Mesozoic Fundy rift basin: penecontemporaneous faulting and sedimentation. Geology, 18, 695–8.2.3.CO;2>CrossRefGoogle Scholar
Olsen, P. E., Schlische, R.W., and Gore, P. J. W. (1990). Tectonic, depositional, and paleoecological history of early Mesozoic rift basins, eastern North America. Geology, 18, 174.Google Scholar
Ord, A. and Hobbs, B. E. (1989). The strength of the continental crust, detachment zones and the development of plastic instabilities. Tectonophysics, 158, 269–89.CrossRefGoogle Scholar
Orr, W. L. (1986). Kerogen/asphaltene/sulfur relationships in sulfur-rich Monterey oils. Organic Geochemistry, 10, 499506.CrossRefGoogle Scholar
Orr, W. L. (1993). Guidelines for Type II-S Kerogens in basin modeling: organic sulfur content of kerogens and crude oils. American Association of Petroleum Geologists Bulletin, 77, 1652.Google Scholar
Ortoleva, P., Merino, E., Moore, C., and Chadam, J. (1987). Geochemical self-organization, I. Reaction-transport feedbacks and modeling approach. American Journal of Science, 287, 9791007.CrossRefGoogle Scholar
Osborne, M. J. and Swarbrick, R. E. (1997). Mechanisms for generating overpressures in sedimentary basins: a reevaluation. American Association of Petroleum Geologists Bulletin, 81, 1023–41.Google Scholar
Ostermeier, R. M. (2001). Compaction effects on porosity ad permeability: deepwater Gulf of Mexico turbidites. Journal of Petroleum Technology, 53, 6874.CrossRefGoogle Scholar
O’Sullivan, P. B., Hanks, C. I., Wallace, W. K., and Green, P. F. (1995). Multiple episodes of Cenozoic denudation in the northeastern Brooks Range: fission-track data from the Okpilak batholith, Alaska. Canadian Journal of Earth Sciences, 32, 1106–18.Google Scholar
Ottesen, E. O., Knipe, R. J., Olsen, T. S., Fisher, Q. J., and Jones, G. (1998). Fault controlled communication in the Sleipner vest Field, Norwegian Continental Shelf: detailed, quantitative input for reservoir simulation and well planning. In Faulting, Fault Sealing and Fluid Flow in Hydrocarbon Reservoirs, ed. Jones, G., Fisher, Q. L., and Knipe, R. J.. Geological Society of London Special Publications, 147, 269–82.Google Scholar
Pajdušák, P., Plomerová, J., and Babuška, V. (1988). Lithosphere deep structure of the West Carpathians determined on the basis of P – residue. In Proceedings of the Conference on Deep Structure Investigation of Czechoslovakia. VEDA, Bratislava, Slovak Academy of Sciences, pp. 151–8.Google Scholar
Palciauskas, V. V. (1986). Models for thermal conductivity and permeability in normally compacting basins. In Thermal Modeling of Sedimentary Basins, ed. Burrus, J.. Technip, Paris, pp. 323–36.Google Scholar
Palciauskas, V. V. and Domenico, P. A. (1982). Characterization of drained and undrained response of thermally loaded repository rocks. Water Resources Research, 18, 281–90.CrossRefGoogle Scholar
P’An, C.-H. (1982). Petroleum in basement rocks. American Association of Petroleum Geologists Bulletin, 66, 1597–643.Google Scholar
Pandey, O. P., Agarwal, P. K., and Negi, J. G. (1995). Lithospheric structure beneath Laxmi Ridge and late Cretaceous geodynamic events. Geo-Marine Letters, 15, 8591.CrossRefGoogle Scholar
Parker, R. H. (1991). The Ivanhoe and Rob Roy Fields, Blocks 15/21a-b, UK North Sea. Memoir of the Geological Society of London, 14, 331–8.CrossRefGoogle Scholar
Parmentier, E. M. (1987). Dynamic topography in rift zones: implications for lithospheric heating. Royal Society of London Philosophical Transactions, 321, 23–5.Google Scholar
Parnell, J. (1987). Secondary porosity in hydrocarbon-bearing transgressive sandstones on an unstable lower Paleozoic continental shelf, Welsh Borderland. In Diagenesis of Sedimentary Sequences, ed. Marshall, J. D.. Queen’s University of Belfast, pp. 297312.Google Scholar
Parnell, J. (2002). Diagenesis and fluid flow in response to uplift and exhumation. In Diagenesis and Fluid Flow in Response to Uplift and Exhumation, Dore, A. G., Cartwright, J. A., Stoker, M. S. T., Jonathan, N. J., and White, P.. Geological Society of London Special Publications, pp. 433–46.CrossRefGoogle Scholar
Parris, T. M., Burruss, R. C., and O’Sullivan, P. B. (2003). Deformation and the timing of gas generation and migration in the eastern Brooks Range foothills, Arctic National Wildlife Refuge, Alaska. American Association of Petroleum Geologists Bulletin, 87, 1823–46.CrossRefGoogle Scholar
Parrish, J. T. (1982). Upwelling and petroleum source beds with reference to the Paleozoic. American Association of Petroleum Geologists Bulletin, 66, 750–74.Google Scholar
Parrish, R. R. (1983). Cenozoic thermal evolution and tectonics of the Coast Mountains of British Columbia, 1. Fission-track dating, apparent uplift rates, and patterns of uplift. Tectonics, 2, 601–31.CrossRefGoogle Scholar
Parry, W. T. and Bruhn, R. L. (1987). Fluid inclusion evidence for minimum 11 km vertical offset on the Wasatch fault, Utah. Geology, 15, 6770.2.0.CO;2>CrossRefGoogle Scholar
Parry, W. T. and Bruhn, R. L. (1990). Fluid pressure transients on seismogenic normal faults. Tectonophysics, 179, 335–44.CrossRefGoogle Scholar
Parry, W. T., Chan, M. A., and Beitler, B. (2004). Chemical bleaching indicates episodes of fluid flow in deformation bands in sandstone. American Association of Petroleum Geologists Bulletin, 88, 117.CrossRefGoogle Scholar
Parry, W. T., Wilson, P., and Bruhn, R. L. (1988). Pore fluid chemistry and chemical reactions on the Wasatch normal fault, Utah. Geochimica et Cosmochimica Acta, 52, 2053–63.CrossRefGoogle Scholar
Parsons, B. and Sclater, J. G. (1977). An analysis of the variation of ocean floor bathymetry and heat flow with age. Journal of Geophysical Research, 82, 803–27.CrossRefGoogle Scholar
Partington, M. A., Copestake, P., Mitchener, B. C., and Underhill, J. R. (1993). Biostratigraphic calibration of genetic stratigraphic sequences in the Jurassic – lowermost Cretaceous (Hettangian – Ryazanian) of the North Sea and adjacent areas. In Petroleum Geology of the Northwest Europe, ed. J. R. Parker. Proceedings of the 4th Conference, Geological Society of London, pp. 371–86.CrossRefGoogle Scholar
Pasquale, V. (2006). Effects of crustal heat source redistribution on the strength envelopes. Geophysical Research Abstracts, 8.Google Scholar
Patton, T. L. (1984). Surface studies of normal-fault geometries in the pre-Miocene stratigraphy of west central Sinai Peninsula. Egyptian General Petroleum Corporation, Sixth Exploration Seminar, pp. 437–52.Google Scholar
Patton, T. L., Moustafa, A. R., Nelson, R. A., and Abdine, A. S. (1994). Tectonic evolution and structural setting of the Suez Rift. American Association of Petroleum Geologists Memoir, 59, 955.Google Scholar
Peacock, D. C. P. and Sanderson, D. J. (1991). Displacements, segment linkage and relay ramps in normal fault zones. Journal of Structural Geology, 13, 721–33.Google Scholar
Peakall, J. (1998). Axial river evolution in response to half-graben faulting: Carson River, Nevada, U.S.A. Journal of Sedimentary Research, 68, 788–99.CrossRefGoogle Scholar
Peate, D. (1997). The Paraná-Etendeka Province. In Large Igneous Provinces: Continental Oceanic and Planetary Flood Volcanism, ed. J. Mahoney and M. F. Coffin. American Geophysical Union Geophysical Monograph 100, pp. 217–45.Google Scholar
Pecova, J., Petr, V., Praus, O., Babuška, V., and Plancar, J. (1979). Internal electrical conductivity distribution on Czechoslovak territory. In Geodynamical Investigations in Czechoslovakia, Bratislava, ed. Vanek, J.. Veda-SAV, pp. 119–27.Google Scholar
Pécskay, Z., Lexa, J., Szakács, A., Balogh, K., Seghedi, I., Konecný, V., Kovács, M., Márton, E., Kaliciak, M., Széky-Fux, V., Póka, T., Gyarmati, P., Edelstein, O., Rosu, E., and Zec, B. (1995). Space and time distribution of Neogene-Quaternary volcanism in the Carpatho-Pannonian region. Acta Vulcanologica, 7, 1528.Google Scholar
Peddy, C., Pinet, B., Masson, D., Scrutton, R., Sibuet, J.-C., Warner, M. R., Leford, J.-P., and Shroeder, I. J. (1989). Crustal structure of the Goban Spur continental margin, Northeast Atlantic, from deep seismic reflection profiling. Journal of the Geological Society of London, 146, 427–37.CrossRefGoogle Scholar
Pedersen, T. F. and Calvert, S. E. (1990). Anoxia vs. productivity: What controls the formation of organic-carbon-rich sediments and sedimentary rocks? American Association of Petroleum Geologists Bulletin, 74, 454–66.Google Scholar
Pedersen, T. and Ro, H. E. (1992). Finite duration extension and decompression melting. Earth and Planetary Science Letters, 113, 1522.CrossRefGoogle Scholar
Pedersen, T. and Skogseid, J. (1989). Vøring Plateau volcanic margin: extension, melting and uplift. Proceedings of the Ocean Drilling Program, Scientific Results, 104, 985–91.Google Scholar
Peel, F. J., Scott, E., and Evans, N. (2001). Abnormal processes as major influences on deepwater seafloor morphology and deposition. Annual Meeting Expanded Abstracts, American Association of Petroleum Geologists, p. 155.Google Scholar
Peltier, W. R., Jarvis, G. T., Forte, A. M., and Solheim, L. P. (1989). The radial structure of the mantle general circulation. The Fluid Mechanics of Astrophysics and Geophysics, 4, 765815.Google Scholar
Peng, Z. and Ben-Zion, Y. (2005). Temporal changes of shallow seismic velocity around the Daradere-Duzce branch of the north Anatolian fault and strong ground motion. Pure and Applied Geophysics, 163, 567–99.Google Scholar
Pennington, J. J. (1975). The geology of the Argyll field. In Petroleum and the Continental Shelf of North-west Europe, ed. Woodland, A. W.. Applied Science Publishers, Barking, pp. 285–91.Google Scholar
PennWell (1982). Oil and Gas Fields of the United States. PennWell Publishing Company, Tulsa, Scale 1: 3 380 952.Google Scholar
Pepper, A. S. (1991). Estimating the petroleum expulsion behavior of source rocks: a novel quantitative approach. In Petroleum Migration, ed. England, W. A. and Fleet, A. J.. Special Publication of Geological Society of London, 59, 931.Google Scholar
Pepper, A. S. and Corvi, P. J. (1995). Simple kinetic models of petroleum formation. Part I: Oil and gas generation from kerogen. Marine and Petroleum Geology, 12, 291319.CrossRefGoogle Scholar
Pepper, A. S. and Dodd, T. A. (1995). Simple kinetic models of petroleum formation. Part II: Oil-gas cracking. Marine and Petroleum Geology, 12, 321–40.Google Scholar
Pe-Piper, G. and Reynolds, P. H. (2000). Early Mesozoic alkaline mafic dykes, southwestern Nova Scotia. The Canadian Mineralogist, 38, 217–32.CrossRefGoogle Scholar
Pereira, E. B., Hamza, V. M., Furtado, V. V., and Adams, J. A. (1986). U, Th, and K content, heat production and thermal conductivity of Sao Paulo, Brazil, continental shelf sediments: a reconnaissance work. Chemical Geology, 58, 217–26.CrossRefGoogle Scholar
Pereira, M. J. (1990). Revised stratigraphic column of the Santos Basin (in Portuguese). Parameters controlling porosity and permeability in clastic reservoirs of the Merluza deep field, Santos Basin, Brazil. Boletim de Geociencias da PETROBRAS, 4, 451–66.Google Scholar
Pereira, M. J. and Macedo, J. M. (1990). A Bacia de Santos; perspectivas de uma nova provincia petrolifera na plataforma continental Sudeste Brasileira. Boletim de Geociencias da PETROBRAS, 4, 311.Google Scholar
Perez, R. J. and Boles, J. R. (2004). Mineralization, fluid flow, and sealing properties associated with an active thrust fault: San Joaquin Basin, California. American Association of Petroleum Geologists Bulletin, 88, 1295–314.CrossRefGoogle Scholar
Pérez-Gussinyé, M., Ranero, C. R., Reston, T. J., and Sawyer, D. (2003). Mechanisms of extension at nonvolcanic margins: evidence from the Galicia interior basin, west of Iberia. Journal of Geophysical Research, 108.CrossRefGoogle Scholar
Pérez-Gussinyé, M. and Reston, T. J. (2001). Rheological evolution during extension at nonvolcanic rifted margins: onset of serpentinization and development of detachments leading to continental breakup. Journal of Geophysical Research, 106, 3961–76.CrossRefGoogle Scholar
Perón-Pinvidic, G. and Manatschal, G. (2010). From microcontinents to extensional allochthons: witnesses of how continents rift and break apart? Petroleum Geoscience, 16, 189–97.CrossRefGoogle Scholar
Perón-Pinvidic, G., Manatschal, G., Minshull, T. A., and Sawyer, D. S. (2007). Tectonosedimentary evolution of the deep Iberia-Newfoundland margins: evidence for a complex breakup history. Tectonics, 26.CrossRefGoogle Scholar
Perrodon, A. (1980). Géodynamique pétrolière. Genése et répartition des gisements d’hydrocarbures. Paris, Masson Elf-Aquitaine Pierce.Google Scholar
Perrodon, A. (1995). Petroleum systems and global tectonics. Journal of Petroleum Geology, 18, 471–6.CrossRefGoogle Scholar
Perrot, K., Geoffroy, L., and Dyment, J. (2003). Magnetic structure of the Greenland Volcanic Passive margin. Abstracts, AGU-EUG-EGS meeting, Nice. April 6–11.Google Scholar
Person, M. (1990). Hydrologic constraints on the thermal evolution of continental rift basins: implications for petroleum maturation. PhD thesis, John Hopkins University, Baltimore, Maryland, p. 271.Google Scholar
Person, M. and Garven, G. (1992). Hydrologic constraints on petroleum generation within continental rift basins: theory and application to the Rhine Graben. American Association of Petroleum Geologists Bulletin, 76, 468–88.Google Scholar
Person, M. and Garven, G. (1994). A sensitivity study of the driving forces on fluid flow during continental-rift basin evolution. Geological Society of America Bulletin, 106, 461–75.2.3.CO;2>CrossRefGoogle Scholar
Peters, K. E. (1986). Guidelines for evaluating petroleum source rock using programmed pyrolysis. American Association of Petroleum Geologists Bulletin, 70, 318–29.Google Scholar
Peters, K. E. and Cassa, M. R. (1994). Applied source rock geochemistry. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 93120.Google Scholar
Peters, K. E., Cunningham, A. E., Walters, C. C., Jigang, J., and Zhaoan, F. (1996). Petroleum systems in the Jiangling-Dangyang area, Jianghan basin, China. Organic Geochemistry, 24, 1035–60.CrossRefGoogle Scholar
Peters, K. E. and Fowler, M. G. (2002). Applications of petroleum geochemistry to exploration and reservoir management. Organic Geochemistry, 33, 536.CrossRefGoogle Scholar
Peters, K. E. and Moldowan, J. M. (1993). The Biomarker Guide: Interpreting Molecular Fossils in Petroleum and Ancient Sediments. Prentice Hall, Englewood Cliffs, New Jersey, p. 363.Google Scholar
Peters, K. E., Moldowan, J. M., Driscole, A. R., and Demaison, G. J. (1989). Origin of Beatrice oil by co-sourcing from Devonian and Middle Jurassic source rocks, Inner Morray Firth, United Kingdom. AAPG Bulletin, 73, 454–71.Google Scholar
Peters, K. E., Pytte, M. H., Elam, T. D., and Sundararaman, P. (1994). Identification of petroleum systems adjacent to the San Andreas fault, California. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 423–36.Google Scholar
Peters, K. E., Ramos, L. S., Zumberge, J. E., Valin, Z. C., Scotese, C. R., and Gautier, D. L. (2007). Circum-Artic petroleum systems identified using decision-tree chemometrics. American Association of Petroleum Geologists Bulletin, 91, 877913.CrossRefGoogle Scholar
Peters, K. E., Snedden, J. W., Sulaeman, A., Sarg, J. F., and Enrico, R. J. (2000). A new geochemical-stratigraphic model for the Mahakam delta and Makassar Slope, Kalimantan, Indonesia. American Association of Petroleum Geologists Bulletin, 84, 1244.Google Scholar
Peters, T. and Stettler, A. (1987). Radiometric age, thermobarometry, and mode of emplacement of the Totalp peridotite in the Eastern Swiss Alps. Schweizerische Mineralogische und Petrographische Mitteilungen = Bulletin Suisse de Mineralogie et Petrographie, 67, 285–94.Google Scholar
Petersen, H. I., Bojesen-Koefoed, J. A., and Thomsen, E. (1995). Cola-derived petroleum in the Middle Jurassic Bryne Formation in the Danish North Sea – a new type of play. In Organic Geochemistry: Developments and Applications to Energy, Climate, Environment and Human History, ed. J. O. Grimalt and C. Dorronsoro. Proceedings of the 17th International Meeting on Organic Geochemistry, San Sebastian, Spain, September 4–8, pp. 473–5.Google Scholar
Petersen, H. I., Rosenberg, P., and Andsbjerg, J. (1996). Organic geochemistry in relation to the depositional environments of Middle Jurassic coal seams, Danish Central Graben, and implications for hydrocarbon generative potential. American Association of Petroleum Geologists Bulletin, 80, 4762.Google Scholar
Peterson, J. A. and Clarke, J. W. (1989). West Siberian oil-gas province. United States Geological Survey, Open File Report, 89–192, 142.Google Scholar
Petit, J. P. (1987). Criteria for the sense of movement on fault surfaces in brittle rocks. Journal of Structural Geology, 9, 597608.CrossRefGoogle Scholar
Philip, H., Rogozhin, E., Cisternas, A., Bosquet, J. C., Borisov, B., and Karakhanian, A. (1992). The Armenian Earthquake of 1988 December 7: Faulting and folding, neotectonics and palaeoseismicity. Geophysical Journal International, 110, 141–58.CrossRefGoogle Scholar
Phillips, O. M. (1991). Flow and Reactions in Permeable Rocks. Cambridge University Press, New York.Google Scholar
Philp, R. P. (1994). Geochemical characteristics of oils derived predominantly from terrigenous source materials. In Coal and Coal-bearing Strata as Oil-prone Source Rocks, ed. Scott, A. C. and Fleet, A. J.. Geological Society of London Special Publications, 77, 7191.Google Scholar
Philp, R. P. and Fan, Z. (1987). Geochemical investigation of oils and source rocks from Qianjiang Depression of Jianghan basin, a terrigenous saline basin, China. Organic Geochemistry, 11, 549–62.CrossRefGoogle Scholar
Philp, R. P. and Lewis, C. A. (1987). Organic geochemistry of biomarkers. Annual Review of Earth and Planetary Science, 15, 363–95.CrossRefGoogle Scholar
Pickup, S. L. B., Whitmarsh, R. B., Fowler, C. M. R., and Reston, T. J. (1996). Insight into the nature of the ocean-continent transition off West Iberia from a deep multichannel seismic reflection profile. Geology, 24, 1079–82.2.3.CO;2>CrossRefGoogle Scholar
Pierce, C., Whitmarsh, R. B., Scrutton, R. A., Pontoisse, B., Sage, R., and Mascle, J. (1996). Cote d’Ivoire-Ghana margin: seismic imaging of passive rifted crust adjacent to a transform continental margin. Geophysics Journal International, 125, 781–95.Google Scholar
Pierce, W. H. and Khalifa bin Mbarak, A.-H. (1993). Southern Arabian Basin oil habitat: seals and gathering areas. Society of Petroleum Engineers of the American Institute of Mining, Metallurgical and Petroleum Engineers, 8, 103–11.Google Scholar
Pindell, J. (1993). Mesozoic-Cenozoic paleogeographic evolution of northern South America. American Association of Petroleum Geologists Bulletin, 77, 340.Google Scholar
Pinet, P. and Souriau, M. (1988). Continental erosion and large-scale relief. Tectonics, 7, 563–82.CrossRefGoogle Scholar
Pinnock, S. J., Clitheroe, A. R., and Rose, P. T. S. (2002). Appraisal and development of a viscous oil field. In United Kingdom Oil and Gas Fields, Commemorative Millennium Volume, ed. Gluyas, J. G. and Hichens, H. M.. Geological Society of London Memoir, 20, 431–41.Google Scholar
Piper, D. Z. and Link, P. K. (2002). An upwelling model for the Phosphoria sea: a Permian, ocean-margin sea in the northwest United States. American Association of Petroleum Geologists Bulletin, 86, 1217–35.Google Scholar
Piqué, A., Jeannette, D., and Michard, A. (1980). The Western Meseta shear zone, a major and permanent feature of the Hercynian Belt in Morocco. Journal of Structural Geology, 2, 5561.CrossRefGoogle Scholar
Piqué, A., Le Roy, P., and Amrhar, M. (1998). Transtensive synsedimentary tectonics associated with ocean opening: the Essaouira-Agadir segment of the Moroccan Atlantic margin. Journal of the Geological Society of London, 155, 913–28.CrossRefGoogle Scholar
Pitman, W. C. and Andrews, J. A. (1985). Subsidence and thermal history of small pull-apart basins. In Strike-Slip Deformation, Basin Formation and Sedimentation, ed. Biddle, K. T. and Christie-Blick, N.. Special Publication – Society of Economic Paleontologists and Mineralogists, San Antonio, Texas, pp. 45119.CrossRefGoogle Scholar
Planert, M. and Williams, J. S. (1995). Groundwater Atlas of the United States, Segment, California, Nevada. HA-0730-B, US Geological Survey.Google Scholar
Planke, S., Symonds, P. A., Alvestad, E., and Skogseid, J. (2000). Seismic volcanostratigraphy of large-volume basaltic extrusive complexes on rifted margins. Journal of Geophysical Research, 105, 19335–52.CrossRefGoogle Scholar
Platte River Associates, 1995. BasinMod 1-D for WindowsTM. Basin Modeling System. Document Version: 5.0.Google Scholar
Plothner, D. (1988). Entwicklung und Anwending einer erdolgeologisch-geochemischen Explorationsmethode unter besonderer Berucksichtigung der Hydraulik im Pechbronner Gebiet-Fachbericht: das Pechelbronner Feld. Bundensanstalt fur Geowissenschaft und Rohstoffe, BMFT-Forschungsvorhaben, 032 6476, 103 464, 1–8.Google Scholar
Poag, C. W. (1979). Stratigraphy and depositional environments of Baltimore Canyon Trough. AAPG Bulletin, 63, 1452–66.Google Scholar
Podladchikov, Y. Y., Poliakov, A. N. B., and Yuen, D. A. (1994). The effect of lithospheric phase transitions on subsidence of extending continental lithosphere. Earth and Planetary Science Letters, 124, 95103.CrossRefGoogle Scholar
Poelchau, H. S., Baker, D. R., Hantschel, T., Horsfield, B., and Wygrala, B. (1997). Basin simulation and the design of the conceptual basin model. In Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling, ed. Welte, D. H., Horsfield, B., and Baker, D. R.. Springer, Berlin, p. 370.CrossRefGoogle Scholar
Pollack, H. N. (1986). Cratonization and thermal evolution of the mantle. Earth and Planetary Science Letters, 80, 175–82.CrossRefGoogle Scholar
Pollack, H. N. and Chapman, D. S. (1977). On the regional variation of heat flow, geotherms, and lithospheric thickness. Tectonophysics, 38, 279–96.CrossRefGoogle Scholar
Popescu, B. M. (1995). The Guyanas (Guyana, Suriname, French Guyana). In Regional Petroleum Geology of the World, ed. Kulke, H.. 22, 603–12.Google Scholar
Popoff, M. (1988). South Atlantic Gondwana: connections of the Benue Trench with northeastern Brazilian basins up to the opening of the Gulf of Guinea in the Lower Cretaceous. Journal of African Earth Sciences, pp. 409–31.Google Scholar
Porath, H. (1971). Magnetic variation anomalies and seismic low-velocity zone in the western United States. Journal of Geophysical Research, 76, 2643–8.CrossRefGoogle Scholar
Porter, J. R., Knipe, R. J., Fisher, Q. J., Farmer, A. B., Allin, N. S., Jones, L. S., Palfrey, A. J., Garrett, S. W., and Lewis, G. (2000). Deformation processes in Britannia fields, UKCS. Petroleum Geoscience, 6, 241–54.CrossRefGoogle Scholar
Posamentier, H. W. and Allen, G. P. (1993). Variability of the sequence stratigraphic model; effects of local basin factors. Sedimentary Geology, 86 , 91109.CrossRefGoogle Scholar
Posamentier, H. W. and Allen, G. P. (1999). Siliciclastic sequence stratigraphy: concepts and applications. SEPM Concepts in Sedimentology and Paleontology, 7, 210.Google Scholar
Posamentier, H. W., Jervey, M. T., and Vail, P. R. (1988). Eustatic controls on clastic deposition; I, Conceptual framework. Special Publication – Society of Economic Paleontologists and Mineralogists, 42, 109–24.Google Scholar
Posamentier, H. W. and Kolla, V. (2003). Seismic geomorphology and stratigraphy of depositional elements in deep-water settings. Journal of Sedimentary Research, 73, 367–88.CrossRefGoogle Scholar
Posamentier, H. W. and Vail, P. R. (1988). Eustatic controls on clastic deposition: II, Sequence and systems tract models. In Sea-level Changes: An Integrated Approach, ed. Wilgus, C. K., Hastings, B. S., Ross, C. A., Posamentier, H. W., Van Wagoner, J., and Kendall, C.. Special Publication – Society of Economic Paleontologists and Mineralogists, 42, 125–54.Google Scholar
Posgay, K., Albu, I., Mayerová, M., Nakládalová, Z., Ibrmayer, I., Blizkovsky, M., Aric, K., and Gutdeutsch, R. (1991). Contour map of the Mohorovicic discontinuity beneath central Europe. Geophysical Transactions, 36, 713.Google Scholar
Posgay, K., Bodoky, T., Hajnal, Z., Tóth, T. M., Fancsik, T., Hegedüs, E., Kovács, A. C., and Takács, E. (1996). International deep reflection survey along the Hungarian geotraverse. Geophysical Transactions, 40, 144.Google Scholar
Pospíšil, L., Ádám, A., Bimka, J., Bodlak, P., Bodoky, T., Dövényi, P., Granser, H., Hegedüs, E., Joó, I., Kendzera, A., Lenkey, L., Nemčok, M., Posgay, K., Pylypyshyn, B., Sedlák, J., Stanley, W. D., Starodub, G., Szalaiová, V., Šály, B., Šutora, A., Várga, G., and Zsíros, T. (2006). Regional geophysical data on the Carpathian-Pannonian lithosphere. In The Carpathians and Their Foreland: Geology and Hydrocarbon Resources, ed. Pícha, F. and Golonka, J.. American Association of Petroleum Geologists Memoir, pp. 651–97.Google Scholar
Postma, G. (1984). Mass-flow conglomerate in a submarine canyon: Abrioja fan delta, Pliocene, southwest Spain. In Sedimentology of Gravels and Conglomerates, ed. Koster, E. H. and Steel, R. J.. Canadian Society of Petroleum Geologists Memoir, 10, 237–8.Google Scholar
Postma, G. (1995). Sea-level-related architectural trends in coarse- grained delta complexes. Sedimentary Geology, 98, 312.CrossRefGoogle Scholar
Postma, G. and Roep, T. B. (1985). Resedimented conglomerates in the bottomsets of Gilbert-type gravel deltas. Journal of Sedimentary Petrology, 55, 874–85.Google Scholar
Poudjom-Djomani, Y. H., O’Reilly, S. Y., Griffin, W. L., and Morgan, P. (2001). The density structure of subcontinental lithosphere through time. Earth and Planetary Science Letters, 184, 605–21.CrossRefGoogle Scholar
Powell, T. G. (1986). Petroleum geochemistry and deposition setting of lacustrine source rocks. Marine and Petroleum Geology, 3, 200–19.CrossRefGoogle Scholar
Powell, T. G. and Boreham, C. J. (1994). Terrestrially sources oils: Where do they exist and what are our limits of knowledge? – A geochemical perspective. In Coal and Coal-bearing Strata as Oil-prone Source Rocks, ed. Scott, A. C. and Fleet, A. J.. Geological Society of London Special Publications, 77, 1129.Google Scholar
Powell, W. G. (1997). Thermal state of the lithosphere in the Colorado Plateau-Basin and Range transition zone, Utah. PhD thesis, University of Utah, p. 232.Google Scholar
Powley, D. E. (1980). Pressures, normal and abnormal. American Association of Petroleum Geologists Advanced Exploration Schools Unpublished Lecture Notes.Google Scholar
Prat, P., Montenat, C., Ott d’Estevou, P., and Bolze, J. (1986). The western margin of the Gulf of Suez from the study of Gebel Zeit and Gebel Mellaha. Documents et Travaux Institut Géologique Albert de Lapparent, Paris, 10, 4574.Google Scholar
Price, L. C. and Barker, C. E. (1985). Suppression of vitrinite reflectance in amorphous rich kerogen: a major unrecognised problem. Journal of Petroleum Geology, 8, 5984.CrossRefGoogle Scholar
Price, L. C. and Schoell, M. (1995). Constraints on the origins of hydrocarbon gas from compositions of gases at their site of origin. Nature, 378, 368–71.CrossRefGoogle ScholarPubMed
Price, L. C. and Wenger, L. M. (1992). The influence of pressure on petroleum generation and maturation as suggested by aqueous pyrolysis. Organic Geochemistry, 19, 141–59.CrossRefGoogle Scholar
Priest, S. S., Sass, J. H., Ellsworth, B., Farrar, C. D., Sorey, M. L., Hill, D. P., Bailey, R., Jacobson, R. D., Finger, J. T., McConnell, V. S., and Zoback, M. (1998). Scientific drilling in Long Valley, California – what will we learn? USGS Fact Sheet, pp. 77–97.Google Scholar
Prince, C. M. (1999). Textural and diagenetic controls on sandstone permeability. Gulf Coast Association of Geological Societies Transactions, 49, 4253.Google Scholar
Prosser, S. (1993). Rift-related linked depositional systems and their seismic expression. In Tectonics and Seismic Sequence Stratigraphy, ed. Williams, G. D. and Dobbs, A.. Geological Society Special Publications, 71, 3566.Google Scholar
Prost, G. and Aranda, M. (2001). Tectonics and hydrocarbon systems of the Veracruz Bain, Mexico. In The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems, ed. Bartolini, C., Buffler, R. T., and Cantu-Chapa, A.. American Association of Petroleum Geologists Memoir, 75, 271–91.Google Scholar
Purser, B. H. and Bosence, D. W. J. (1998). Sedimentation and Tectonics in Rift Basins: Red 354 Sea – Gulf of Aden. Chapman and Hall, London, p. 663.CrossRefGoogle Scholar
Purvis, K., Kao, J., Flanagan, K., Henderson, J., and Duranti, D. (2002). Complex reservoir geometries in a deep water clastic sequence, Gryphon field, UKCS: Injection structures, geological modelling and reservoir simulation. Marine and Petroleum Geology, 19, 161–79.CrossRefGoogle Scholar
Puteanus, D., Glasby, G. P., Stoffers, P., and Dunzendorf, H. (1991). Hydrothermal iron-rich deposits from the Teahitia-Mehitia and Mackonalk hot spot areas, Southwest Pacific. Marine Geology, 98, 389409.CrossRefGoogle Scholar
Quigley, T. M. and Mackenzie, A. S. (1988). The temperatures of oil and gas formation in the subsurface. Nature, 333, 549–52.CrossRefGoogle Scholar
Raab, M. J., Brown, R. W., Gallagher, K., and Weber, K. (2001). The geomorphic response of the passive continental margin of Namibia to post break-up tectonics. In International Conference and Annual Meeting of the Deutsche Geologische Gesellschaft and Geologische Vereinigung, ed. S. Roth and A. Rueggeberg. Schriftenreihe der Deutschen Geologischen Gesellschaft, Kiel, Federal Republic of Germany, p. 157.Google Scholar
Rabinowicz, M., Dandurand, J. L., Jakubowski, M., Schott, J., and Casson, J. P. (1985). Convection in a North Sea reservoir: influences on diagenesis and hydrocarbon migration. Earth and Planetary Science Letters, 74, 387404.CrossRefGoogle Scholar
Radhakrishna, M., Verma, R. K., and Purushotham, A. K. (2002). Lithospheric structure below the eastern Arabian Sea and adjoining west coast of India based on integrated analysis of gravity and seismic data. Marine Geophysical Researchers, 23, 2542.CrossRefGoogle Scholar
Radney, B. and Byerlee, J. D. (1988). Laboratory studies of the shear strength of montmorillonite and illite under crustal conditions. Eos, Transactions, American Geophysical Union, 69, 1463.Google Scholar
Rădulescu, D. P., Sandulescu, M., and Veliciu, S. (1983). A geodynamic model of the East Carpathians and the thermal field in the lithosphere. Anuarul Institutului de Geologie si Geofizica – Annuaire de l’Institut de Geologie et de Geophysique, 63, 135–44.Google Scholar
Rădulescu, F. (1988). Seismic models of the crustal structure in Romania. Revue Roumaine de Géologie, Géophysique et Géographie, 32, 1319.Google Scholar
Raffensperger, J. P. and Garven, G. (1995). The formation of unconformity-type uranium ore deposits. 1. Coupled groundwater flow and heat transport modeling. American Journal of Science, 295, 581636.CrossRefGoogle Scholar
Raharjo, I., Wannamaker, P., Moore, J. N., Allis, R., Chapman, D., et al., 2002. Magmatic chimney beneath Telaga Bodas revealed by magnetotellurics profiling: a case study at the Karaha Bodas geothermal system, Indonesia. Eos, Transactions, American Geophysical Union, 83, 1423.Google Scholar
Rahe, B., Ferrill, D., Morris, A., and Alan, P. (1998). Physical analog modeling of pull-apart basin evolution. Tectonophysics, 285, 2140.CrossRefGoogle Scholar
Răileanu, V., Diaconescu, C., and Radulescu, F. (1994). Characteristics of Romanian lithosphere from deep seismic reflection profiling. Tectonophysics, 239, 165–85.CrossRefGoogle Scholar
Rampone, E. and Piccardo, G. B. (2000). The ophiolite-oceanic lithosphere analogue: new insights from the Northern Apennines (Italy). Geological Society of America, 349, 2134.Google Scholar
Ranalli, G. (1995). Rheology of the Earth. Chapman and Hall, London, p. 413.Google Scholar
Ranalli, G. and Murphy, D. C. (1987). Rheological stratification of the lithosphere. Tectonophysics, 132, 281–95.CrossRefGoogle Scholar
Ranalli, G. and Rybach, L. (2005). Heat flow, heat transfer and lithosphere rheology in geothermal areas: features and examples. Journal of Volcanology and Geothermal Research, 148, 319.CrossRefGoogle Scholar
Ranganathan, V. and Hanor, J. S. (1987). A numerical model for the formation of saline waters due to diffusion of dissolved NaCl in subsiding sedimentary basins with evaporates. Journal of Hydrology, 92, 97102.CrossRefGoogle Scholar
Rankenburg, K., Lassiter, J. C., and Brey, G. (2005). The role of continental crust and lithospheric mantle in the genesis of Cameroon Volcanic Line lavas: constraints from isotopic variations in lavas and megacrysts from the Biu and Jos plateaux. Journal of Petroleum, 46, 169–90.Google Scholar
Rasskazov, S. V., Boven, A., Ivanov, A. V., and Semenova, V. G. (2000). Middle Quanternary volcanic impulse in the Olekma-Stanovoy mobile system: 40Ar-39Ar dating of volcanics from the Tokinsky Stanovik. Geology of the Pacific Ocean, 19, 1928.Google Scholar
Ratcliffe, N. M. and Burton, W. C. (1985). Fault reactivation models for origin of the Newark Basin and studies related to eastern U.S. seismicity. U. S. Geological Survey Circular, 946, 3645.Google Scholar
Rattey, R. P. and Hayward, A. P. (1993). Sequence stratigraphy of a failed rift system: the Middle Jurassic to Early Cretaceous basin evolution of the Central and Northern North Sea. In Petroleum Geology of Northwest Europe, ed. Parker, J. R.. Proceedings of the 4th Conference. The Geological Society of London, pp. 215–49.Google Scholar
Ravnas, R. and Steel, R. J. (1998). Architecture of marine rift-basin successions. American Association of Petroleum Geologists Bulletin, 82, 100–46.Google Scholar
Ray, D. K. (1963). Tectonic Map of India (1:2,000,000). Geological Society of India, Calcutta.Google Scholar
Raymo, M. E., Hodell, D., and Jansen, E. (1992). Response of deep ocean circulation to initiation of northern hemisphere glaciation (3–2 Ma). Paleoceanography, 7, 645–72.CrossRefGoogle Scholar
Reading, H. G. (1991). The classification of deep-sea depositional systems by sediment calibre and feeder system. Journal of the Geological Society of London, 148, 427–30.CrossRefGoogle Scholar
Reddy, P. R., Venkateswarulu, N., Koteswara Rao, P., and Prasad, A. S. S. S. R. S. (1999). Crustal structure of peninsular shield India from DSS studies. Current Science, 77, 1606–11.Google Scholar
Rees, B. A., Detrick, R. S., and Coakley, B. J. (1993). Seismic stratigraphy of the Hawaiian flexural moat. Geological Society of America Bulletin, 105, 189205.2.3.CO;2>CrossRefGoogle Scholar
Regalli, M. S. P. (1989). A idade dos evaporitos da plataforma continental do Ceara, Brasil, e sua relacao com os outros evaporitos das bacias nordestinas. Boletin de Instituto de Geociencias, Universidade de Sao Paul, Publicacao Especial, 7, 139–43.CrossRefGoogle Scholar
Regelous, M., Niu, Y. L., Abouchanmi, W., and Castillo, P. R. (2009). Shallow origin for South Atlantic Dupal Anomaly from lower continental crust: geochemical evidence from the Mid-Atlantic Ridge at 26 degrees S. Lithos, 112, 5772.CrossRefGoogle Scholar
Rehrig, W. A. and Reynolds, S. J. (1980). Geologic and geochronologic reconnaissance of a northwest-trending zone of metamorphic core complexes in southern and western Arizona. In Cordilleran Metamorphic Core Complexes, ed. Crittenden, M. D., Coney, P. J., and Davis, G. H.. Geological Society of America Memoir, 153, 131–57.Google Scholar
Reid, M. E., Sisson, T. W., and Brien, D. L. (2001). Volcano collapse promoted by hydrothermal alteration and edifice shape, Mount Rainier, Washington. Geology, 29, 779–82.2.0.CO;2>CrossRefGoogle Scholar
Reid, S. A. and McIntyre, J. L. (2001). Monterey Formation porcelanite reservoirs of the Elk Hills field, Kern County, California. American Association of Petroleum Geologists Bulletin, 85, 169–89.Google Scholar
Reiners, P. W., Campbell, I. S., Nicolescu, S., Allen, C., Garver, J. I., and Palin, J. M. (2002). Single crystal helium-lead dating of detrital zircon. Abstracts with Programs – Geological Society of America, 34, 484.Google Scholar
Reisser, K. (2000). Petroleum geology of the western Ust-Yurt Basin, Republic of Kazakhstan. The Bulletin of the Houston Geological Society, 43, 20–1.Google Scholar
Reiter, M., Edwards, C. L., Hartman, H., and Weidman, C. (1975). Terrestrial heat flow along the Rio Grande rift, New Mexico and southern Colorado. Geological Society of America Bulletin, 86, 811–18.2.0.CO;2>CrossRefGoogle Scholar
Reiter, M. A., Mansure, A. J., and Shearer, C. (1979). Geothermal characteristics of the Colorado Plateau. Tectonophysics, 61, 183–95.CrossRefGoogle Scholar
Ren, S., Faleide, J. I., Eldholm, O., Skogseid, J., and Gradstein, F. (2003). Late Cretaceous-Paleocene tectonic development of the NW Vøring Basin. Marine and Petroleum Geology, 20, 177206.CrossRefGoogle Scholar
Renne, P., Deckart, K., Ernesto, M., Feraud, G., and Piccirillo, E. (1996a). Age of the Ponta Grossa dyke swarm (Brazil), and implications to the Parana flood volcanism. Earth and Planetary Science Letters, 144, 199211.CrossRefGoogle Scholar
Renne, P., Ernesto, M., Pacca, I. G., Coe, R. S., Glen, J. M., Prevot, M., and Perrin, M. (1992). The age of Paraná flood volcanism, rifting of Gondwanaland and the Jurassic-Cretaceous boundary: Science, 258, 975–9.CrossRefGoogle ScholarPubMed
Renne, P., Glen, J. M., Milner, S. C., and Duncan, A. R. (1996b) Age of Etendeka flood volcanism and associated intrusions in south-western Africa. Geology, 24(7), 659–62.2.3.CO;2>CrossRefGoogle Scholar
Reston, T. J. (1996). The S reflector west of Galicia: the seismic signature of a detachment fault. Geophysical Journal International, 127, 230–44.CrossRefGoogle Scholar
Reston, T. J., Krawczyk, C. M., and Hoffmann, H. J. (1995). Detachment tectonics during Atlantic rifting: analysis and interpretation of the S reflection, the West Galicia margin. Geological Society Special Publications, 90, 93109.CrossRefGoogle Scholar
Reston, T. J., Krawczyk, C. M., and Klaeschen, D. (1996). The S reflector west of Galicia 727 (Spain): evidence from prestack depth migration for detachment faulting during 728 continental breakup. Journal of Geophysical Research, 101, 8075–91.CrossRefGoogle Scholar
Reston, T. J., Pennell, J., Stubenrauch, A., Walker, I., and Perez-Gussinye, M. (2001). Detachment faulting, mantle serpentinization, and serpentinite-mud volcanism beneath the Porcupine Basin, southwest of Ireland. Geology, 29, 587–90.2.0.CO;2>CrossRefGoogle Scholar
Reynolds, A. D., Simmons, M. D., Bowman, M. B. J., Henton, J., Brayshaw, A. C., Ali-Zade, A. A., Guliyev, I. S., Syleymanova, S. F., Ateava, E. Z., Mamedova, D. N., and Koshkarly, R. O. (1998). Implications of outcrop geology for reservoirs in the Neogene productive series: Apsheron Peninsula, Azerbaijan. American Association of Petroleum Geologists Bulletin, 82, 2549.Google Scholar
Reynolds, D. J. (1984). Structural and dimensional repetition in continental rifting. Master´s thesis, Duke University, Durham, North Carolina, p. 175.Google Scholar
Reynolds, J. G. and Burnham, A. K. (1995). Comparison of kinetic analysis of source rocks and kerogen concentrates. Organic Geochemistry, 23, 1119.CrossRefGoogle Scholar
Reynolds, S. J. (1985). Geology of the South Mountains, central Arizona. Bulletin-State of Arizona, Bureau of Geology and Mineral Technology, Geological Survey Branch, p. 61.Google Scholar
Reynolds, T. (1994). Quantitative analysis of submarine fans in the Tertiary of the North Sea Basin. Marine and Petroleum Geology, 11, 202–7.CrossRefGoogle Scholar
Ribe, N. M. (1996). The dynamics of plume-ridge interaction, 2. Off- ridge plumes. Journal of Geophysical Research, 101, 16195–204.CrossRefGoogle Scholar
Ribe, N. M., Christensen, U. R., and Theiβing, J. (1995). The dynamics of plume-ridge interaction, 1: Ridge-centered plumes. Earth and Planetary Science Letters, 134, 155–68.CrossRefGoogle Scholar
Ribe, N. M. and Delattre, W. L. (1998). The dynamics of plume-ridge interaction, III: The effects of ridge migration. Geophysical Journal International, 133, 511–18.CrossRefGoogle Scholar
Rice, D. D., Fouch, T. D., and Johnson, R. C. (1992). Influence of source rock type, thermal maturity, and migration on composition and distribution of natural gases, Uinta basin, Utah. In Hydrocarbon and Mineral Resources of the Uinta Basin, Utah and Colorado, ed. Fouch, T. D., Nuccio, V. F., and Chidsey, T. C.. Utah Geological Association Guidebook, 20, 95109.Google Scholar
Rice, J. R. (1992). Fault stress states, pore pressure distributions, and the weakness of the San Andreas fault. In Fault Mechanics and Transport Properties of Rocks, ed. Evans, B. and Wong, T.-F.. Academic Press, San Diego, California, pp. 475503.Google Scholar
Richard, P. and Krantz, R. W. (1991). Experiments on fault reactivation in strike-slip mode: experimental and numerical modelling of continental deformation. 5th meeting of the European Union of Geosciences, Symposium on Experimental and Numerical Modelling of Continental Deformation, Strasbourg, France, March 20–23, 1989, 188, 117–31.Google Scholar
Richard, P. D., Naylor, M. A., and Koopman, A. (1995). Experimental models of strike-slip tectonics. Petroleum Geoscience, 1, 7180.CrossRefGoogle Scholar
Richards, M. A., Duncan, R. A., and Courtillot, V. E. (1989). Flood basalts and hot-spot tracks – Plume heads and tails. Science, 246, 105–7.CrossRefGoogle ScholarPubMed
Richardson, R. M. and Coblentz, D. D. (1994). Stress modeling in the Andes: constraints on the South American intraplate stress magnitudes. Journal of Geophysical Research, 99, 22015–26.CrossRefGoogle Scholar
Richardson, S. W. and Oxburg, E. R. (1979). The heat flow field in mainland UK. Nature, 282, 565–7.CrossRefGoogle Scholar
Ridley, J. (1993). The relations between mean rock stress and fluid flow in the crust; with reference to vein- and lode-style gold deposits. Ore Geology Reviews, 8, 2337.CrossRefGoogle Scholar
Ritchie, J. D., Gatliff, R. W., and Richards, P. C. (1999). Early Tertiary magmatism in the offshore NW UK margin and surrounds. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, London, pp. 573–84.Google Scholar
Ritsema, J. and van Heist, H. (2000). New seismic model of the upper mantle beneath Africa. Geology, 28, 63–6.2.0.CO;2>CrossRefGoogle Scholar
Ritter, O., Ryberg, T., Weckmann, U., Hoffmann-Rothe, A., Abueladas, A., Garfunkel, Z., and DESERT Research Group (2003). Geophysical images of the Dead Sea Transform in Jordan reveal an impermeable barrier for fluid flow. Geophysical Research Letters, 30.CrossRefGoogle Scholar
Robert, P. (1988). Organic Metamorphism and Geothermal History. Elf-Aquitaine and Reidel Publishing, Dordrecht.Google Scholar
Roberts, A. M., Lundin, E. R., and Kusznir, N. J. (1997). Subsidence of the Vøring Basin and the influence of the Atlantic continental margin. Journal of the Geological Society of London, 154, 551–7.Google Scholar
Roberts, A. M., Yielding, G., and Badley, M. E. (1993). Tectonic and bathymetric controls on stratigraphic sequences within evolving half-graben. In Tectonics and Seismic Sequence Stratigraphy, ed. Williams, G. D. and Dobb, A.. Geological Society of London Special Publications, 71, 87121.Google Scholar
Roberts, D. G., Montadert, L., and Searle, R. C. (1979). The western Rockall Plateau: stratigraphy and structural evolution. Initial Reports of the Deep Sea Drilling Project, 48, 1061–88.CrossRefGoogle Scholar
Roberts, G. P. (1991). Structural controls on fluid migration through the Rencurel thrust zone, Vercors, French Sub-Alpine chains. In Petroleum Migration, ed. England, W. A. and Fleet, A. J.. Geological Society of London Special Publications, 59, 245–62.Google Scholar
Roberts, G. P. (2007). Fault orientation variations along the strike of active normal fault systems in Italy and Greece: implications for predicting the orientations of subseismic-resolution faults in hydrocarbon reservoirs. American Association of Petroleum Geologists Bulletin, 91, 120.CrossRefGoogle Scholar
Roberts, G. P. and Ganas, A. (2000). Fault-slip directions in central and southern Greece measured from striated and corrugated fault planes: comparison with focal mechanism and geodetic data. Journal of Geophysical Research, 105, 23443–62.CrossRefGoogle Scholar
Roberts, H. H., Aharon, P., Carney, R., and Sassen, R. (1990). See floor responses to hydrocarbon seeps, Louisiana continental slope. Geo-Marine Letters, 10, 232–43.CrossRefGoogle Scholar
Roberts, H. H. and Carney, R. S. (1997). Evidence of episodic fluid, gas and sediment venting on the Northern Gulf of Mexico continental slope. Economic Geology, 92, 863–79.CrossRefGoogle Scholar
Roberts, J. A. and Cramp, A. (1996). Sediment stability on the western flanks of the Canary Islands. Marine Geology, 134, 1330.CrossRefGoogle Scholar
Roberts, M. J. (1991). The South Brae Field, Block 16/07a, U.K. North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. I. L. Abbotts. Memoir of the Geological Society of London, 14, 55–62.Google Scholar
Roberts, S. J. and Nunn, J. A. (1995). Episodic fluid expulsion from geopressured sediments. Marine and Petroleum Geology, 12, 195204.CrossRefGoogle Scholar
Roberts, S. J., Nunn, J. A., Cathles, L., and Cipriani, F. D. (1996). Expulsion of abnormally pressured fluids along faults. Journal of Geophysical Research, 101, 28231–52.CrossRefGoogle Scholar
Robinson, D. and Santana Zamora, A. (1999). The smectite to chlorite transition in the Chipilapa geothermal system, El Salvador. American Mineralogist, 84, 607–19.CrossRefGoogle Scholar
Robinson, N., Eglington, G., Brassel, S. C., and Cranwell, P. A. (1984). Dinoflagellate origin for sedimentary 4-Alpha-methylsteroids and 5-Alpha (H)-stanols. Nature, 308, 439–42.CrossRefGoogle Scholar
Robinson, V. D. and Engel, M. H. (1993). Characterization of the source horizons within the Late Cretaceous transgressive sequence of Egypt. In Source Rocks in a Sequence Stratigraphic Framework, ed. Katz, B. J. and Pratt, L. M.. American Association of Petroleum Geologists Studies in Geology, 37, 101–17.Google Scholar
Robson, D. (1991). The Argyll, Duncan and Innes fields, Blocks 30/24 and 30/25a, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 219–25.Google Scholar
Rockwell, T. K., Keller, E. A., and Dembroff, G. R. (1988). Quaternary rate of folding of the Ventura anticline, western Transverse Ranges, southern California. American Geological Society Bulletin, 100, 850–8.2.3.CO;2>CrossRefGoogle Scholar
Roden-Tice, M. K. and Wintsch, R. P. (2002). Early Cretaceous normal faulting in southern New England: evidence from apatite and zircon fission-track ages. Journal of Geology, 110, 159–78.CrossRefGoogle Scholar
Rodrigues, R., Santos, A. S., and Costa, L. A. (1984). Avaliacao geoquimica da Bacia de Barreirinhas. Internal report Petrobras, Siex 1304017.Google Scholar
Roehler, H. W. (1992). Correlation, composition, areal distribution, and thickness of Eocene stratigraphic units, greater Green River basin, Wyoming, Utah, and Colorado. USGS Professional Paper, 1506, 149.Google Scholar
Rona, P. A., Bougault, H., Charlou, J. L., Appriou, P., Nelsen, T. A., Trefy, J. H., Eberhart, G. L., Barone, A., and Needham, H. D. (1992). Hydrothermal circulation, serpentinization, and degassing at a rift valley-fracture zone intersection: Mid-Atlantic Ridge near 15 degrees N, 45 degrees W. Geology, 20, 783–6.2.3.CO;2>CrossRefGoogle Scholar
Rook, S. H. and Williams, G. C. (1942). Imperial carbon dioxide gas field: summary of operations, California oil fields. Annual Report of the State Oil and Gas Supervisor, 28, 1233.Google Scholar
Rooksby, S. K. (1991). The Miller Field, Blocks 16/7B, 16/8B, UK North Sea. In United Kingdom oil and gas fields, 25 Years Commemorative Volume, ed. I. L. Abbotts. Memoir of the Geological Society of London, 14, 159–64.CrossRefGoogle Scholar
Rose, P. R. (2001). Risk analysis and management of petroleum exploration ventures. American Association of Petroleum Geologists Methods in Exploration Series, 12, 178.Google Scholar
Rosendahl, B. R. (1987). Architecture of continental rifts with special reference to East Africa. Annual Review of Earth and Planetary Sciences, 15, 445503.CrossRefGoogle Scholar
Rosendahl, B. R. and Groschel-Becker, H. (1999). Deep seismic structure of the continental margin in the Gulf of Guinea: a summary report. Geological Society Special Publications, 153, 7583.CrossRefGoogle Scholar
Rosendahl, B. R., Groschel-Becker, H., Meyers, J., and Kaczmarick, K. (1991). Deep seismic reflection study of a passive margin, southeastern Gulf of Guinea. Geology, 19, 291–5.2.3.CO;2>CrossRefGoogle Scholar
Rosendahl, B. R., Mohriak, W. U., Nemčok, M., Odegard, M. E., Turner, J. P., and Dickson, W. G. (2005). West African and Brazilian conjugate margins: crustal types, architecture, and plate configurations. The 4th HGS/PESGB International Conference on African E7P, Houston Program, 3.CrossRefGoogle Scholar
Rosendahl, B. R., Reynolds, D. J., Lorber, P. M., Burgess, C. F., McGill, J., Scott, D., Lambiase, J. J., and Derksen, S. J. (1986). Structural expressions of rifting: lessons from Lake Tanganyika, Africa. In Sedimentation in the African Rifts, ed. Frostrick, L. E.. Geological Society of London Special Publications, 25, 2943.Google Scholar
Rowan, M. G., Hart, B. S., Nelson, S., Flemings, P. B., and Trudgill, B. D. (1998). Three-dimensional geometry and evolution of a salt-related growth-fault array: Eugene Island 330 Field, offshore Louisiana, Gulf of Mexico. Marine and Petroleum Geology, 15, 309–28.CrossRefGoogle Scholar
Rowan, M. G., Jackson, M. P. A., and Trudgill, B. D. (1999). Salt-related fault families and fault welds in the northern Gulf of Mexico. American Association of Petroleum Geologists Bulletin, 83, 1454–84.Google Scholar
Rowan, M. G. and Weimer, P. (1998). Salt–sediment interaction, northern Green Canyon and Ewing Bank (offshore Louisiana), northern Gulf of Mexico. American Association of Petroleum Geologists Bulletin, 82, 1055–82.Google Scholar
Roy, R. F., Beck, A. E., and Toulokian, Y. S. (1981). Thermophysical properties of rocks. In Physical Properties of Rocks and Minerals, ed. Toulokian, Y. S., Judd, W. R., and Roy, R. F.. McGraw-Hill, New York, pp. 409–88.Google Scholar
Royden, L. H. (1985). The Vienna Basin: a thin-skinned pull-apart basin. Strike-slip deformation, basin formation, and sedimentation. SEPM, Special Publications, 37, 319–38.Google Scholar
Royden, L. and Keen, C. E. (1980). Rifting process and thermal evolution of the continental margin of eastern Canada determined from subsidence curves. Earth and Planetary Science Letters, 51, 343–61.CrossRefGoogle Scholar
Rubey, W. W. and Hubbert, M. K. (1960). Role of fluid pressure in mechanics of overthrust faulting, II: Overthrust belt in geosynclinal area of western Wyoming in light of fluid pressure hypothesis. Geological Society of America Bulletin, 60, 167205.Google Scholar
Ruble, T. E., Lewan, M. D., and Philp, R. P. (2001). New insights on the Green River petroleum system in the Uinta Basin from hydrous pyrolysis experiments. American Association of Petroleum Geologists Bulletin, 85, 1333–71.Google Scholar
Ruble, T. E., Lewan, M. D., and Philp, R. P. (2003). New insights on the Green River petroleum system in the Uinta basin from hydrous pyrolysis experiments. Reply. American Association of Petroleum Geologists Bulletin, 87, 1535–41.CrossRefGoogle Scholar
Ruble, T. E. and Philp, R. P. (1998). Stratigraphy, depositional environments and organic geochemistry of source-rocks in the Green River petroleum system, Uinta basin, Utah. In Modern and Ancient Lake Systems: New Problems and Perspectives, ed. Pitman, J. K. and Carroll, A. R.. Geological Association Guidebook, Salt Lake City, Utah, 26, 289328.Google Scholar
Rudnick, R. L., Ireland, T. R., Gehrels, G., Irving, A. J., Chesley, J. T., and Hanchar, J. M. (1999). Dating mantle metasomatism: U-Pb geochronology of zircons in cratyonic mantle xenoliths from Montana and Tanzania. In The Nixon Volume, ed. Gurney, J. J., Gurney, M. D., Pascoe, S. and Richardson, S. R.. Proceedings of the VIIth International Kimberlite Conference, pp. 503–21.Google Scholar
Rudnick, R. L., McDonough, W. F., and O’Connell, R. J. (1998). Thermal structure, thickness and composition of continental lithosphere. Chemical Geology, 145, 395411.CrossRefGoogle Scholar
Ruppel, C. (1995). Extensional processes in continental lithosphere. Journal of Geophysical Research, 100, 24187–215.CrossRefGoogle Scholar
Russell, S. M. and Whitmarsh, R. B. (2003). Magmatism at the west Iberia non-volcanic rifted continental margin: evidence from analyses of magnetic anomalies. Geophysical Journal International, 154, 706–30.CrossRefGoogle Scholar
Rust, D. J. and Summerfeld, M. A. (1990). Isopach and borehole data as indicators of rifted margin evolution in southwestern Africa. Marine and Petroleum Geology, 7, 277–87.CrossRefGoogle Scholar
Rutter, E. H. (1976). The kinetics of rock deformation by pressure solution. Philosophical Transactions of the Royal Society of London, 283, 203–20.Google Scholar
Rutter, E. H. (1983). Pressure solution in nature, theory and experiment. Journal of Geological Society of London, 140, 725–40.CrossRefGoogle Scholar
Rutter, E. H. and Brodie, K. H. (1988). Experimental approaches to the study of deformation metamorphism relationships. Mineralogical Magazine, 52, 3542.CrossRefGoogle Scholar
Rutter, E. H. and White, S. H. (1979). The microstructures and rheology of fault gouges produced experimentally under wet and dry conditions at temperatures up to 400 degrees centigrade. Bulletin de Mineralogie, 102, 102–9.Google Scholar
Rybach, L. (1976). Radioactive heat production: a physical property determined by the chemistry of rocks. In The Physics and Chemistry of Minerals and Rocks, ed. R. G. J. Strens. Wiley and Sons, London, pp. 309–18.Google Scholar
Rybach, L. (1985). The interpretation of heat flow measurements in terms of crustal and lithospheric structure. Terra cognita, 5, 33–4.Google Scholar
Rybach, L. (1986a). Radioactive heat production in rocks and its relation to other petrophysical parameters. Pure and Applied Geophysics, 114, 309–17.Google Scholar
Rybach, L. (1986b). Amount and significance of radioactive heat sources in sediments. In Thermal Modeling in Sedimentary Basins, ed. Burrus, J.. Technip, Paris, pp. 311–22.Google Scholar
Rybach, L. and Buntebarth, G. (1984). The variation of heat generation, density and seismic velocity with rock type in the continental lithosphere. Tectonophysics, 103, 335–44.CrossRefGoogle Scholar
Sage, F., Basile, C., Mascle, J., Pontoise, B., and Whitmarsh, R. B. (2000). Crustal structure of the continent–ocean transition off the Cote d’Ivoire-Ghana transform margin: implications for thermal exchanges across the palaeotransform boundary. Geophysics Journal International, 143, 62678.CrossRefGoogle Scholar
Sakai, H. et al. (1997). Paleomagnetic study with (super 40) Ar/ (super 39) Ar dating of Rajmahal Hills and Mahanadi Graben in India: reconstruction of Gondwanaland. Chishitsugaku Zasshi. Journal of the Geological Society of Japan, 103, 192202.Google Scholar
Sales, J. K. (1997). Seal strength vs. trap closure: a fundamental control on the distribution of oil and gas. In Seals, Traps and the Petroleum System, ed. Surdam, R. C.. American Association of Petroleum Geologists Memoir, 67, 5783.Google Scholar
Salmon, P., Taruno, R. J., and Sadirsan, W. S. (1996). Hydrocarbon exploration in Indonesia in the first half of 1990s decade. International Geological Congress, Abstracts – Congres Geologique Internationale, Resumes, 30, 565.Google Scholar
Sandwell, D. T. and Smith, W. H. F. (1997). Marine gravity anomaly from Geosat and ERS 1 satellite altimetry. Journal of Geophysical Research, 102, 10039–54.CrossRefGoogle Scholar
Sartain, S. M. and See, B. E. (1997). The South Georgia Basin: an integration of Landsat, gravity, magnetics and seismic data to delineate basement structure and rift basin geometry. In 1997 Gulf Coast Association of Societies: Geology across the Gulf, New Offshore Technologies, Keys to Onshore Revitalization, ed. Craig, W. W. and Kohl, B.. Transactions – Gulf Coast Association of Geological Societies, 47, 493–7.Google Scholar
Sassen, R., Roberts, H. H., Aharon, P., Larkin, J., Chinn, E. W., and Carney, R. (1993). Chemosynthetic bacterial mats at cold hydrocarbon seeps, Gulf of Mexico continental slope. Organic Geochemistry, 20, 7789.CrossRefGoogle Scholar
Saunders, A. D., Fitton, J. G., Kerr, A. C., Norry, M. J., and Kent, R. W. (1997). The North Atlantic igneous province. In Large Igneous Provinces: Continental, Oceanic and Planetary Flood Volcanism, ed. Mahoney, J. and Coffin, M. F.. AGU Geophysical Monograph, 100, 4594.Google Scholar
Saunders, I. and Young, A. (1983). Rates of surface processes on slopes, slope retreat and denudation. Earth Surface Processes and Landforms, 8, 473501.CrossRefGoogle Scholar
Sawyer, D. S., Swift, A., Sclater, J. G., and Toksoz, N. (1982). Extensional model for the subsidence of the northern United States Atlantic continental margin, Geology, 10, 134–40, 3/82.Google Scholar
Sawyer, M. J. and Keegan, J. B. (1996). Use of palynofacies characterization in sand-dominated sequences, Brent Group, Ninian Field, UK North Sea. Petroleum Geoscience, 2, 289–97.CrossRefGoogle Scholar
Schaltegger, U., Gebauer, D., and von Quadt, A. (2002). The mafic-ultramafic rock association of Loderio-Biasca (lower Pennine nappes, Ticino, Switzerland): Cambrian oceanic magmatism and its bearing on early Paleozoic paleogeography. Chemical Geology, 186, 265–79.CrossRefGoogle Scholar
Schamel, S. (1993). Hydrocarbon Prospects in the Pechora Basin. Earth Science & Resources Institute, University of South Carolina, ESRI Technical Report, 92-09-373(1).Google Scholar
Schamel, S. and Ressetar, R. (1998). Tectonics and sedimentation of the Great Salt Lake and Uinta Basins, Utah: analogues for nonmarine basins in China. EGI Field Seminar, I00743, 170.Google Scholar
Schardt, C., Large, R., and Yang, J. (2006). Controls on heat flow, fluid migration, and massive sulfide formation of an off-axis hydrothermal system – the Lau Basin perspective. American Journal of Science, 306, 103–34.CrossRefGoogle Scholar
Scheck, M. and Bayer, U. (1999). Evolution of the Northeast German Basin: inferences from a 3D structural model and subsidence analysis. Tectonophysics, 313, 145–69.CrossRefGoogle Scholar
Schenk, C. J., Higley, D. K., and Magoon, L. B. (2003). Region 6 Assessment Summary – Central and South America. U.S. Geological Survey Digital Data Series, 60.Google Scholar
Schenk, H. J. and Horsfield, B. (1998). Using natural maturation series to evaluate the utility of parallel reaction kinetics models: an investigation of Toarcian shales and Carboniferous coals, Germany. Organic Geochemistry, 29, 137–4.CrossRefGoogle Scholar
Schenk, H. J., Horsfield, B., Krooss, B., Schaefer, R. G., and Schwochau, K. (1997). Kinetics of petroleum formation and cracking. In Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling, ed. Welte, D. H., Horsfield, B., and Baker, D. R.. Springer-Verlag, Berlin, pp. 233–69.Google Scholar
Schilling, J.-G. (1973). Iceland mantle plume: geochemical evidence along Reykjanes Ridge. Nature, 242, 565–71.CrossRefGoogle Scholar
Schilling, J.-G. (1985). Upper mantle heterogeneities and dynamics. Nature, 314, 62–7.CrossRefGoogle Scholar
Schilling, J.-G., Kingsley, R., Fontignie, D., Pordea, R., and Xue, S. (1999). Dispersion of the Jan Mayen and Iceland mantle plumes in the Arctic: a He-Pb-Nd-Sr isotope tracer study of basalts from the Kolbeinsey, Mohns, and Knipovich Ridges. Journal of Geophysical Research, 104, 10543–69.CrossRefGoogle Scholar
Schleicher, A. M., van der Pluijm, B. A., Solum, J. G., and Warr, L. N. (2006). Origin and significance of clay-coated fractures in mudrock fragments of the SAFOD borehole (Parkfield, California). Geophysical Research Letters, 33, 16313.CrossRefGoogle Scholar
Schlindwein, V. and Jokat, W. (1999). Structure and evolution of the continental crust of northern east Greenland from integrated geophysical studies. Journal of Geophysical Research, 104, 15227–45.CrossRefGoogle Scholar
Schlische, R. W. (1991). Half-graben basin filling models: new constraints on continental extensional basin development. Basin Research, 3, 123–41.CrossRefGoogle Scholar
Schlische, R. W. (1992). Structural and stratigraphic development of the Newark extension basin, eastern North America: implications for the growth of the basin and its bounding structures. Bulletin of Geological Society of America, 104, 1246–63.2.3.CO;2>CrossRefGoogle Scholar
Schlische, R. W. (1993). Anatomy and evolution of the Triassic-Jurassic continental rift system, eastern North America. Tectonics, 12, 1026–42.CrossRefGoogle Scholar
Schlische, R. W. (1995). Geometry and origin of fault-related folds in extensional settings. American Association of Petroleum Geologists Bulletin, 79, 1661–78.Google Scholar
Schlische, R. W. (2003). Progress in understanding the structural geology, basin evolution, and tectonic history of the eastern North American Rift System. In Aspects of Triassic-Jurassic Rift Basin Geoscience, ed. LeTorneau, P. M. and Olsen, P. E.. Columbia University Press, New York.Google Scholar
Schlische, R. W. and Anders, M. H. (1996). Stratigraphic effects and tectonic implications of the growth of normal faults and extensional basins. In Reconstructing the History of Basin and Range Extension Using Sedimentology and Stratigraphy, Beratan, K. K.. Special Paper – Geological Society of America, pp. 183203.Google Scholar
Schmitt, H. R. H. (1991). The Chanter Field, Block 15/17, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 261–8.Google Scholar
Schmitt, H. R. H. and Gordon, A. F. (1991). The Piper Field, Block 15/17, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 361–8.Google Scholar
Schmucker, U. (1969). Conductivity anomalies, with special reference to the Andes. In The Application of Modern Physics to the Earth and Planetary Interiors, ed. Runcorn, S. K.. Wiley-Interscience, pp. 125–38.Google Scholar
Schoell, M. (1984). Recent advances in petroleum isotope geochemistry. Organic Geochemistry, 6, 645–63.CrossRefGoogle Scholar
Schoell, M., Hwang, R. J., Carlson, R. M. K., and Welton, J. E. (1994). Carbon isotopic composition of individual biomarkers in gilsonites. Organic Geochemistry, 21, 673–83.CrossRefGoogle Scholar
Schöellkopf, N. B. and Patterson, B. A. (1997). Petroleum systems of Cabinda, Angola. AAPG/ ABGP Hedberg Research Symposium, Extended Abstracts Volume, Rio de Janeiro, Brazil.Google Scholar
Scholl, D. W., Christensen, M. N., Von Huene, R., and Marlow, M. S. (1970). Peru-Chile trench sediments and sea-floor spreading. Geological Society of America Bulletin, 81, 1339–60.CrossRefGoogle Scholar
Scholten, J. C., Scott, S. D., Garbe-Schoenberg, D., Fietzke, J., Thomas, B., and Kennedy, C. B. (2004). Hydrothermal iron and manganese crusts from the Pitcairn Hotspot region. In Oceanic Hotspots: Intraplate Submarine Magmatism and Tectonism, ed. Hekinian, R., Stoffers, P., and Cheminee, J.-L.. Springer, Berlin.Google Scholar
Scholz, C. H. (1968). Microfracturing and the inelastic deformation of rock in compression. Journal of Geophysical Research, 73, 1417–32.CrossRefGoogle Scholar
Schon, J. (1983). Petrophysik: Physikalische Eigenschaften von Gesteinen und Mineralen. Ferdinand Enke Verlag, Stuttgart.Google Scholar
Schou, L., Eggen, S., and Schoell, M. (1985). Oil-oil and oil-source rocks correlation, northern North Sea. In Petroleum Geochemistry in Exploration of the Norwegian Shelf, ed. Thomas, B. M., Dore, A. G., Eggen, S. S., Home, P. C., and Larsen, R. Mange. Graham and Trotman, London, pp. 101–17.Google Scholar
Schowalter, T. T. (1979). Mechanics of secondary hydrocarbon migration and entrapment. American Association of Petroleum Geologists Bulletin, 63, 723–60.Google Scholar
Schubert, C. (1980). Late-Cenozoic pull-apart basins, Bocono fault zone, Venezuelan Andes. Journal of Structural Geology, 2, 463–8.CrossRefGoogle Scholar
Schueller, S., Gueydan, F., and Davy, P. (2005). Brittle-ductile coupling: role of ductile viscosity on brittle fracturing. Geophysical Research Letters, 32, 4.CrossRefGoogle Scholar
Schulte, W. M., Van Rossem, P. A. H., and Van de Vover, W. (1994). Current challenges in the Brent Field. Journal of Petroleum Technology, 46, 1073–9.CrossRefGoogle Scholar
Schultz, R. A. and Moore, J. M. (1996). New observations of grabens from the Needles district, Canyonlands National Park, Utah. In Geology and Resources of the Paradox Basin, ed. Huffman, A. C., Lund, W. R., and Godwin, L. H.. Utah Geological Association, pp. 295302.Google Scholar
Sclater, J. G. and Christie, P. A. F. (1980). Continental stretching: an explanation of the post mid-Cretaceous subsidence of the central North Sea basin. Journal of Geophysical Research, 85, 3711–39.CrossRefGoogle Scholar
Sclater, J. G., Royden, L., Horváth, F., Burchfield, C., Sempken, S., and Stegena, S. (1980). The formation of the intra-Carpathian basins as determined from subsidence data. Earth and Planetary Science Letters, 51, 137–62.CrossRefGoogle Scholar
Scotese, C. R. (1998). Computer software to produce plate tectonic reconstructions. In Gondwana 10: Event Stratigraphy of Gondwana, ed. Almond, J., Anderson, J., Booth, P., Chinsamy-Turan, A., Cole, D., De Wit, M. J., Rubridge, B., Smith, R., van Bever Donker, J., and Storey, B. C.. Journal of African Earth Sciences, 27, 171–2.Google Scholar
Scrutton, R. A. (1984). Modeling of mnagmatic and gravity anomalies at Goiban Spur, northeast Atlantic. In Initial Reports of the Deep Sea Drilling Project 80, ed. P. C. de Graciansky and C. W. Poag. U.S. Government Printing Office, Washington, DC.Google Scholar
Sebai, A., Stutzmann, E., Montager, J. P., Sicilia, D., and Beucler, E. (2006). Anisotropic structure of the African upper mantle from Rayleigh and Love wave tomography. Physics of the Earth and Planetary Interiors, 155, 4862.CrossRefGoogle Scholar
Sedwick, P. N., McMurtry, G. M., and Macdougall, J. D. (1992). Chemistry of hydrothermal solutions from Pele’s Vents, Loihi Seamount, Hawaii. Geochimica et Cosmochimica Acta, 56, 3643–67.CrossRefGoogle Scholar
Seewald, J. S., Benitez-Nelson, B. C., and Whelan, J. K. (1998). Laboratory and theoretical constraints on the generation and composition of natural gas. Geochimica et Cosmochimica Acta, 62, 1599–617.CrossRefGoogle Scholar
Segall, M. P., Nash, G., Umbriaco, J., Kessler, C., Dudley-Murphy, E., and Sawyer, R. K. (2003). BHP-Billiton – EGI DSDP-ODP digital database project. Phase I – Northwest Africa. EGI Report, 50500985-03-30-03, p. 20, 32 enclosures, 3 CDs.Google Scholar
Segall, M. P. and Pollard, D. D. (1980). Mechanics of discontinuous faults. Journal of Geophysical Research, 85, 4337–50.CrossRefGoogle Scholar
Seiglie, G. and Baker, M. (1984). Relative sea level changes during the Middle and Late Cretaceous from Zaire to Cameroon (central West Africa). In Interregional Unconformities and Hydrocarbon Accumulation, ed. J. S. Schlee. AAPG Memoir 36, pp. 81–8.Google Scholar
Self, S., Thordarson, T., and Keszthelyi, L. (1997). Emplacement of continental flood basalt lava flows. In Large Igneous Provinces: Continental, Oceanic and Planetary Flood Volcanism, ed. Mahoney, J. J. and Coffin, M. F.. AGU Geophysical Monograph, 100, 381410.Google Scholar
Sengör, A. M. C. (1976). Collision of irregular continental margins: implications for foreland deformation of Alpine-type orogeny. Geology, 4, 779–85.2.0.CO;2>CrossRefGoogle Scholar
Sengör, A. M. C. (2001). Elevation as indicator of mantle-plume activity. In Mantle Plumes: Their Identification through Time, ed. R. E. Ernst and K. L. Buchan. Geological Society of America Special Paper, 352, 183–225.Google Scholar
Sengör, A. M. C. and Burke, K. (1978). Relative timing of rifting and volcanism on Earth and its tectonic implications. Geophysical Research Letters, 5, 419–21.CrossRefGoogle Scholar
Shaler, N. S. and Woodworth, J. B. (1899). Geology of the Richmond Basin, Virginia. U.S. Geological Survey Annual Report, 19, 1246–63.Google Scholar
Sharp, I. R., Gawthorpe, R. L., Armstrong, B., and Underhill, J. R. (2000). Propagation history and passive rotation of mesoscale normal faults: implications for synrift stratigraphic development. Basin Research, 12, 285305.CrossRefGoogle Scholar
Sharp, J. M. (1978). Energy and momentum transport model of the Ouachita basin and its possible impact on formation of economic mineral deposits. Economic Geology, 73, 1057–68.CrossRefGoogle Scholar
Sharp, J. M. Jr. (1983). Permeability controls on aquathermal pressuring. American Association of Petroleum Geologists Bulletin, 67, 2057–61.Google Scholar
Shaw, W. J. and Lin, J. (1996). Models of ocean ridge lithospheric deformation: dependence on crustal thickness, spreading rate, and segmentation. Journal of Geophysical Research, 101, 17977–93.CrossRefGoogle Scholar
Shearman, D. J., Mossop, G., Dunsmore, H., and Martin, M. (1972). Origin of gypsum veins by hydraulic fracture. Institution of Mining and Metallurgy, Transactions, Applied Earth Science, 81, 149–55.Google Scholar
Sheng, G., Fu, J., Brassell, S. C., Gowar, A. P., Eglinton, G., Damste, J. S. S., de Leeuw, J. W., and Schenk, P. A. (1987). Sulfur-containing compounds in sulfur-rich crude oils from hypersaline lake sediments and their geochemical implications. Geochemistry, 6, 115–26.CrossRefGoogle Scholar
Shepherd, M. (1991). The Magnus Field, Blocks 211/7a, 12a, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 153–7.Google Scholar
Sheridan, R. E., Musser, D. L., Glover, L., Talwani, M., Ewing, J. I., Holbrook, W. S., Purdy, G. M., Hawman, R., and Smithson, S. (1993). Deep seismic reflection data of EDGE U.S. Atlantic continent margin experiment: implications for Appalachian sutures and Mesozoic rifting and magmatic underplating. Geology, 21, 563–7.2.3.CO;2>CrossRefGoogle Scholar
Sheth, H. C. (2005). From Deccan to Réunion: no trace of a mantle plume. Geological Society of America Special Paper, 388, 477501.Google Scholar
Shevenell, L. (2005a). Crustal Thickness Map. Great Basin Center, University of Nevada, Reno.Google Scholar
Shevenell, L. (2005b). Shear Strain Distribution. Great Basin Center, University of Nevada, Reno.Google Scholar
Shi, J. Y., Mackenzie, A. S., Alexander, R., Eglinton, G., Gowar, A. P., Wolff, G., and Maxwell, J. R. (1982). A biological marker investigation of petroleums and shales from the Shengli oilfield, the People’s Republic of China. Chemical Geology, 35, 131.Google Scholar
Shi, Y. and Wang, C. Y. (1986). Pore pressure generation in sedimentary basins: overloading versus aquathermal. Journal of Geophysical Research, 91, 2153–62.CrossRefGoogle Scholar
Shikazono, N. and Holland, H. D. (1983). The partitioning of strontium between anhydrite and aqueous solutions from 150 degrees to 250 degrees C. Economic Geology Monograph, 5, 320–8.Google Scholar
Shillington, D. J., Holbrook, W. S., Tucholke, B. E., Hopper, J. R., Louden, K. E., Larsen, H. C., Van Avendonk, H. J. A., Deemer, S., and Hall, J. (2004). Data report: marine geophysical data on the Newfoundland nonvolcanic rifted margin around SCREECH Transect 2. Proceedings of the Ocean Drilling Program, Initial Report, 210.Google Scholar
Shinohara, H., Kazahaya, K., and Lowenstern, J. B. (1995). Volatile transport in a convecting magma column: implications for porphyry Mo mineralization. Geology, 23, 1091–4.2.3.CO;2>CrossRefGoogle Scholar
Shirley, D. N. (1986). Compaction of igneous cumulates. Journal of Geology, 94, 795809.CrossRefGoogle Scholar
Shirley, D. N. (1987). Differentiation and compaction in the Palisades Sill. Journal of Petrology, 28, 835–65.CrossRefGoogle Scholar
Sial, A. N. (1976). The post-Paleozoic volcanism of Northeast Brazil and its tectonic significance. Anais da Academia Brasileira de Ciencias, 48, 299311.Google Scholar
Sial, A. N., Long, L. E., Pessoa, D. A. R., and Kawashita, K. (1981). Potassium-argon ages and strontium isotope geochemistry of Mesozoic and Tertiary basaltic rocks, northeastern Brasil. Anais da Academia Brasileira de Ciencias, 53, 115–22.Google Scholar
Sibson, R. H. (1974). Frictional constraints on thrust, wrench and normal faults. Nature, 249, 542–4.CrossRefGoogle Scholar
Sibson, R. H. (1981a). Controls on low-stress hydro-fracture dilatancy in thrust, wrench and normal fault terrains. Nature, 289, 665–7.CrossRefGoogle Scholar
Sibson, R. H. (1981b). Fluid flow accompanying faulting: field evidence and models. In Earthquake Prediction: An International Review, ed. Simpson, D. W. and Richards, P. G.. American Geophysical Union, Maurice Ewing Series, 4, 593603.Google Scholar
Sibson, R. H. (1982). Fault zone models, heat flow, and the depth distribution of seismicity in the continental crust of the United States. Seismological Society of America Bulletin, 72, 151–63.Google Scholar
Sibson, R. H. (1986). Brecciation processes in fault zones: inferences from earthquake rupturing. Pure and Applied Geophysics, 124, 159–75.CrossRefGoogle Scholar
Sibson, R. H. (1990a). Conditions for fault-valve behaviour. In Deformation Mechanisms, Rheology and Tectonics, ed. Knipe, R. J. and Rutter, E. H.. Geological Society of London Special Publications, pp. 1528.Google Scholar
Sibson, R. H. (1990b). Faulting and Fluid Flow. In Fluids in Tectonically Active Regimes of the Continental Crust, ed. Nesbitt, B. E.. Mineralogical Association of Canada Short Course, 18, 93132.Google Scholar
Sibson, R. H. (1994). Hill fault/fracture meshes as migration conduits for overpressured fluids. In Proceedings of Workshop LXIII, US Geological Survey Red-Book Conference on the Mechanical Involvement of Fluids in Faulting, ed. H. Hickman, R. H. Sibson, R. L. Bruhn, and M. L. Jacobson. Open-File Report, 94–0228, pp. 224–30.Google Scholar
Sibson, R. H. (1996). Structural permeability of fluid-driven fault-fracture meshes. Journal of Structural Geology, 18, 1031–42.CrossRefGoogle Scholar
Sibson, R. H. (2003). Brittle-failure controls on maximum sustainable overpressure in different tectonic regimes. American Association of Petroleum Geologists Bulletin, 87, 901–8.CrossRefGoogle Scholar
Sibson, R. H., Robert, F., and Poulsen, K. H. (1988). High-angle reverse faults, fluid-pressure cycling, and mesothermal gold-quartz deposits. Geology, 16, 551–5.2.3.CO;2>CrossRefGoogle Scholar
Siddiqui, F. I. (1996). A dynamic theory of hydrocarbon migration and trapping. PhD thesis, University of Texas, Austin.Google Scholar
Siddiqui, F. I. and Lake, L. W. (1997). A comprehensive dynamic theory of hydrocarbon migration and trapping. Proceedings of the SPE 72nd Annual Technical Conference, San Antonio, October 5–8, 1997, pp. 395–410.CrossRefGoogle Scholar
Silver, L. T., James, E. W., and Chappell, B. W. (1988). Petrological and geochemical investigations at the Cajon Pass deep drillhole. Geophysical Research Letters, 15, 961–4.CrossRefGoogle Scholar
Sikora, P. J., Bergen, J. A., and Farmer, C. A. (1999). Chalk palaeoenvironments and depositional model, Valhall-Hod fields, southern Norwegian North Sea. In Biostratigraphy in Production and Development Geology, ed. Jones, R. W. and Simmons, M. D.. Geological Society of London Special Publications, 152, 113–38.Google Scholar
Simancas, J. F., Carbonnell, R., Gonzalez Lodeiro, F., Perez Estaun, A., Juhlin, C., Ayarza, P., Kashubin, A., Azor, A., Martinez Poyatos, D., Almodovar, G. R., Pascula, E., Saez, R., and Exposito, I. (2003). Crustal structure of the transpression Variscan orogen of SW Iberia: SW Iberia deep seismic reflection profile (IBERSEIS). Tectonics, 22, 19.CrossRefGoogle Scholar
Simon, N. S. C. and Podladchikov, Y. Y. (2006). The effect of mantle petrology on lithosphere dynamics during extension. EGU General Assembly Vienna, Scientific Programme, 279.Google Scholar
Simoneit, B. R. T. and Stuermer, D. H. (1982). Organic geochemical indicators for sources of organic matter and paleoenvironmental conditions in Cretaceous oceans. In Nature and Origin of Cretaceous Carbon-rich Facies, ed. Schlanger, S. O. and Cita, M. B.. Academic Press, London, pp. 145–63.Google Scholar
Simons, F. J., Zielhuis, A., and van der Hilst, R. D. (1999). The deep structure of the Australian continent from surface wave tomography. Lithos, 48, 1743.CrossRefGoogle Scholar
Sims, D., Ferrill, D. A., and Stamatakos, J. A. (1999). Role of a ductile decollement in the development of pull-apart basins; experimental results and natural examples. Journal of Structural Geology, 21, 533–54.CrossRefGoogle Scholar
Sinclair, H. D. (1997). Tectonostratigraphic model for underfilled peripheral foreland basins: an Alpine perspective. Geological Society of America Bulletin, 109, 324–46.2.3.CO;2>CrossRefGoogle Scholar
Sinclair, H. D., Coakley, B. J., Allen, P. A., and Watts, A. B. (1991). Simulation of foreland basin stratigraphy using a diffusion model of mountain belt uplift and erosion: an example from the Central Alps, Switzerland. Tectonics, 10, 599620.CrossRefGoogle Scholar
Singh, A. P. (1999). The deep crustal accretion beneath the Laxmi Ridge in the northeastern Arabian Sea: the plume model again. Journal of Geodynamics, 27, 609–22.CrossRefGoogle Scholar
Singh, A. P. (2002). Impact of Deccan volcanism on deep crustal structure along western part of Indian mainland and adjoining Arabian Sea. Current Science, 82, 316–25.Google Scholar
Singh, A. P. and Mall, D. M. (1998). Crustal accretion beneath Koyana coastal region (India) and late Cretaceous geodynamics. Tectonophysics, 290, 285–97.CrossRefGoogle Scholar
Singh, R. N. (1976). Measurement and analysis of strata deformation around mining excavations. PhD thesis, University of Wales, Cardiff.Google Scholar
Sinha, S. T. (2013). Factors controlling a micro-continent development during continental break up: the Elan Bank case study. PhD thesis, Masaryk University, Brno.Google Scholar
Sinha, S. T., Nemčok, M., Choudhuri, M., Misra, A. A., Sharma, S. P., Sinha, N., and Sujata Venkatraman, S. (2010). The crustal architecture and continental break up of East India Passive margin: an integrated study of deep reflection seismic interpretation and gravity modeling, AAPG Annual Convention & Exhibition, New Orleans, USA, Search and Discovery Article #40611.Google Scholar
Sinninghe Damste, J. S., Kenig, F., Koopmans, M. P., Koster, J., Schouten, S., Haynes, J. M., and de Leeuw, J. W. (1995). Evidence for gammacerane as an indicator of water column stratification. Geochimica et Cosmochimica Acta, 59, 1895–900.CrossRefGoogle ScholarPubMed
Sixsmith, P. J., Hampson, G. J., Gupta, S., and Johnson, H. D. (2004a). Sedimentology and facies architecture of a transgressive sandstone: the Cretaceous Hosta Sandstone, New Mexico, USA. American Association of Petroleum Geologists Annual Meeting Expanded Abstracts, 13, 129.Google Scholar
Sixsmith, P. J., Hampson, G. J., Gupta, S., Johnson, H. D. and Fofana, J. F. (2004b). Facies architecture of a net transgressive sandstone reservoir analog: the Cretaceous Hosta Tongue, New Mexico. AAPG Bulletin, 92(4), 513–47.Google Scholar
Sixsmith, P. J., Hampson, G. J., Gupta, S., Johnson, H. D., and Fofana, J. F. (2008). Facies architecture of a net transgressive sandstone reservoir analog: the Cretaceous Hosta Tongue, New Mexico. American Association of Petroleum Geologists Bulletin, 92, 513–47.CrossRefGoogle Scholar
Skempton, A. W. (1985). Residual strength of clays in landslides, folded strata and the laboratory. Geotechnique, 35, 318.CrossRefGoogle Scholar
Skjerven, J., Rijs, F., and Kalheim, J. E. (1983). Late Paleozoic to early Cenozoic structural development of the south- southeastern Norwegian North Sea. Geologie en Mijnbouw. Netherland Journal of Geosciences, 62, 3545.Google Scholar
Skogseid, J. (2001). Volcanic margins: geodynamic and exploration aspects. Marine and Petroleum Geology, 18, 457–61.CrossRefGoogle Scholar
Skogseid, J., Pedersen, T., Eldholm, O., and Larsen, B. T. (1992). Tectonism and magmatism during NE Atlantic continental break-up: the Vøring Margin. Geological Society of London Special Publications, 68, 305–20.CrossRefGoogle Scholar
Skogseid, J., Planke, S., Faleide, J. I., Pedersen, T., Eldholm, O., and Neverdal, F. (2000). NE Atlantic continental rifting and volcanic margin formation. In Dynamics of Norwegian Margin, ed. Nottvedt, A., Larsen, B. T., Olaussen, S., Tørudbakken, B., Skogseid, J., Gabnelson, R. H., Brekke, H., and Birkeland, O.. Geological Society of London Special Publications, 167, 295326.Google Scholar
Sleep, N. H. (1990). Hotspots and mantle plumes: some phenomenology. Journal of Geophysical Research, 95, 6715–36.CrossRefGoogle Scholar
Sleep, N. H. (2005). Evolution of the continental lithosphere. Annual Reviews of the Earth and Planetary Sciences, 33, 369–93.CrossRefGoogle Scholar
Sloss, L. L. (1979). Global sea level change: a view from the craton. American Association of Petroleum Geologists Memoir, 29, 461–7.Google Scholar
Sloss, L. L. (1988). Forty years of sequence stratigraphy. Geological Society of America Bulletin, 100, 1661–5.2.3.CO;2>CrossRefGoogle Scholar
Sluijk, D. and Nederlof, M. H. (1984). Worldwide geological experience as a systematic basis for prospect appraisal. In Petroleum Geochemistry and Basin Evaluation, ed. Demaison, G. and Murris, R. J.. American Association of Petroleum Geologists Memoir, 35, 15–26.Google Scholar
Smallwood, J. R., Staples, R. K., Richardson, K. R., and White, R. S. (1999). Crust generated above the Iceland mantle plume: from continental rift to oceanic spreading center. Journal of Geophysical Research, 104, 22885–902.CrossRefGoogle Scholar
Smallwood, J. R. and White, R. S. (2002). Ridge-plume interaction in the North Atlantic and its influence on continental breakup and seafloor spreading. In The North Atlantic Igneous Province: Stratigraphy, Tectonic, Volcanic and Magmatic Processes, ed. Jolley, D. W. and Bell, B. R.. Geological Society of London Special Publications, 197, 1537.Google Scholar
Smith, A. D. and Lewis, C. (1999). The planet beyond the plume hypothesis. Earth Science Reviews, 48, 135–82.CrossRefGoogle Scholar
Smith, J. T. (1994). The petroleum system logic as an exploration tool in a frontier setting. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 2549.Google Scholar
Smith, K. and Ritchie, J. D. (1993). Jurassic volcanic centres in the Central North Sea. In Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference, ed. Parker, J. R.. The Geological Society of London, pp. 519–31.Google Scholar
Smith, L. and Chapman, D. S. (1983). On the thermal effects of groundwater flow, 1: Regional scale systems. Journal of Geophysical Research, 88, 593608.CrossRefGoogle Scholar
Smith, L., Forster, C. B., and Evans, J. P. (1990). Interaction of fault zones, fluid flow, and heat transfer at the basin scale. International Association of Hydrogeologists, Selected Papers, 2, 4167.Google Scholar
Smith, R. B., Nagy, Walter C., Julander, Kelsey A., Viveiros, John J., Barker, Craig A., Gants, Donald G. (1989). Geophysical and tectonic framework of the eastern Basin and Range-Colorado Plateau-Rocky Mountain transition. Memoir – Geological Society of America, 172, 205–33.CrossRefGoogle Scholar
Smith, R. E. and Wiltschko, D. V. (1996). Generation and maintenance of abnormal fluid pressures beneath a ramping thrust sheet: isotropic permeability experiments. Journal of Structural Geology, 18, 951–70.CrossRefGoogle Scholar
Smith, R. I., Hodgson, N., and Fulton, M. (1993). Salt control on Triassic reservoir distribution, UKCS Central North Sea. In Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference, ed. J. R. Parker. Geological Society of London, Bath, pp. 547–57.CrossRefGoogle Scholar
Smith, R. L. (1987). The structural development of the Clyde Field. In Petroleum Geology of North West Europe, ed. Brooks, J. and Glennie, K. W.. Graham and Trotman, London, pp. 523–31.Google Scholar
Smith, W. H. F. and Sandwell, D. T. (1997). Seafloor topography from satellite altimetry and ship depth soundings. Science, 277, 1956–62.CrossRefGoogle Scholar
Smoot, J. P. (1983). Depositional subenvironments in an arid closed basin: the Wilkins Peak Member of the Green River Formation (Eocene), Wyoming, U.S.A. Sedimentology, 30, 801–27.CrossRefGoogle Scholar
Snoke, A. W., Howard, K., and Lush, A. P. (1984). Polyphase Mesozoic-Cenozoic deformational history of the northern Ruby Mountains-East Humboldt Range, Nevada. In Field Trip Guide Book, ed. Lintz, J., Jr. University of Nevada.Google Scholar
Snowdon, L. R. (1991). Oil from type III organic matter: resinite revisited. Organic Geochemistry, 17, 743–7.CrossRefGoogle Scholar
Soderstrom, B., Forsberg, A., Holtar, E., and Rasmussen, B. A. (1991). The Mjolner Field, a deep Upper Jurassic oil field in the Central North Sea. First Break, 9, 156–71.CrossRefGoogle Scholar
Sofer, Z. (1984). Stable carbon isotope composition of crude oils: applications to source depositional environments and petroleum alteration. AAPG Bulletin, 68, 3149.Google Scholar
Solli, T. (1995). Upper Jurassic play concept – an integrated study in Block 34/7, Norway. First Break, 13, 2130.CrossRefGoogle Scholar
Sommer, F. (1978). Diagenesis of Jurassic sandstones in the Viking Graben. Journal of the Geological Society of London, 135, 63–7.CrossRefGoogle Scholar
Sorel, D. (2000). A Pleistocene and still-active detachment fault and the origin of the Corinth-Patras Rift, Greece. Geology, 28, 83–6.2.0.CO;2>CrossRefGoogle Scholar
Sorel, D., Keraudren, B., Muller, C., Bahain, J.-C., and Falgueres, C. (1997). Holocene uplift and palaeoseismicity on the Eliki fault, western Gulf of Corinth, Greece. Annales Geofisicae, 39, 575–88.Google Scholar
Sorey, M., Evans, B., Kennedy, M., Rogie, J., and Cook, A. (1999). Magmatic gas emissions from Mammoth Mountain, Mono County, California. California Geology, 52, 416.Google Scholar
Souza, J. M. (1982). Transmission of seismic energy through the Brazilian Parana Basin layered basalt stack. Master’s thesis, University of Texas, Austin.CrossRefGoogle Scholar
Spaak, P., Almond, J., Salahudin, S., Mohd Salleh, Z., and Tosun, O. (1999). Fulmar: a mature field revisited. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, The Geological Society of London, pp. 1089–100.Google Scholar
Spadini, G., Robinson, A. G., and Cloetingh, S. A. P. L. (1996). Western versus eastern Black Sea tectonic evolution: pre-rift lithospheric controls on basin formation. Tectonophysics, 266, 139–54.CrossRefGoogle Scholar
Spadini, G., Robinson, A. G., and Cloetingh, S. A. P. L. (1997). Thermomechanical modeling of Black Sea basin formation, subsidence, and sedimentation. American Association of Petroleum Geologists Memoir, 68, 1938.Google Scholar
Sparks, R. S. J., Huppert, H. E., Kerr, R. C., McKenzie, D. P., and Tait, S. R. (1985). Postcumulus processes in layered intrusions. Geological Magazine, 122, 555–68.CrossRefGoogle Scholar
Spencer, A. M., Home, P. C., and Wiik, V. (1986). Habitat of hydrocarbons in the Jurassic Ula trend, Central Graben, Norway. In Habitat of Hydrocarbons on the Norwegian Continental Shelf, ed. Spencer, A. M., Campbell, C. J., Hanslein, S. H., Nelson, P. H., Nysaether, E., and Ormaasen, E. G.. Graham and Trotman, London, pp. 111–28.Google Scholar
Spencer, A. M., Leckie, G. G., and Chew, K. J. (1996). North Sea hydrocarbon plays and their resources. First Break, 14, 345–57.CrossRefGoogle Scholar
Spiegelman, M. and McKenzie, D. (1987). Simple 2-D models for melt extraction at mid-ocean ridges and island arcs. Earth and Planetary Science Letters, 83, 137–52.CrossRefGoogle Scholar
Spiegelman, M. and Reynolds, J. R. (1999). Combined dynamic and geochemical evidence for convergent melt flow beneath the East Pacific Rise. Nature, 402, 282–5.CrossRefGoogle Scholar
Spiers, C. J. and Schutjens, P. M. T. (1990). Densification of crystalline aggregates by fluid phase diffusional creep. In Deformation Processes in Minerals, Ceramics and Rocks, ed. Barber, D. J. and Meredith, P. G.. Mineralogical Society Series, 1, 334–52.Google Scholar
Srivastava, S. P., Sibuet, J.-C., Cande, S., Roest, W. R., and Reid, I. D. (2000). Magnetic evidence for slow spreading during the formation of the Newfoundland and Iberian margins. Earth and Planetary Science Letters, 182, 6176.CrossRefGoogle Scholar
Srivastava, S. P. and Tapscott, C. R. (1986). Plate kinematics of the North Atlantic. In The Western North Atlantic Region: The Geology of North America, ed. Vogt, P. R. and Tucholke, B. E.. Geological Society of America, Boulder, pp. 379404.Google Scholar
Stacey, F. D. (1992). Physics of the Earth. Brookfield Press, Brisbane, p. 513.Google Scholar
Stach, E., Mackowsky, M. T., Teichmueller, M., Taylor, G. H., Chandra, D., and Teichmueller, R. (1982). Stach’s Textbook of Coal Petrology, Third Revised and Enlarged Edition. E. Schweizerbart’sche Verlagsbuchhandlung, Stuttgart, p. 569.Google Scholar
Stahle, V., Frenzel, G., Kobe, B., Michard, A., Puchelt, H., and Schneider, W. (1990). Zircon syenite pegmatites in the Finero peridotite (Ivrea Zone): evidence for a syenite from mantle source. Earth and Planetary Science Letters, 101, 196205.CrossRefGoogle Scholar
Steckler, M. S. (1981a). The thermal and mechanical evolution of Atlantic – type continental margins. PhD thesis, New York, Columbia University, p. 261.Google Scholar
Steckler, M.S. (1981b). A thermomechanical model of the subsidence of continental margins. Eos, Transactions, American Geophysical Union, 62, 390.Google Scholar
Steckler, M. S. (1985). Uplift and extension of the Gulf of Suez: indications of induced mantle convection. Nature, 317, 135–9.CrossRefGoogle Scholar
Steckler, M. S., Berthelot, F., Lyberis, N., and Le Pichon, X. (1988). Subsidence in the Gulf of Suez: implications for rifting and plate kinematics. Tectonophysics, 153, 249–70.CrossRefGoogle Scholar
Steckler, M. S., Lavier, L., Feinstein, S., Kohn, B. P., and Eyal, M. (1994). Pattern of mantle thinning from subsidence and heat flow measurements in the Gulf of Suez: a case for along-strike flow from the Red Sea. Eos, Transactions, American Geophysical Union, 75, 589.Google Scholar
Steckler, M. S., Omar, G. I., Karner, G. D., and Kohn, B. P. (1993). Pattern of hydrothermal circulation within the Newark basin from fission-track analysis. Geology, 21, 735–8.2.3.CO;2>CrossRefGoogle Scholar
Steel, R., Gjelberg, J., Helland-Hansen, W., Kleinspehn, K., Nottvedt, A., and Rye-Larsen, M. (1985). Tertiary strike-slip basins and orogenic belt of Spitsbergen. Special Publication – Society of Economic Paleontologists and Mineralogists, 37, 339–59.Google Scholar
Stel, H. (1985). Crystal growth in cataclasites: diagnostic microstructures and implications. Tectonophysics, 78, 585600.CrossRefGoogle Scholar
Stephens, B. P. (2001). Basement controls on hydrocarbon systems, depositional pathways, and exploration plays beyond the Sigsbee Escarpment in the central Gulf of Mexico. In Petroleum Systems of Deep-water Basins: Global and Gulf of Mexico Experience, ed. Fillon, R. H., Rosen, N. C., Weimer, P., Lowrie, A., Pettingill, H., Phair, R. L., Roberts, H. H., and van Hoorn, B.. Program and Abstracts – Society of Economic Paleontologists, Gulf Coast Section, Research Conference, 21, 129–58.Google Scholar
Stephenson, L. P. (1977). Porosity dependence on temperature: limits on maximum possible effect. American Association of Petroleum Geologists Bulletin, 61, 407–15.Google Scholar
Stephenson, M. A. (1991). The North Brae Field, Block 16/07a, U.K. North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. I. L. Abbotts. Memoir of the Geological Society of London, 14, 43–8.Google Scholar
Stephenson, R. and Lambeck, K. (1985). Erosion-isostatic rebound models for uplift: an application to south-eastern Australia. Geophysical Journal of the Royal Astronomical Society, 82, 3155.CrossRefGoogle Scholar
Stevens, D. A. and Wallis, R. J. (1991). The Clyde Field, Block 30/17b, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. I. L. Abbotts. Memoir of the Geological Society of London, 14, 339–45.CrossRefGoogle Scholar
Stewart, D. M. and Faulkner, A. J. G. (1991). The Emerald Field, Blocks 2/10a, 2/15a, 3/11b, U.K. North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. I. L. Abbotts. Memoir of the Geological Society of London, 14, 111–16.Google Scholar
Stewart, I. J. (1993). Structural controls on the Late Jurassic age shelf system, Ula trend, Norwegian North Sea. In Petroleum Geology of Northwest Europe. Proceedings of the 4th Conference, ed. J. R. Parker. Geological Society of London, pp. 469–83.CrossRefGoogle Scholar
Stewart, I. S. and Hancock, P. L. (1988). Normal fault zone evolution and fault scarp degradation in the Aegean region. Basin Research, 1, 139–53.CrossRefGoogle Scholar
Stewart, J. H. (1983). Extensional tectonics in the Death Valley area, California: transport of the Panamint Range structural block 80 km northwestward. Geology, 11, 153–7.2.0.CO;2>CrossRefGoogle Scholar
Stewart, K., Turner, S., Kelley, S., Hawkesworth, C.J., Kirstein, L., and Mantovani, M. (1996). 3-D, 40Ar/39Ar geochronology in the Parana continental flood basalt province. Earth and Planetary Science Letters, 143, 95109.CrossRefGoogle Scholar
Stewart, S. A. (1996). Tertiary extensional fault systems on the western margin of the North Sea Basin. Petroleum Geoscience, 2, 167–76.CrossRefGoogle Scholar
Stewart, S. A. and Clark, J. A. (1999). Impact of salt on the structure of the Central North Sea hydrocarbon fairways. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, The Geological Society of London, pp. 179200.Google Scholar
Stewart, S. A., Fraser, S. I., Cartwright, J. A., Clark, J. A., and Johnson, H. D. (1999). Controls on Upper Jurassic sediment distribution in the Durward-Dauntless area, Blocks 21/11, 21/16. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, The Geological Society of London, pp. 879–96.Google Scholar
Stewart, S. A. and Reeds, A. (2003). Geomorphology of kilometer-scale extensional fault scarps: factors that impact seismic interpretation. American Association of Petroleum Geologists Bulletin, 87, 251–72.CrossRefGoogle Scholar
Stierman, D. J. and Williams, A. E. (1984). Hydrologic and geochemical properties of the San Andreas Fault at the Stone Canyon well. Pure and Applied Geophysics, 122, 403–24.Google Scholar
St John, W. (2000). The role of transform faulting in formation of hydrocarbon traps in the Gulf of Guinea, West Africa. Offshore West Africa Conference, Abidjan, March 21–23, 2000.Google Scholar
Stockbridge, C. P. and Gray, D. I. (1991). The Fulmar Field, Blocks 30/16 and 30/11b, UK North Sea. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, The Geological Society of London, 14, 309–16.Google Scholar
Stockli, D. F., Dumitru, T. A., McWilliams, M. O., and Farley, K. A. (2003). Cenzoic tectonic evolution of the White Mountains, California and Nevada. Geological Society of America Bulletin, 115, 788816.2.0.CO;2>CrossRefGoogle Scholar
Stoffers, P., Glasby, G. P., Stuben, D., Renner, R. M., Pierre, T. G., Webb, J., and Cardile, C. M. (1993). Comparative mineralogy and geochemistry of hydrothermal iron-rich crusts from the Pitcairn, Teahitia-Migetia and Macdonald hot spot areas of the S. W. Pacific. Marine Georesources and Geotechnology, 11, 4586.CrossRefGoogle Scholar
Stoll, R. D. and Bryan, G. M. (1979). Physical properties of sediments containing gas hydrates. Journal of Geophysical Research, 84, 1629–34.CrossRefGoogle Scholar
Storey, M., Kent, R. W., Saunders, A. D., Hergt, J., Salters, V. J. M., Whitechurch, H., Sevigny, J. H., Thirlwall, M. F., Leat, P., Ghose, N. C., and Gifford, M. (1992). Scientific Results, Ocean Driling Program, Leg 120: Lower Cretaceous volcanic rocks on continental margins and their relationship to the Kerguelen Plateau: College Station, Texas, Ocean Drilling Program, pp. 33–53.Google Scholar
Stow, D. A. V. and Atkin, B. P. (1987). Sediment facies and geochemistry of Upper Jurassic mudrocks in the Central North Sea area. In Petroleum Geology of North West Europe, ed. Brooks, J. and Glennie, K. W.. Graham and Trotman, London, pp. 797808.Google Scholar
Strehlau, J. (2006). Earthquakes and nonvolcanic tremor in the lower continental crust: open questions. EGU annual meeting, Vienna, April 2–7, 2006, Scientific Programme, p. 562.Google Scholar
Strehlau, J. and Stange, S. (2006). Earthquakes in the lower crust under the Northern Alpine foreland basin: seismological detection of active metamorphism? EGU annual meeting, Vienna, April 2–7, 2006, Scientific Programme, p. 562.Google Scholar
Streit, J. E. (1997). Low frictional strength of upper crustal faults: a model. Journal of Geophysical Research, 102, 24619–26.CrossRefGoogle Scholar
Strong, G. E. and Milodowski, A. E. (1987). Aspects of the diagenesis of the Sherwood sandstones of the Wessex Basin and their influence on reservoir characteristics. In Diagenesis of Sedimentary Sequences, ed. Marshall, J. D.. Geological Society Special Publications, pp. 325–37.Google Scholar
Stuart, C. J., Nemčok, M., Vangelov, D., Higgins, E. R., Welker, C., and Meaux, D. (2011). Deposition/tectonic interaction in the active and subsequently extensionally dismembered thrustbelt; Eastern Balkans, Bulgaria. American Association of Petroleum Geologists Bulletin, 95, 649–73.Google Scholar
Stuiver, M., Reimer, P. J., and Braziunas, T. F. (1998). High-precision radiocarbon age calibration for terrestrial and marine samples. Radiocarbon, 40, 1127–51.CrossRefGoogle Scholar
Stüwe, K. L., White, L., and Brown, R. (1994). The influence of eroding topography on steady-state isotherms. Applications to fission track analysis. Earth and Planetary Science Letters, 124, 6374.CrossRefGoogle Scholar
Suardy, A., Simbolon, B., and Taruno, P. J. (1987). Two decades of hydrocarbon exploration activities in Indonesia. Transactions of the Circum-Pacific Energy and Mineral Resources Conference, 4, 243–61.Google Scholar
Subrahmanyam, C., Thakur, N. K., Gangadhara Rao, T., Ramesh, K., Ramana, M. V., and Subrahmanyam, V. (1999). Tectonics of the Bay of Bengal: new insights from satellite-gravity and ship-borne geophysical data. Earth and Planetary Science Letters, 171, 237–51.CrossRefGoogle Scholar
Suchy, V., Heijlen, W., Sykorova, I., Muchez, P., Dobes, P., Hladikova, J., Jackova, I., Safanda, J., and Zeman, A. (2000). Geochemical study of calcite veins in the Silurian and Devonian of the Barrandien Basin (Czech Republic): evidence for widespread post-Variscan fluid flow in the central part of the Bohemian Massif. Sedimentary Geology, 131, 201–19.CrossRefGoogle Scholar
Suppe, J. (1985). Principles of Structural Geology. Prentice Hall, Englewood Cliffs, New Jersey.Google Scholar
Surdam, R. C., Crossey, L. J., Hagen, E. S., and Heasler, H. P. (1989). Organic-inorganic and sandstone diagenesis. American Association of Petroleum Geologists Bulletin, 73, 123.Google Scholar
Surdam, R. C., Jiao, Z. S., and MacGowan, D. B. (1993). Redox reactions involving hydrocarbon and mineral oxidants: a mechanism for significant porosity enhancement in sandstones. American Association of Petroleum Geologists Bulletin, 77, 1509–18.Google Scholar
Surlyk, F., Dons, T., Clausen, C. K., and Higham, J. (2002). Chapter 13. Upper Cretaceous. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. Geological Society of London, pp. 490547.Google Scholar
Swanberg, C. A. (1972). Vertical distribution of heat generation in the Idaho batholith. Journal of Geophysical Research, 77, 2508–13.CrossRefGoogle Scholar
Sweeney, J. J. and Burnham, A. K. (1990). Evaluation of a simple model of vitrinite reflectance based on chemical kinetics. American Association of Petroleum Geologists Bulletin, 74, 1559–70.Google Scholar
Sweeney, J. J., Burnham, A. K., and Braun, R. L. (1987). A model of hydrocarbon generation from type I kerogen: application to Uinta Basin, Utah. American Association of Petroleum Geologists Bulletin, 71, 967–85.Google Scholar
Sylvester, A. G. (1988). Strike-slip faults. Geological Society of America Bulletin, 100, 1666–703.2.3.CO;2>CrossRefGoogle Scholar
Syms, R. M., Savory, D. F., Ward, C. J., Ebdon, C. C., and Griffin, A. (1999). Integrating sequence stratigraphy in field development and reservoir management – the Telford Field. In Petroleum Geology of Northwest Europe, ed. Fleet, A. J. and Boldy, S. A. R.. Proceedings of the 5th Conference, The Geological Society of London, pp. 1075–87.Google Scholar
Szafián, P. (1999). Gravity and Tectonics: A Case Study in the Pannonian Basin and the Surrounding Mountain Belt. Vrije University, Amsterdam, p. 154.Google Scholar
Tagami, T. and O’Sullivan, P. B. (2005). Fundamentals of fission-track thermochronology. In Low-Temperature Thermochronology: Techniques, Interpretations and Applications, ed. Reinersand, P. W. and Ehlers, T. A.. Reviews in Mineralogy and Geochemistry, 58, 1948.CrossRefGoogle Scholar
Tait, S. R., Huppert, H. E., and Sparks, R. S. J. (1984). The role of compositional convection in the formation of accumulate rocks. Lithos, 17, 139–46.CrossRefGoogle Scholar
Tait, S. R. and Kerr, R. C. (1987). Experimental modelling of interstitial melt convection in cumulus piles. In Origins of Igneous Layering, ed. Parson, I.. NATO ASI Series C, 196, 569–87.Google Scholar
Talbot, C. J. and Ghebreab, W. (1997). Red Sea detachment and basement core complexes in Eritrea. Geology, 25, 655–8.2.3.CO;2>CrossRefGoogle Scholar
Talbot, M. R. (1988). The origins of lacustrine oil source rocks: evidence from the lakes of tropical Africa. In Lacustrine Petroleum Source Rocks, ed. Fleet, A. J., Kelts, K., and Talbot, M. R.. Geological Society Special Publication, 40, 2943.Google Scholar
Talwani, M., Windisch, C. C., and Langseth, M. G. Jr. (1971). Reykjanes Ridge crest: a detailed geophysical study. Journal of Geophysical Research, 76, 473517.CrossRefGoogle Scholar
Tankard, A. J. and Welsink, H. J. (1987). Extensional tectonics and stratigraphy of Hibernia oil field, Grand Banks, Newfoundland. American Association of Petroleum Geologists Bulletin, 71, 1210–32.Google Scholar
Tankard, A. J. and Welsink, H. J. (1989). Mesozoic extension and styles of basin formation in Atlantic Canada. American Association of Petroleum Geologists Memoir, 46, 175–95.Google Scholar
Tankard, A. J., Welsink, H. J., and Jenkins, W. A. M. (1989). Structural styles and stratigraphy of the Jeanne d’Arc Basin, Grand Banks of Newfoundland. American Association of Petroleum Geologists Memoir, 46, 265–82.Google Scholar
Tard, F., Masse, P., Walgenwitz, F., and Gruneisen, P. (1991). The volcanic passive margin in the vicinity of Aden, Yemen. Bulletin des centres de recherche et d’exploration-production Elf-Aquitaine, 15, 19.Google Scholar
Tari, G., Ashton, P., Coterill, K., Molnar, J., Sorgenfrei, M., Thompson, W. A. P., Valasek, D., and Fox, J. (2002). Are West Africa deepwater salt tectonics analogous to the Gulf of Mexico? Oil and Gas Journal, 4, 116.Google Scholar
Tari, G. and Horváth, F. (2006). Alpine evolution and hydrocarbon geology of the Pannonian Basin: an overview. In The Carpathians and Their Foreland: Geology and Hydrocarbon Resources, ed. Golonka, J. and Pícha, F. J.. American Association of Petroleum Geologists Memoir, 84, 605–18.Google Scholar
Tari, G., Horváth, F., and Rumpler, J. (1992). Styles of extension in the Pannonian Basin. Tectonophysics, 208, 203–19.CrossRefGoogle Scholar
Tari, G., Molnar, J., Ashton, P., and Hedley, R. (2000). Salt tectonics in the Atlantic margin of Morocco. Leading Edge, 19, 1074–76.CrossRefGoogle Scholar
Tari, G., Molnar, J., Ashton, P., and Hedley, R. (2001). Exploration in syn-rift versus post-rift salt basins of West Africa: Are there significant differences? Annual Meeting Expanded Abstracts – American Association of Petroleum Geologists, p. 198.Google Scholar
Tavenas, F., Jean, P., Leblond, P., and Leroueil, S. (1983). The permeability of natural soft clays. Part 2: Permeability characteristics. Canadian Geotechnical Journal, 20, 645–60.Google Scholar
Tayebi, M. (1989). Le segment hercynien du haut-Atlas occidental dans les Ait Chaib, Maroc. Stratigraphie, tectonique et role de la zone failee ouest altasique. PhD thesis, University of d’Aix-Marseille III, p. 234.Google Scholar
Taylor, A. M. and Gawthorpe, R. L. (1993). Application of sequence stratigraphy and trace fossil analysis to reservoir description: examples from the Jurassic of the North Sea. In Petroleum Geology of Northwest Europe, ed. J. R. Parker. Proceedings of the 4th Conference, The Geological Society of London, pp. 317–35.CrossRefGoogle Scholar
Taylor, B., Crook, K., and Sinton, J. (1994). Extensional transform zones and oblique spreading centers. Journal of Geophysical Research, 99, 19707–18.CrossRefGoogle Scholar
Taylor, G. H., Teichmueller, M., Davis, A., Diessel, C. F. K., Littke, R., Robert, P., Glick, D. C., Smyth, M., Swaine, D. J., Vandebroucke, M., and Espitalie, J. (1998). Organic Petrology. Gebrueder Borntraeger, Berlin, p. 704.Google Scholar
Tegelaar, E. W., Matthezing, R. M., Jansen, J. B. H., Horsfield, B., and de Leeuw, J. W. (1989). Possible origin of n-alkanes in high-wax crude oils. Nature, 342, 529–31.CrossRefGoogle Scholar
Tegelaar, E. W. and Noble, T. A. (1994). Kinetics of hydrocarbon generation as a function of the molecular structure of kerogen as revealed by pyrolysis-gas chromatography. Organic Geochemistry, 22, 543–74.CrossRefGoogle Scholar
Teisserenc, P. and Villemin, J. (1990). Sedimentary basin of Gabon-geology and oil systems. In Divergent/Passive Margin Basins, ed. Edwards, J. D. and Santogrossi, P. A.. American Association of Petroleum Geologists Memoir, 48, 117–99.Google Scholar
Ten Brink, U. S. and Ben-Avraham, Z. (1989). The anatomy of a pull-apart basin: seismic reflection observations of the Dead Sea basin. Tectonics, 8, 333–50.CrossRefGoogle Scholar
Ten Brink, U. and Stern, T. (1992). Rift flank uplifts and hinterland basins: comparison of the Transatlantic Mountains with the Great Escarpment of southern Africa. Journal of Geophysical Research, 97, 569–85.CrossRefGoogle Scholar
Ter Voorde, M. and Bertotti, G. (1994). Thermal effects of normal faulting during rifted basin formation: 1, A finite difference model. Tectonophysics, 240, 133–44.CrossRefGoogle Scholar
TGS-NOPEC (2003). Problems in exploration – Part 1. Need to find faults? American Association of Petroleum Geologists Explorer, 11.Google Scholar
Thatcher, W. R., Foulger, G. R., Julian, B. R., Svarc, J. L., and Quilty, E. G. (1998). Present day deformation across the Basin and Range Province, Western United States. Eos, Transactions, American Geophysical Union, 79, 559.Google Scholar
Theil, K. (1995). Wlasciwosci fizyko-mechaniczne i modele masywow skalnych polskich Karpat fliszowych. Gdansk, Institut Budownictwa Wodnego, Polska Akademia Nauk.Google Scholar
Thomas, D. W. (1996a). Tectonic evolution and kinematics of the Mesozoic rift system, UK Northern North Sea. PhD thesis, University of London.Google Scholar
Thomas, D. W. (1996b). Mesozoic regional tectonics and South Viking Graben formation: evidence for localised thin-skinned detachments during rift development and inversion. Marine and Petroleum Geology, 13, 149–77.CrossRefGoogle Scholar
Thomas, M. M. and Clouse, J. A. (1995). Scaled physical model of secondary oil migration. American Association of Petroleum Geologists Bulletin, 79, 1929.Google Scholar
Thompson, A. H., Katz, A. J., and Krohn, C. E. (1987). The microgeometry and transport properties of sedimentary rock. Advances in Physics, 36, 625–94.CrossRefGoogle Scholar
Thompson, G. and Melson, W. G. (1972). The petrology of oceanic crust across fracture zones in the Atlantic Ocean: evidence of a new kind of sea-floor spreading. Journal of Geology, 80, 526–38.CrossRefGoogle Scholar
Thompson, K. F. M. (1987). Fractionated aromatic petroleums and the generation of gas-condensates. Organic Geochemistry, 18, 573–90.Google Scholar
Thompson, K. F. M. (1988). Gas-condensate migration and oil fractionation in deltaic systems. Marine and Petroleum Geology, 5, 237–46.CrossRefGoogle Scholar
Thompson, R. N. (1974). Primary basalts and magma genesis; I, Skye, North-West Scotland. Contributions to Mineralogy and Petrology, 45, 317–41.CrossRefGoogle Scholar
Thompson, R. N. and Morrison, M. A. (1988). Asthenospheric and lower-lithospheric mantle contributions to continental extensional magmatism: an example from the British Tertiary Province. Chemical Geology, 68, 115.CrossRefGoogle Scholar
Thompson, S., Cooper, B. S., and Barnard, P. C. (1994). Some examples and possible explanations for oil generation from coals and coaly sequences. In Coal and Coal-bearing Strata as Oil-prone Source Rocks, ed. Scott, A. C. and Fleet, A. J.. Geological Society of London Special Publications, 77, 119–37.Google Scholar
Thompson, V. (2009). Potential-field and 2D seismic analysis of a volcanic rifted margin: implications for crustal architecture and petroleum maturation off the west coast of South Africa. Master´s thesis, University of Utah, Salt Lake City, p. 142.Google Scholar
Thurber, C., Roecker, S., Ellsworth, W., Chen, Y., and Lutter, W. (1997). Two-dimensional seismic image of the San Andreas Fault in the Northern Gabilan Range, central California: Evidence from fluids in the fault zone. Geophysical Research Letters, 24, 1591–4.CrossRefGoogle Scholar
Thurber, C., Roecker, S., Roberts, K., Gold, M., Powell, L., and Rittger, K. (2003). Earthquake locations and three-dimensional fault zone structure along the creeping section of the San Andreas fault near Parkfield, CA: preparing for SAFOD. Geophysical Research Letters, 30.CrossRefGoogle Scholar
Thurber, C., Roecker, S., Zhang, H., Baher, S., and Ellsworth, W. L. (2004). Fine-scale structure of the San Andreas fault zone and location of the SAFOD target earthquakes. Geophysical Research Letters, 31.CrossRefGoogle Scholar
Tiercelin, J. J. and Faure, H. (1978). Rates of sedimentation and vertical subsidence in neorifts and paleorifts. In Tectonics and Geophysics of Continental Rifts, ed. Ramberg, I. B. and Neumann, E. R.. Reidel Publishing Company, Dordrecht, pp. 41–7.Google Scholar
Tigre, C. A., Schaller, H., Del Luchese, C. Jr., and Possato, S. (1983). Pampo, Linguado, and Badejo fields: their discoveries, appraisals, and early production systems. Proceedings of the Offshore Technology Conference, 15, 407–14.Google Scholar
Tillman, L. E. (1989). Sedimentary facies and reservoir characteristics of the Nugget Sandstone (Jurassic), Painter Reservoir Field, Uinta County, Wyoming. In Petrogenesis and Petrophysics of Selected Sandstone Reservoirs of the Rocky Mountain Region, ed. Coalson, E. B., Kaplan, S. S., Keighin, C. W., Oglesby, C. A., and Robinson, J. W.. Rocky Mountain Association of Geologists, pp. 97108.Google Scholar
Tissot, B. P., Durand, B., Espitalié, J., and Combaz, A. (1974). Influence of nature and diagenesis of organic matter in formation of petroleum. American Association of Petroleum Geologists Bulletin, 58, 499506.Google Scholar
Tissot, B. P. and Espitalié, J. (1975). The thermal evolution of organic matter in sediments: application of a mathematical model simulation. Revue de l’Institut Français du Pétrol, 30, 743–77.Google Scholar
Tissot, B. P., Pelet, R., and Ungerer, P. (1987). Thermal history of sedimentary basins, maturation indices, and kinetics of oil and gas generation. American Association of Petroleum Geologists Bulletin, 71, 1445–66.Google Scholar
Tissot, B. P. and Welte, D. H. (1984). Petroleum Formation and Occurrence, a New Approach to Oil and Gas Exploration. Springer-Verlag, New York.Google Scholar
Toksoz, M. N., Cheng, C. H., and Timur, A. (1976). Velocities of seismic waves in porous rocks. Geophysics, 41, 621–43.CrossRefGoogle Scholar
Tokunaga, T., Mogi, K., Matsubara, O., Tosaka, H., and Kojima, K. (2000). Buoyancy and interfacial force effects on two-phase displacement patterns: an experimental study. American Association of Petroleum Geologists Bulletin, 84, 6574.Google Scholar
Tomek, Č., Dvořáková, L., and Ibrmajer, I. (1987). Crustal profiles of active continental collisional belt: Czechoslovak deep seismic profiling in the West Carpathians. Geophysical Journal of the Royal Astronomical Society, 89, 383–8.Google Scholar
Tomek, C., Ibrmajer, I., and Korab, T. (1989). Kôrové štruktúry Západních Karpát na hlubinném reflexním seizmickém profilu. Crustal structure of the Western Carpathians as determined by deep seismic sounding. Mineralia Slovaca, 21, 326.Google Scholar
Tomic, J., Behar, F., Vandenbroucke, M., and Tang, Y. (1995). Artificial maturation of Monterey kerogen (Type II-S) in a closed system and comparison with Type II kerogen: implications on the fate of sulfur. Organic Geochemistry, 23, 647–60.CrossRefGoogle Scholar
Torgensen, T. and Clarke, W. B. (1992). Geochemical constraints on formation fluid ages, hydrothermal heat flux, and crustal mass transport mechanisms at Cajon Pass. Journal of Geophysical Research, 97, 5031–8.Google Scholar
Toulokian, Y. S., Judd, W. R., and Roy, R. F. (1981). Physical properties of rocks and minerals. McGraw-Hill, New York, p. 548.Google Scholar
Trewin, N. H. and Bramwell, M. G. (1991). The Auk field, Block 30/16 UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 227–36.Google Scholar
Trinidade, L. A. F., Dias, J. L., and Mello, M. R. (1995). Sedimentological and geochemical characterization of the Lagoa Feia Formation, rift phase of the Campos Basin. In Petroleum Source Rocks, ed. Katz, B. J.. Springer-Verlag, Berlin, pp. 149–65.Google Scholar
Tron, V. and Brun, J.-P. (1991). Experiments on oblique rifting in brittle-ductile systems. Tectonophysics, 188, 7184.CrossRefGoogle Scholar
Trudgill, B. D. (2002). Structural controls on drainage development in the Canyonlands grabens of Southeast Utah. American Association of Petroleum Geologists Bulletin, 86, 1095–112.Google Scholar
Trudgill, B. and Cartwright, J. (1994). Relay-ramp forms and normal-fault linkages, Canyonlands National Park, Utah. Geological Society of America Bulletin, 106, 1143–211.2.3.CO;2>CrossRefGoogle Scholar
Trumbull, R. B., Sobolev, S. V., and Bauer, K. (2002). Petrophysical modeling of high seismic velocity crust at the Namibian volcanic margin. In Volcanic Rift Margins, ed. Menzies, M. A., Klemperer, S. L., , C, Ebinger, , and Baker, J.. Geological Society of America Special Paper, 362, pp. 221–30.Google Scholar
Tsenn, M. C. and Carter, N. L. (1987). Upper limits of power law creep of rocks. Tectonophysics, 136, 126.CrossRefGoogle Scholar
Tucholke, B. E., Lin, J., and Kleinrock, M. C. (1998). Megamullions and mullion structure defining oceanic metamorphic core complexes on the Mid-Atlantic Ridge. Journal of Geophysical Research, 103, 9857–66.CrossRefGoogle Scholar
Tucker, G. E. and Slingerland, R. (1996). Predicting sediment flux from fold and thrust belts. Basin Research, 8, 329–49.CrossRefGoogle Scholar
Turcotte, D. L. and Emerman, S. H. (1983). Mechanisms of active and passive rifting. Tectonophysics, 94, 3950.CrossRefGoogle Scholar
Turcotte, D. L. and Schubert, G. (1982). Geodynamics: Applications of Continuum Physics to Geological Problems. John Wiley and Sons, New York, p. 450.Google Scholar
Turcotte, D. L. and Schubert, G. (2002). Geodynamics, Second edition. Cambridge University Press, Cambridge, p. 456.CrossRefGoogle Scholar
Turner, C. C. and Allen, P. J. (1991). The Central Brae Field, Block 16/07a, U.K. North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. I. L. Abbotts. Memoir of the Geological Society of London, 14, 49–54.Google Scholar
Turner, C. C., Cohen, J. M., Connell, E. R., and Cooper, D. M. (1987). A depositional model for the South Brae oilfield. Proceedings of the Third Canadian/American Conference on Hydrogeology, June 22–6, 1986, Bannf, Alberta, pp. 853–64.Google Scholar
Turner, C. C. and Connell, E. R. (1991). Stratigraphic relationships between Upper Jurassic submarine fan sequences in the Braere, U.K. North Sea: the implications for reservoir distribution. 23rd Annual Offshore Technology Conference, Dallas, pp. 83–91.Google Scholar
Turner, F. J. and Weiss, L. E. (1963). Structural Analysis of Metamorphic Tectonites. New York, McGraw-Hill, p. 545.Google Scholar
Turner, P. J. (1993). Clyde: reappraisal of a producing field. In Petroleum Geology of Northwest Europe, ed. Parker, J. R.. Proceedings of the 4th Conference, Geological Society of London, pp. 1503–12.Google Scholar
Turner, S., Regelous, M., Kelley, S., Hawkesworth, C., and Mantovani, M. (1994). Magmatism and continental break-up in the South Atlantic: high precision 40Ar-39Ar geochronology. Earth and Planetary Science Letters, 121, 333–48.CrossRefGoogle Scholar
Tuttle, M. L. W., Charpentier, R. R., and Brownfield, M. E. (1999). The Niger Delta petroleum system: Niger Delta province, Nigeria, Cameroon, and Equatorial Guinea, Africa. USGS Open-File Report, 99-50-H, p. 65.CrossRefGoogle Scholar
Tye, R. S. (1991). Fluvial-sandstone reservoirs of the Travis Peak Formation, East Texas Basin. Concepts in Sedimentology and Paleontology, 3, 172–88.Google Scholar
Tye, R. S. (2004a). Geomorphology; an approach to determining subsurface reservoir dimensions. American Association of Petroleum Geologists Bulletin, 88, 1123–47.CrossRefGoogle Scholar
Tye, R. S. (2004b). Reservoir description and unique horizontal-well designs boost. Primary and EOR production from the fluvio-deltaic Prudhoe Bay field, Alaska. AAPG Distinguished Lecture Funded by the AAPG Foundation in honor of Roy M. Huffington.CrossRefGoogle Scholar
Tye, R. S. and Coleman, J. M. (1989a). Depositional processes and stratigraphy of fluvially dominated lacustrine deltas: Mississippi delta plain. Journal of Sedimentary Petrology, 59, 973–96.Google Scholar
Tye, R. S. and Coleman, J. M. (1989b). Evolution of Atchafalaya lacustrine deltas, south-central Louisiana. Sedimentary Geology, 65, 95112.CrossRefGoogle Scholar
Uenzelmann Neben, G., Siess, V., and Bleil, U. (1997). A seismic reconnaissance survey of the northern Congo fan. Marine Geology, 140, 283306.CrossRefGoogle Scholar
Ukstins, I., Renne, P., Wolfenden, E., Baker, J., Ayalew, D., and Menzies, M. A. (2002). Matching conjugate volcanic rifted margins: 40Ar/39Ar chronostratigraphy of pre- and syn-rift bimodal flood volcanism in Ethiopia and Yemen. Earth and Planetary Science Letters, 198, 289306.CrossRefGoogle Scholar
Ulmishek, G. F. and Klemme, H. D. (1990). Depositional controls, distribution, and effectiveness of world’s petroleum source rocks. U. S. Geological Survey Bulletin, 1931.Google Scholar
Underhill, J. R. (1991a). Implications of Mesozoic – recent basin development in the Inner Moray Firth, UK. Marine and Petroleum Geology, 8, 359–69.CrossRefGoogle Scholar
Underhill, J. R. (1991b). Controls on Late Jurassic seismic sequences, Inner Moray Firth: a critical test of Exxon’s original sea-level chart. Basin Research, 3, 7998.CrossRefGoogle Scholar
Underhill, J. R. (1994). Discussion on palaeoecology and sedimentology across a Jurassic fault scarp, NE Scotland. Journal of the Geological Society of London, 151, 729–31.Google Scholar
Underhill, J. R. and Partington, M. A. (1993). Jurassic thermal doming and deflation in the North Sea: implication of the sequence stratigraphic evidence. In Petroleum Geology of Northwest Europe, ed. Parker, J. R.. Proceedings of the 4th Conference, The Geological Society of London, pp. 337–46.Google Scholar
Underhill, J. R., Sawyer, M. J., Hodgson, P., Shallcross, M. D., and Gawthorpe, R. L. (1997). Implications of fault scarp degradation for Brent Group prospectivity, Ninian Field, northern North Sea. American Association of Petroleum Geologists Bulletin, 81, 9991022.Google Scholar
Underwood, M. B. and Howell, D. G. (1987). Thermal maturity of the Cambria Slab, an inferred trench-slope basin in Central California. Geology, 15, 216–19.2.0.CO;2>CrossRefGoogle Scholar
Ungerer, P. (1990). State of the art of research in kinetic modeling of oil formation and expulsion. Organic Geochemistry, 16, 125.CrossRefGoogle Scholar
Ungerer, R., Behar, F., and Discamp, D. (1983). Tentative calculation of the overall volume expansion of organic matter during hydrocarbon genesis from geochemistry data. Implications for primary migration. In Advances in Organic Geochemistry 1981, ed. Bjoroy, M., Albrecht, P., Cornford, C., de Groot, K., Elington, G., and Galimov, E.. John Wiley, Chichester, pp. 129–35.Google Scholar
Ungerer, P., Burrus, J., Doligez, B., Chenet, P. Y., and Bessis, F. (1990). Basin evaluation by integrated two-dimensional modeling of heat transfer, fluid flow, hydrocarbon generation, and migration. American Association of Petroleum Geologists Bulletin, 74, 309–35.Google Scholar
Ungerer, P. and Pelet, R. (1987). Extrapolation of the kinetics of oil and gas formation from laboratory to sedimentary basin. Nature, 327, 52–4.CrossRefGoogle Scholar
United States Geological Survey (USGS) (1996). Global 30-Arc-Second Elevation Data Set. United States Geological Survey, Sioux Falls, South Dakota.Google Scholar
United States Geological Survey (USGS) (2006). Quaternary Fault and Fold Database of the United States. http://earthquake.usgs.gov/regional/qfaults.Google Scholar
Unsworth, M. and Bedrosian, P. A. (2004). Electrical resistivity structure at the SAFOD site from magnetotelluric exploration. Geophysical Research Letters, 31.CrossRefGoogle Scholar
Unsworth, M., Bedrosian, P. A., Eisel, M., Egbert, G., and Siripunvaraporn, W. (2000). Along strike variations in the electrical structure of the San Andreas Fault at Parkfield, California. Geophysical Research Letters, 27, 3021–4.CrossRefGoogle Scholar
Unsworth, M., Egbert, G., and Booker, J. (1999). High- resolution electromagnetic imaging of the San Andreas fault in Central California. Journal of Geophysical Research, 104, 1131–50.CrossRefGoogle Scholar
Unsworth, M., Malin, P. E., Egert, G. D., and Booker, J. R. (1997). Internal structure of the San Andreas fault at Parkfield, California. Geology, 25, 359–62.2.3.CO;2>CrossRefGoogle Scholar
Unternehr, P., Curie, D., Olivet, J. L., Goslin, J., and Beuzart, P. (1988). South Atlantic fits and intraplate boundaries in Africa and South America. Tectonophysics, 155, 169–79.CrossRefGoogle Scholar
Urgeles, R., Canals, M., Baraza, J., and Alonso, B. (1998). Seismostratigraphy of the western flanks of El Hierro and La Palma (Canary Islands): record of Canary Island volcanism. Marine Geology, 146, 225–41.CrossRefGoogle Scholar
Urgeles, R., Canals, M., Baraza, J., Alonso, B., and Masson, D. (1997). The most recent megalandslides of the Canary Islands: El Golfo debris avalanche and Canary debris flow, West El Hierro Island. Journal of Geophysical Research, 102, 20305–23.CrossRefGoogle Scholar
Urgeles, R., Canals, M., Roberts, J., and the SNV “Las Palmas” Shipboard party (2000). Fluid flow from pore pressure measurements off La Palma, Canary Islands. Journal of Volcanology and Geothermal Research, 101 , 253–71.CrossRefGoogle Scholar
Urgeles, et al. (1999). Fluid flow from pore pressure measurements off La Palma, Canary Islands. Journal of Volcanology and Geothermal Research, 94(1–4), 305–21.CrossRefGoogle Scholar
Vagnes, E., Boavida, J., Jeronimo, P., de Brito, M., and Peliganga, J. M. (2005). Crustal architecture of West African rift basins in the deep water province. In Petroleum Systems of Divergent Continental Margin Basins, ed. P. Post. 25th Gulf Coast Section, Bob F. Perkins Research Conference, 4–7 December 2005, CD-ROM.CrossRefGoogle Scholar
Vail, P. R. (1987). Seismic stratigraphy interpretation using sequence stratigraphy: Part 1, Seismic stratigraphy interpretation procedure. American Association of Petroleum Geologists Studies in Geology, 27, 110.Google Scholar
Vail, P. R., Mitchum, R. M. Jr., and Thompson, S. (1977). Seismic stratigraphy and global changes of sea level: Part 3, Relative changes of sea level from coastal onlap. Seismic stratigraphy: applications to hydrocarbon exploration. American Association of Petroleum Geologists, pp. 6381.Google Scholar
Vail, P. R. and Todd, R. G. (1981). Cause of northern North Sea Jurassic unconformities. American Association of Petroleum Geologists Bulletin, 65, 1003.Google Scholar
Vail, P. R., Todd, R. G., and Sangree, J. B. (1977). Seismic stratigraphy and global changes of sea level: Part 5. Chronostratigraphic significance of seismic reflections: Section 2. Application of Seismic Reflection Configuration to Stratigraphic Interpretation Memoir, 26, 99116.Google Scholar
Van Avendonk, H. J. A., Holbrook, W. S., Nunes, G. T., Shillington, D. J., Tucholke, B. E., Louden, K. E., Larsen, H. C., and Hopper, J. R. (2006). Seismic velocity structure of the rifted margin of the eastern Grand Banks of Newfoundland, Canada. Journal of Geophysical Research, 111, B11404.CrossRefGoogle Scholar
Van Avendonk, H. J. A., Lavier, L. L., Shillington, D. J., and Manatschal, G. (2009). Extension of continental crust at the margin of the eastern Grand Banks, Newfoundland. Tectonophysics, 468, 131–48.CrossRefGoogle Scholar
Van Balen, R. T., Van der Beek, P. A., and Cloetingh, S. A. P. L. (1995). The effect of rift shoulder erosion on stratal patterns at passive margins: implications for sequence stratigraphy. Earth and Planetary Science Letters, 134, 527–44.CrossRefGoogle Scholar
Van der Beek, P., Andriessen, P., and Cloetingh, S. (1995). Morphotectonic evolution of rifted continental margins: inferences from a coupled tectonic-surface processes model and fission track thermochronology. Tectonics, 14, 406–21.CrossRefGoogle Scholar
Van der Beek, P., Cloetingh, S., and Andriessen, P. (1994). Mechanisms of extensional basin formation and vertical motions at rift flanks: constraints from tectonic modelling and fission-track thermochronology. Earth and Planetary Science Letters, 121, 417–33.CrossRefGoogle Scholar
Van den Beukel, P. J. (1990). Thermal and mechanical modelling of convergent plate margins. Geologica Ultraiectina, 62, 126.Google Scholar
Van der Helm, A. A., Gray, D. I., Cook, M. A., and Schulte, A. M. (1990). Fulmar, the development of a large North Sea field. In North Sea Oil and Gas Reservoirs, Vol. 2, ed. Buller, A. T. et al. Proceedings of the North Sea Oil and Gas Reservoirs Conference, Graham and Trotman, London, pp. 2545.CrossRefGoogle Scholar
Van der Kamp, G. and Bachu, S. (1989). Use of dimensional analysis in the study of thermal effects of various hydrogeological regimes. In Hydrogeological Regimes and Their Subsurface Thermal Effects, ed. Beck, A. E., Garven, G., and Stegena, L.. Geophysical Monograph, 47, 23–8.Google Scholar
Van Heijst, M. W. I. M. and Postma, G. (2001). Fluvial response to sea-level changes: a quantitative analogue, experimental approach. Basin Research, 13, 269–92.CrossRefGoogle Scholar
Van Heijst, M. W. I. M., Postma, G., van Kesteren, W. P., and de Jongh, R. G. (2002). Control of syndepositional faulting on systems tract evolution across growth-faulted shelf margins: an analog experimental model of the Miocene Imo River Field, Nigeria. American Association of Petroleum Geologists Bulletin, 86, 1335–66.Google Scholar
Van Schmus, W. R. (1984). Radioactivity properties of minerals and rocks. In Handbook of Physical Properties of Rocks, ed. Carmichael, R. S.. CRC Press, Boca Raton, pp. 281–93.Google Scholar
Van Wagoner, J. C., Posamentier, H. W., Mitchum, R. M., Vail, P. R., Sarg, J. F., Loutit, T. S., and Hardenbol, J. (1988). An overview of the fundamentals of sequence stratigraphy and key definitions. In Sea-Level Changes: An Integrated Approach, ed. Wilgus, C. K., Hastings, B. S., Kendall, C. G. C. St., Posamentier, H. W., Ross, C. A., and Van Wagoner, J. C.. Special Publication – Society of Economic Paleontologists and Mineralogists, pp. 3945.CrossRefGoogle Scholar
Van Wijk, J. W. and Cloetingh, S. A. P. L. (2002). Basin migration caused by slow lithospheric extension. Earth and Planetary Science Letters, 198, 275–88.CrossRefGoogle Scholar
Van Wijk, J. W., Huismans, R. S., ter Voorde, M., and Cloetingh, S. A. P. L. (2001). Melt generation at volcanic continental margins: no need for a mantle plume? Geophysical Research Letters, 28, 3995–8.CrossRefGoogle Scholar
Vandenbroucke, M. (1993). Migration of hydrocarbons. In Applied Petroleum Geochemistry, ed. Bordenave, M. L.. Technip, Paris, pp. 125–48.Google Scholar
Vandenbroucke, M., Behar, F., and Rudkiewicz, J. L. (1999). Kinetic modeling of petroleum formation and cracking: implication from the high pressure/high temperature Elgin Field (UK, North Sea). Organic Geochemistry, 30, 1105–25.CrossRefGoogle Scholar
Vandervoort, D. S. (1997). Stratigraphic response to saline lake-level fluctuations and the origin of cyclic nonmarine evaporite deposits: the Pleistocene Blanca Lila Formation, northwest Argentina. Geological Society of America Bulletin, 109, 210–24.2.3.CO;2>CrossRefGoogle Scholar
Vasseur, G. and Burrus, J. (1990). Contraintes hydrodynamiques et thermiques sur la genese des gisements stratiformes a plomb-zinc. In Mobilite et concentration des metaux de base dans les couvertures sedimentaires, manifestations, mecanismes, prospection; actes du colloque international, ed. Pelissonnier, H. and Sureau, J. F.. Documents – B. R. G. M., 183, 305–54.Google Scholar
Vasseur, G. and Demongodin, L. (1995). Convective and conductive heat transfer in sedimentary basins. Basin Research, 7, 6779.CrossRefGoogle Scholar
Vauchez, A., Dineur, F., and Rudnick, R. L. (2005). Microstructure, texture and seismic anisotropy of the lithospheric mantle above a mantle plume: insights from the Labait volcano xenoliths (Tanzania). Earth and Planetary Science Letters, 232, 295314.CrossRefGoogle Scholar
Vavra, C. L., Kaldi, J. G., and Sneider, R. M. (1992). Geological applications of capillary pressure: a review. American Association of Petroleum Geologists Bulletin, 76, 840–50.Google Scholar
Veenstra, E. (1985). Rift and drift in the Dampier sub-basin, a seismic and structural interpretation. The APEA Journal, 25, 177–89.Google Scholar
Veldkamp, J. J., Gailllard, M. G., Jonkers, H. A., and Levell, B. K. (1996). A Kimmeridgian time-slice through the Humber Group of the central North Sea: a test of sequence stratigraphic methods. In Geology of the Humber Group: Central Graben and Moray Firth, UKCS, ed. Hurst, A., Johnson, H. D., Burley, S. D., Canham, A. C., and Mackertich, D. S.. Special Publication of the Geological Society of London, 114, 128Google Scholar
Vendeville, B., Cobbold, P. R., Davy, P., Brun, J. P., and Choukroune, P. (1987). Physical models of extensional tectonics at various scales. Geological Society of London Special Publications, 28, 95107.CrossRefGoogle Scholar
Vendeville, B. C. and Jackson, M. P. A. (1992). The fall of diapirs during thinskinned extension. Marine and Petroleum Geology, 9, 354–71.Google Scholar
Venkatakrishnan, R. and Lutz, R. (1988). A kinematic model for the evolution of Richmond Triassic basin, Virginia. Developments in Geotectonics, 22, 445–62.CrossRefGoogle Scholar
Verhoef, J. and Srivastava, S. P. (1989). Correlation of sedimentary basins across the North Atlantic as obtained from gravity and magnetic data, and its relation to the early evolution of the North Atlantic. American Association of Petroleum Geologists Memoir, 46, 131–47.Google Scholar
Versfelt, J. A. and Rosendahl, B. R. (1989). Relationships between pre-rift structure and rift architecture in Lakes Tanganyika and Malawi, East Africa. Nature, 337, 354–7.CrossRefGoogle Scholar
Villemin, T., Alvarez, F., and Angelier, J. (1986). The Rhinegraben: extension, subsidence and shoulder uplift. Tectonophysics, 128(1–2), 4759.CrossRefGoogle Scholar
Villemin, T., Angelier, J., and Sunwoo, C. (1995). Fractal distribution of fault length and offsets: implications of brittle deformation evaluation: the Lorraine coal basin. In Fractals in the Earth Sciences, ed. Barton, C. and La Pointe, P. R.. Plenum Press, New York, pp. 205–26.Google Scholar
Vincent, M. W. and Ehlig, P. L. (1988). Laumontite mineralization in rocks exposed north of San Andreas Fault at Cajon Pass, southern California. Geophysical Research Letters, 15, 977–80.CrossRefGoogle Scholar
Volkman, J. K. (1988). Biological marker compounds as indicators of the depositional environments of petroleum source rocks. In Lacustrine Petroleum Source Rocks, ed. Kelts, A. J. and Talbot, M. R.. Geological Society of London Special Publications, 40, 103–22.Google Scholar
Volkman, J. K., Alexander, R., Kagi, P. I., and Woodhouse, G. W. (1983). Demethylated hopanes in crude oils and their applications in petroleum geochemistry. Geochimica et Cosmochimica Acta, 47, 785–94.CrossRefGoogle Scholar
Vollset, J. and Dore, A. G. (1984). A revised Triassic and Jurassic lithostratigraphic nomenclature for the Norwegian North Sea. Norwegian Petroleum Directorate Bulletin, 3.Google Scholar
Von der Dick, H., Meloche, J. D., Dwyer, J., and Gunther, P. (1989). Source-rock geochemistry and hydrocarbon generation in the Jeanne d’Arc Basin, Grand Banks, offshore eastern Canada. Journal of Petroleum Geology, 12, 5168.CrossRefGoogle Scholar
Von Rad, U. and Thurow, J. (1992). Bentonitic clay as indicators of early Neocomian post-breakup volcanism off Northwest Australia: proceedings of the Ocean Drilling Program, Exmouth Plateau, covering Leg 122 of the cruises of the drilling vessel JOIDES Resolution, Singapore, Republic of Singapore, sites 759–764: College Station, Texas, Ocean Drilling Program, pp. 213–32.Google Scholar
Vrolijk, P., Donelick, R. A., Queng, J., and Cloos, M. (1992). Testing models of fission track annealing in apatitie in a simple thermal setting: Site 800. Leg 129. In Proceedings of the Ocean Drilling Program Scientific Results, ed. R. I. Larson, Y. Lancelot, and others. Ocean Drilling Program, College Station, 129, 169–76.Google Scholar
Vutukuri, V. S., Lama, R. D., and Saluja, S. S. (1974). Handbook on Mechanical Properties of Rocks: Testing Techniques and Results, 1. Trans Tech Publications, Bay Village.Google Scholar
Waddams, P. and Clark, N. M. (1991). The Petronella Field, Block 14/20b, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 353–60.Google Scholar
Wade, D. N., Lawrence, D. A., and Riley, L. A. (1995). The Rowan Sandstone Member (Upper Jurassic to Lower Cretaceous): stratigraphic definition and implications for North Sea exploration. Journal of Petroleum Geology, 18, 223–30.CrossRefGoogle Scholar
Wade, J. A. and MacLean, B. C. (1990). The geology of the southeastern margin of Canada. In Geology of the Continental Margin of Eastern Canada: Geology of North America, ed. Keen, M. J. and Williams, G. L.. Geological Survey of Canada, 2, 190238.Google Scholar
Wade, S. C. and Reiter, M. (1994). Hydrothermal estimation of vertical ground-water flow, Canutillo, Texas. Groundwater, 325, 735–42.Google Scholar
Wakefield, L. L., Droste, H., Giles, M. R., and Janssen, R. (1993). Late Jurassic plays along the western margin of the Central Graben. In Petroleum Geology of Northwest Europe, ed. Parker, J. R.. Proceedings of the 4th Conference, Geological Society of London, pp. 1167–78.Google Scholar
Walker, I. M., Berry, K. A., Bruce, J. R., Bystol, L., and Snow, J. H. (1997). Structural modelling of regional depth profiles in the Vøring Basin: implications for the structural and stratigraphic development of the Norwegian passive margin. Journal of the Geological Society of London, 154, 537–44.CrossRefGoogle Scholar
Walsh, J. B. (1981). Effect of pore pressure and confining pressure on fracture permeability. International Journal of Rock Mechanics and Mining Sciences, 18, 429–35.Google Scholar
Walsh, J. J. and Watterson, J. (1988). Analysis of the relationship between displacements and dimensions of faults. Journal of Structural Geology, 10, 239–47.CrossRefGoogle Scholar
Walter, M. J. (1999). Melting residues of fertile peridotite and the origin of cratonic lithosphere. In Mantle Petrology: Field Observations and High Pressure Experimentation: A Tribute to Francis R. (Joe) Boyd, ed. Fei, Y., Bertka, C. M. and Mysen, B. O.. Geochemical Society Special Publication, 6, 225–39.Google Scholar
Walters, M. and Combs, J. (1989). Heat flow regime in the Geysers-Clear Lake region of northern California. Geothermal Research Council Transactions, 13, 491502.Google Scholar
Walters, R. F. (1946). Burier pre-Cambrian hills ion northeastern Barton County, central Kansas. American Association of Petroleum Geologists Bulletin, 30, 660710.Google Scholar
Walters, R. F. (1953). Oil production from fractured pre-Cambrian basement rocks in central Kansas. American Association of Petroleum Geologists Bulletin, 37, 300–13.Google Scholar
Walters, R. F. and Price, A. S. (1948). Kraft-Prusa oil field, Barton County, Kansas. Oil and Gas Journal, 45, 249–80.Google Scholar
Walton, N. R. G. (1981). A detailed hydrogeochemical study of groundwaters from the Triassic sandstone aquifer of south-west England. Natural Environment Research Council, Institute of Geological Sciences, p. 43.Google Scholar
Wang, B. and Qian, K. (1992). Geology Research and Exploration Practice in the Shengli Petroleum Province (in Chinese). Petroleum University Press, Dongying, p. 239.Google Scholar
Wang, C.-Y. (1984). On the constitution of the San Andreas fault zone in Southern California. Journal of Geophysical Research, 89, 5858.CrossRefGoogle Scholar
Wang, C.-Y., Rui, F., Zengsheng, Y., and Xingjue, S. (1986). Gravity anomaly and density structure of the San Andreas fault zone. Pure and Applied Geophysics, 124, 127–40.CrossRefGoogle Scholar
Wannamaker, P. E., Doerner, W. M., Stodt, J. A., and Johnston, J. M. (1997). Subdued state of tectonism of the Great Basin interior relative to its eastern margin based on deep resistivity structure. Earth and Planetary Science Letters, 150, 4153.CrossRefGoogle Scholar
Wannamaker, P. E. and Hohman, G. W. (1991). Electromagnetic induction studies. U.S. National Report to the IUGG, invited paper, Review of Geophysics, pp. 405–15.Google Scholar
Wannesson, J., Icart, J.-C., and Ravat, J. (1991). Structure and evolution of two adjoining segments of the West African margin from deep seismic profiling. In Continental Lithosphere: Deep Seismic Reflections, ed. Meissner, R., Brown, L., Duerbaum, H.-J., Franke, W., Fuchs, K., and Seifert, F.. American Geophysical Union Geodynamics Series, 22, 275–89.Google Scholar
Waples, D. W. (1980). Time and temperature in petroleum formation: application of Lopatin’s method to petroleum exploration. American Association of Petroleum Geologists Bulletin, 64, 916–26.Google Scholar
Waples, D. W. (1985). Geochemistry in Petroleum Exploration. Reidel Publishing Company, Boston, p. 232.CrossRefGoogle Scholar
Waples, D. W. (1992). Future developments in basin modeling. 29th International Geological Congress, Abstracts – Congres Geologique Internationale, Resumes, 29, 812.Google Scholar
Waples, D. W. (1994a). Maturity modeling: thermal indicators, hydrocarbon generation, and oil cracking. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 285306.Google Scholar
Waples, D. W. (1994b). Modeling of sedimentary basins and petroleum systems. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 307–22.Google Scholar
Waples, D. W., Kamata, H., and Suizu, M. (1992a). The art of maturity modeling, Part 1: Finding a satisfactory geologic model. American Association of Petroleum Geologists Bulletin, 76, 3146.Google Scholar
Waples, D. W., Kamata, H., and Suizu, M. (1992b). The art of maturity modeling, Part 2: Alternative models and sensitivity analysis. American Association of Petroleum Geologists Bulletin, 76, 4766.Google Scholar
Waples, D. W. and Machihara, T. (1991). Biomarkers for geologists. American Association of Petroleum Geologists Methods in Exploration Series, 9.Google Scholar
Warnock, A. C., Zeitler, P. K., Wolf, R. A., and Bergman, S. C. (1997). An evaluation of low-temperature apatite U-Th/He thermochronometry. Geochimica et Cosmochimica Acta, 61, 5371–7.CrossRefGoogle Scholar
Warrington, G., Cope, J. C. W., and Ivimaycook, H. C. (1994). St. Audries Bay, Somerset, England: a candidate global stratotype section and point for the base of the Jurassic System. Geological Magazine, pp. 191200.CrossRefGoogle Scholar
Waschbusch, P. and Beaumont, C. (1996). Effect of slab retreat on crustal deformation in simple regions of plate convergence. Journal of Geophysical Research, 101, 28133–48.CrossRefGoogle Scholar
Waterston, C. D. (1950). Note on the sandstone injections of Eathie Haven, Cromarty. Geological Magazine, 87, 133–9.CrossRefGoogle Scholar
Watkeys, M. K. (2002). Development of the Lebombo rifted volcanic margin of Southeast Africa. In Volcanic Rift Margins, ed. M. A. Menzies, S. L. Klemperer, C. Ebinger, , and Baker, J.. Geological Society of America Special Paper, 362, pp. 2746.Google Scholar
Watkins, E. (2008). Tupi area presalt oil discoveries likely contiguous. Oil and Gas Journal, 106, 32–4.Google Scholar
Watkins, J. S., Ladd, J. W., Buffler, R. T., Shaub, F. J., Houston, M. H., and Worzel, J. L. (1978). Occurrence and evolution of salt in deep Gulf of Mexico. In Framework, Facies, and Oil-Trapping Characteristics of the Upper Continental Margin, ed. Bouma, A. H., Moore, G. T., and Coleman, J. M.. American Association of Petroleum Geologists Studies in Geology, 7, 4365.Google Scholar
Watson, E. B., Wark, D. A., and Thomas, J. B. (2006). Crystallization thermometers for zircon and rutile. Contributions to Mineralogy and Petrology, 151, 413–33.CrossRefGoogle Scholar
Watts, A. B. (1978). An analysis of isostasy in the world’s oceans: 1. Hawaiian-Emperor seamount chain. Journal of Geophysical Research, 83, 59896004.CrossRefGoogle Scholar
Watts, A. B. (2001). Isostasy and Flexure of the Lithosphere. Cambridge University Press, Cambridge, p. 458.Google Scholar
Watts, A. B., Karner, G. D., and Steckler, M. (1982). Lithospheric flexure and the evolution of sedimentary basins. Philosophical Transactions of the Royal Society of London, 305, 249–81.Google Scholar
Watts, A. B., Peirce, C., Collier, J., Dalwood, R., Canales, J. P., and Henstock, T. J. (1997). A seismic study of lithospheric flexure in the vicinity of Tenerife, Canary Islands. Earth and Planetary Science Letters, 146, 431–47.CrossRefGoogle Scholar
Watts, A. B. and Stewart, J. (1998). Gravity anomalies and segmentation of the continental margin offshore West Africa. Earth and Planetary Science Letters, 156, 239–52.CrossRefGoogle Scholar
Watts, A. B. and Thorne, J. (1984). Tectonics, global changes in sea level and their relationship to stratigraphic sequences at the U.S. Atlantic continental margin. Marine and Petroleum Geology, 1, 319–39.CrossRefGoogle Scholar
Watts, A. B. and Zhong, S. (2000). Observations of flexure and the rheology of oceanic lithosphere. Geophysical Journal International, 142, 855–75.CrossRefGoogle Scholar
Watts, N. L. (1987). Theoretical aspects of cap-rock and fault seals for single- and two-phase hydrocarbon columns. Marine and Petroleum Geology, 4, 274307.CrossRefGoogle Scholar
Wdowinski, S. and Axen, G. J. (1992). Isostatic rebound due to tectonic denudation; a viscous flow model of a layered lithosphere. Tectonics, 11, 303–15.CrossRefGoogle Scholar
Weast, R. C. (1974). CRC Handbook of Chemistry and Physics. CRC Press, Cleveland.Google Scholar
Weber, K. J. and Daukoru, E. (1975). Petroleum geology of the Niger delta. 9th World Petroleum Congress Proceedings, Tokyo, 2, 209–21.Google Scholar
Weeks, L. G. (1952). Factors of sedimentary basin development that control oil occurrence. American Association of Petroleum Geologists Bulletin, 36, 2071–124.Google Scholar
Weimer, P., Varnai, P., Budhijanto, F. M., Acosta, Z. M., Martinez, R. E., Navarro, A. F., Rowan, M. G., McBride, B. C., Villamil, T., Arango, C., Crews, J. R., and Pulham, A. J. (1998). Sequence stratigraphy of Pliocene and Pleistocene turbidite systems, northern Green Canyon and Ewing Bank (offshore Louisiana), northern Gulf of Mexico. American Association of Petroleum Geologists Bulletin, 82, 918–60.Google Scholar
Weinberg, R. F. and Podladchikov, Y. Y. (1994). Diapiric ascent of magmas through power law crust and mantle. Journal of Geophysical Research, 99, 9543–59.CrossRefGoogle Scholar
Weinberg, R. F. and Podladchikov, Y. Y. (1995). The rise of solid-state diapirs. Journal of Structural Geology, 17, 1183–95.CrossRefGoogle Scholar
Weissel, J. K. and Karner, G. D. (1989). Flexural uplift of rift flanks due to mechanical unloading of the lithosphere during extension. Journal of Geophysical Research, 94, 13919–50.CrossRefGoogle Scholar
Welsink, H. J., Dwyer, J. D., and Knight, R. J. (1989). Tectono-stratigraphy of the passive margin off Nova Scotia. American Association of Petroleum Geologists Memoir, 46, 215–31.Google Scholar
Welte, D. H., Hantschel, T., and Wygrala, B. (1997). Petroleum systems and the role of multi-dimensional basin modeling. Proceedings of an International Conference on Petroleum Systems of SE Asia and Australasia, pp. 23–4.CrossRefGoogle Scholar
Wenger, L. M., Davis, C. L., and Isaksen, G. H. (2002). Multiple controls on petroleum biodegradation and impact on oil quality. Society of Petroleum Engineers Reservoir Formation Evaluation and Engineering, 5, 375–83.Google Scholar
Wernicke, B. (1985). Uniform-sense simple shear of the continental lithosphere. Canadian Journal of Earth Sciences, 22, 108–25.CrossRefGoogle Scholar
Wernicke, B. P., Axen, G. J., and Snow, J. K. (1988). Basin and Range extensional tectonics at the latitude of Las Vegas, Nevada. Geological Society of America Bulletin, 100, 1738–57.2.3.CO;2>CrossRefGoogle Scholar
Wernicke, B. and Burchfiel, B. C. (1982). Modes of extensional tectonics. Journal of Structural Geology, 4, 105–15.CrossRefGoogle Scholar
Wesson, R. L. (1988). Dynamics of fault creep. Journal of Geophysical Research, 93, 8929–51.CrossRefGoogle Scholar
Wheatley, T. J., Biggins, D., Buckingham, J., and Holloway, N. H. (1987). The geology and exploration of the transitional shelf, an area to the west of the Viking Graben. In Petroleum Geology of North West Europe, ed. Brooks, J. and Glennie, K. W.. Graham and Trotman, London, pp. 979–89.Google Scholar
Whelan, J., Kennicut, M., Brooks, J., Schumacher, D., and Eglington, L. (1994). Organic geochemical indicators of dynamic fluid flow processes in petroleum basins. Organic Geochemistry, 22, 587615.CrossRefGoogle Scholar
White, D. A. (1993). Geologic risking guide for prospects and plays. American Association of Petroleum Geologists Bulletin, 77, 2048–64.Google Scholar
White, N. and Lovell, B. (1997). Measuring the pulse of a plume with the sedimentary record. Nature, 387, 888–91.CrossRefGoogle Scholar
White, N. and McKenzie, D. (1988). Formation of the “steer’s head” geometry of sedimentary basins by differential stretching of the crust and mantle. Geology, 16, 250–3.2.3.CO;2>CrossRefGoogle Scholar
White, R. S. (1992). Crustal structure and magmatism of North Atlantic continental margins. Journal of the Geological Society of London, 149, 841–54.CrossRefGoogle Scholar
White, R. S. (1997). Rift-plume interaction in the North Atlantic. Series A. Philosophical Transactions of the Royal Society of London, 355, 319–39.Google Scholar
White, R. S. and McKenzie, D. (1989). Magmatism at rift zones: the generation of volcanic continental margins and flood basalts. Journal of Geophysical Research, 94, 7685–729.CrossRefGoogle Scholar
White, R. S., McKenzie, D., and O’Nions, R. K. (1992). Oceanic crustal thickness from seismic measurements and rare earth element inversions. Journal of Geophysical Research, 97, 19683–715.CrossRefGoogle Scholar
White, R. S., Spence, G. D., Fowler, S. R., McKenzie, D. P., Westbrook, G. K., and Bowen, A. N. (1987). Magmatism at rifted continental margins: Nature, 330, 439–44.CrossRefGoogle Scholar
Whitehead, M. and Pinnock, S. J. (1991). The Highlander Field, Block 14/20b, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 323–9.Google Scholar
Whiticar, M. J. (1994). Correlation of natural gases with their sources. In The Petroleum System – From Source to Trap, ed. Magoon, L. B. and Dow, W. G.. American Association of Petroleum Geologists Memoir, 60, 261–83.Google Scholar
Whiticar, M. J. and Faber, E. (1986). Methane oxidation in sediment and water column environments – isotopic evidence. In Advances in Organic Geochemistry, ed. Leythaeuser, D. and Rulkotter, J.. Pergamon Press, New York, 10, 759–68.Google Scholar
Whitmarsh, R. B., Dean, S. M., Minshull, T. A., and Tompkins, M. (2000). Tectonic implications of exposure of lower continental crust beneath the Iberia abyssal plain, Northeast Atlantic Ocean: geophysical evidence. Tectonics, 19, 919–42.CrossRefGoogle Scholar
Whitmarsh, R. B., Manatschal, G., and Minshull, T. A. (2001). Evolution of magma-poor continental margins from rifting to seafloor spreading. Nature, 413, 150–4.CrossRefGoogle ScholarPubMed
Whitmarsh, R. B. and Miles, P. R. (1995). Models of the development of the West Iberia rifted continental margin at 40°30’N deduced from surface and deep-tow magnetic anomalies. Journal of Geophysical Research, 100, 3789–806.CrossRefGoogle Scholar
Whitmarsh, R. B., Miles, P. R., and Mauffret, A. (1990). The ocean-continent boundary off western continental margin of Iberia – I. Crustal structure at 40°30’N. Geophysical Journal International, 103, 509–31.CrossRefGoogle Scholar
Whitmarsh, R. B., White, R. S., Horsefield, S. J., Sibuet, J.-C., Recq, M., and Louvel, V. (1996). The ocean-continent boundary off the western continental margin of Iberia: crustal structure west of Galicia Bank. Journal of Geophysical Research, 101, 28291–314.CrossRefGoogle Scholar
Wignall, P. B. and Maynard, J. R. (1993). The sequence stratigraphy of transgressive black shales. In Source Rocks in a Sequence Stratigraphic Framework, ed. Katz, B. J. and Pratt, L. M.. American Association of Petroleum Geologists Studies in Geology, 37, 3547.Google Scholar
Wilhelm, B. and Somerton, W. H. (1967). Simultaneous measurement of pore and elastic properties of rocks under triaxial stress condition. Journal of the Society of Petroleum Engineers, 7, 283–94.Google Scholar
Willemse, E. J. M., Peacock, D. C. P., and Aydin, A. (1997). Nucleation and growth of strike-slip faults in limestones from Somerset, UK. Journal of Structural Geology, 19, 1461–77.CrossRefGoogle Scholar
Williams, H. H. and Eubank, R. T. (1995). Hydrocarbon habitat in the rift graben of the central Sumatra Basin, Indonesia. In Hydrocarbon Habitat in Rift Basins, ed. Lambiase, J. J.. Geological Society of London Special Publications, 80, 331–71.Google Scholar
Williams, J. J., Conner, D. C., and Peterson, K. E. (1975). Piper oilfield, North Sea: fault-block structure with Upper Jurassic beach/bar reservoir sands. American Association of Petroleum Geologists Bulletin, 59, 1581–601.Google Scholar
Williams, R. R. (1991). The Deveron field, Block 211/18a, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. Abbotts, I. L.. Memoir of the Geological Society of London, 14, 83–7.Google Scholar
Wilson, D. L. (2002). Planning for success: Profit vs. reserves. Oil and Gas Journal, 100, 30–1.Google Scholar
Wilson, D. S., Teagle, D. A. H., Alt, J., Banerjee, N., Umino, S., Miyashita, S., and Acton, G. D. (2005). Drilling a complete in situ section of upper oceanic crust formed at superfast spreading rate: Hole 1256D. EGU General Assembly, Vienna, p. 374.Google Scholar
Wilson, M. (1997). Thermal evolution of the Central Atlantic passive margins: continental break-up above a Mesozoic super-plume. Journal of the Geological Society, 154, 491–5.CrossRefGoogle Scholar
Wilson, P. G., Turner, J. P., and Westbrook, G. K. (2003). Structural architecture of the ocean-continent boundary at an oblique transform margin through deep-imaging seismic interpretation and gravity modelling: Equatorial Guinea, West Africa. Tectonophysics, 374, 1940.CrossRefGoogle Scholar
Wilson, R. C. L., Manatschal, G., and Wise, S. (2001). Rifting along non-volcanic passive margins: stratigraphic and seismic evidence from the Mesozoic successions of the Alps and western Iberia. Geological Society Special Publications, 187, 429–52.CrossRefGoogle Scholar
Wilson, W. R. (1980). Thermal studies in a geothermal area. PhD thesis, University of Utah, Salt Lake City, p. 145.Google Scholar
Winterer, E. L. (1989). Sediment thickness map of the Northeast Pacific. In The Eastern Pacific Ocean and Hawaii. The Geology of North America, ed. Winterer, E. L. H., Decker, D. M., and Robert, W.. Geological Society of America, Denver, Colorado.Google Scholar
Wise, D. U. (1992). Dip domain method applied to the Mesozoic Connecticut Valley rift basins. Tectonics, 11, 1357–68.CrossRefGoogle Scholar
Witcher, J. C. (1988). Geothermal resources of southwestern New Mexico and southeastern Arizona. Field Conference Guidebook, New Mexico Geological Society, 39, 191–8.Google Scholar
Withjack, M. O., Meisling, K. E., and Russell, L. R. (1988). Forced folding and basement-detached normal faulting in the Haltenbanken area, offshore Norway. American Association of Petroleum Geologists Bulletin, 72, 259.Google Scholar
Withjack, M. O., Olsen, P. E., and Schlische, R. W. (1995). Tectonic evolution of the Fundy rift basin, Canada: evidence of extension and shortening during passive margin Development. Tectonics, 14, 390405.CrossRefGoogle Scholar
Withjack, M. O., Olson, J., and Peterson, E. (1990). Experimental models of extensional forced folds. American Association of Petroleum Geologists Bulletin, 74, 1038–54.Google Scholar
Withjack, M. O., Schlische, R. W., and Baum, M. S. (2009). Extensional development of the Fundy rift basin, southeastern Canada, Geological Journal, 44 (6), 631–51.CrossRefGoogle Scholar
Withjack, M. O., Schlische, R. W., and Olsen, P. E. (1998): Diachronous rifting, drifting, and inversion on the passive margin of Central Eastern North America: an analog for other passive margins. American Association of Petroleum Geologists Bulletin, 82, 817–35.Google Scholar
Witkind, I. J., Myers, W. B., Hadley, J. B., Hamilton, W., and Fraser, G. D. (1962). Geologic features of the earthquake at Hebgen Lake, Montana, August 17, 1959. Bulletin of the Seismological Society of America, 52, 163–80.CrossRefGoogle Scholar
Witlox, H. W. M. (1986). Finite element simulation of basal extension faulting within a sediment overburden. Numerical Methods in Geomechanics – European Conference, Stuttgard, 2, 765.Google Scholar
Wojtal, S. and Mitra, G. (1986). Strain hardening and strain softening in fault zones from foreland thrusts. Geological Society of America Bulletin, 97, 674–87.2.0.CO;2>CrossRefGoogle Scholar
Wolf, R. A., Farley, K. A., and Silver, L. T. (1996). Helium diffusion and low-temperature thermochronometry of apatite. Geochimica et Cosmochimica Acta, 60, 4231–40.CrossRefGoogle Scholar
Wolfe, C. J., McNutt, M. K., and Detrick, R. S. (1994). The Marquesas archipelagic apron: seismic stratigraphy and implications for volcano growth, mass wasting, and crustal underplating. Journal of Geophysical Research, 99, 13591–608.CrossRefGoogle Scholar
Wolff, G. A., Lamb, N. A., and Maxwell, J. R. (1986a). The origin and fate of 4-methyl steroid hydrocarbons. I. Diagenesis of 4-methyl steranes. Geochimica et Cosmochimica Acta, 50, 335–42.CrossRefGoogle Scholar
Wolff, G. A., Lamb, N. A., and Maxwell, J. R. (1986b). The origin and fate of 4-methyl steroid hydrocarbons. II. Dehydration of steroids and occurrence of C30 4-methylsteranes. Geochimica et Cosmochimica Acta, 50, 965–74.CrossRefGoogle Scholar
Wollard, G. P. (1959). Crustal structure from gravity and seismic measurements. Journal of Geophysical Research, 64, 1521–44.CrossRefGoogle Scholar
Wood, D. (2003). More aspects of E&P asset and portfolio risk analysis. Oil and Gas Journal, 101, 2832.Google Scholar
Wood, J. A. and Hewett, T. A. (1982). Fluid convection and mass transfer in porous sandstones – a theoretical approach. Geochimica et Cosmochimica Acta, 46, 1701–13.CrossRefGoogle Scholar
Wood, J. A. and Hewett, T. A. (1984). Reservoir diagenesis and convective fluid flow. In Clastic Diagenesis, ed. McDonald, D. A. and Surdam, R. C.. American Association of Petroleum Geologists Memoir, 37, 99110.Google Scholar
Woodbury, H. O., Murray, I. B., Pickford, P. J., and Akers, W. H. (1973). Pliocene and Pleistocene depocenters, outer continental shelf of Louisiana and Texas. American Association of Petroleum Geologists Bulletin, 57, 2428–39.Google Scholar
Woodside, W. and Messmer, J. H. (1961). Thermal conductivity of porous media I: Unconsolidated sand. Journal of Applied Geophysics, 32, 1688–706.Google Scholar
Workman, R. K. and Hart, S. R. (2005). Major and trace element composition of the depleted MORB mantle (DMM). Earth and Planetary Science Letters, 231, 5372.CrossRefGoogle Scholar
Wride, V. C. (1994). Structural features and structural styles from the five countries area of the North Sea Central Graben. First Break, 13, 395407.Google Scholar
Wright, L. A. (1974). Fault Map of the Region of Central and Southern Death Valley, Eastern California and Western Nevada, Guidebook; Death Valley Region, California and Nevada. Death Valley Publication Company.Google Scholar
Wright, L. D. (1977). Sediment transport and deposition at river mouths: a synthesis. American Association of Petroleum Geologists Bulletin, 88, 857–68.Google Scholar
Wright, T. L. (1991). Structural geology and tectonic evolution of the Los Angeles basin, California. In Active Margin Basins, ed. Biddle, K. T.. American Association of Petroleum Geologists Memoir, 52, 35135.Google Scholar
Wygrala, B. P. (1989). Integrated Study of an Oil Field in the Southern Po Basin, Northern Italy. University of Koln.Google Scholar
Xie, X., Jiao, J. J., Tang, Z., and Zheng, C. (2003). Evolution of abnormally low pressure and its implications for the hydrocarbon system in the southeast uplift zone of Songliao Basin, China. American Association of Petroleum Geologists Bulletin, 87, 99119.Google Scholar
Xie, X. N., Li, S. T., Dong, W. L., and Hu, Z. L. (2001). Evidence for episodic expulsion of hot fluids along faults near diapiric structures of the Yinggehai Basin, South China Sea. Marine and Petroleum geology, 18, 715–28.CrossRefGoogle Scholar
Xiong, Y., Wang, Y., and Wang, Y. (2007). Selective chemical degradation of kerogen from Nenjiang Formation of the southern Songliao Basin. Science in China, Series D, Earth Series, 50, 1504–9.Google Scholar
Xue, L. and Galloway, W. E. (1993). Genetic sequence stratigraphic framework, depositional style, and hydrocarbon occurrence of the Upper Cretaceous QYN formations in the Songliao lacustrine basin, northeastern China. American Association of Petroleum Geologists Bulletin, 77, 1792–808.Google Scholar
Yalcin, M. N., Littke, R., and Sachsenhofer, R. F. (1997). Thermal history of sedimentary basins. In Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling, ed. Welte, D. H., Horsfield, B., and Baker, D. R.. Springer-Verlag, Berlin, pp. 71167.CrossRefGoogle Scholar
Yaliz, A. 1991. The Crawford Field, block 9/28a, UK North Sea. In United Kingdom Oil and Gas Fields, 25 Years Commemorative Volume, ed. I. L. Abbotts. Memoir of the Geological Society of London, 14, 287–93.Google Scholar
Yamasaki, T. and Nakada, M. (1997). The effects of the spinel-garnet phase transition on the formation of rifted sedimentary basins. Geophysical Journal International, 130, 681–92.CrossRefGoogle Scholar
Yang, J., Large, R. R., Bull, S., and Scott, D. L. (2006). Basin-scale numerical modeling to test the role of buoyancy-driven fluid flow and heat transfer in the formation of stratiform Zn-Pb-Ag deposits in the northern Mount Isa Basin. Economic Geology, 101, 1275–92.CrossRefGoogle Scholar
Yang, J., Latychev, K., and Edwards, R. N. (1998). Numerical computation of hydrothermal fluid circulation in fractured earth structures. Geophysical Journal International, 135, 627–49.CrossRefGoogle Scholar
Yardley, B. W. D. (1986). Is there water in the deep continental crust? Nature, 323, 111.CrossRefGoogle Scholar
Ye, S., Canales, J. P., Rihm, R., Danobeitia, J. J., and Gallart, J. (1999). A crustal transect through the northern and northeastern part of the volcanic edifice of Gran Canaria, Canary Islands. Journal of Geodynamics, 28, 326.CrossRefGoogle Scholar
Yiang, W., Li, Y., and Gao, R. (1985). Formation and evolution of nonmarine petroleum in Songliao Basin, China. American Association of Petroleum Geologists Bulletin, 69, 1112–22.Google Scholar
Yielding, G., Badley, M. E., and Roberts, A. M. (1992). The structural evolution of the Brent Province. In Geology of the Brent Group, ed. Morton, A. C., Haszeldine, R. S., Giles, M. R., and Brown, S.. Geological Society Special Publications, pp. 2743.Google Scholar
Yielding, G., Jackson, J. A., King, G. C. P., Sinval, H., Vita-Finzi, C., and Wood, R. M. (1981). Relations between surface deformation, fault geometry, seismicity and rupture characteristics during the El Asnam (Algeria) earthquake of 10 October 1980. Earth and Planetary Science Letters, 56, 287304.CrossRefGoogle Scholar
Younes, A. I. (1996). Fracture distribution on faulted basement blocks, Gulf of Suez, Egypt: reservoir characterization and tectonic implications. PhD. thesis, Pennsylvania State University at University Park, University Park, p. 170.Google Scholar
Younes, A. I. and McClay, K. R. (2002). Development of accommodation zones in the Gulf of Suez–Red Sea Rift, Egypt. American Association of Petroleum Geologists Bulletin, 86, 1003–26.Google Scholar
Yuen, D. A. and Fleitout, L. (1985). Thinning of the lithosphere by small-scale convective destabilization. Nature, 313, 125–8.CrossRefGoogle Scholar
Yukler, M. A. and Erdogan, T. L. (1996). Effects of geological parameters on temperature histories of basins. Annual Meeting Abstracts – American Association of Petroleum Geologists and Society of Economic Paleontologists and Mineralogists, 5, 157–8.Google Scholar
Zalán, P. V., Nelson, E. P., Warme, J. E., and Davis, T. (1985). The Piauí Basin, rifting and wrenching in an equatorial Atlantic transform basin. In Strike-slip Deformation, Basin Formation, and Sedimentation, ed. K. T. Biddle and N. Christie-Blick. Society of Economic Paleontologists and Mineralogists. Tulsa, OK, Special Publication 37, pp. 177–92.Google Scholar
Zalán, P. V., Severino, M. C. G., Rigoti, C. A., Magnavita, L. P., Oliveira, J. A. B., and Vianna, A. R. (2011). An entirely new 3-D view of the crustal and mantle structure of a South Atlantic passive margin – Santos, Campos and Espírito Santo Basins, Brazil. Search and Discovery Article No. 30177, Adapted from expanded abstract presentation at AAPG Annual Convention and Exhibition, Houston, Texas, April 10–13, 2011.Google Scholar
Zanella, E. and Coward, M. P. (2003). Chapter 4. Structural framework. In The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea, ed. Evans, D., Graham, C., Armour, A., and Bathurst, P.. Geological Society of London, pp. 4559.Google Scholar
Zappone, A., Burlini, L., Connolly, J., Husen, S., and Kissling, E. (2006). Lower crustal seismic velocities and seismicity in the northern Alpine foreland: I. Observations and petrophysical experiments. Geophysical Research Abstracts, 8.Google Scholar
Zeitler, P. K., Herczeg, A. L., McDougall, I., and Honda, M. (1987). U-Th-He dating of apatite: a potential thermochronometer. Geochimica et Cosmochimica Acta, 51, 2865–8.CrossRefGoogle Scholar
Zeng, J. H. and Jin, Z. J. (2003). Experimental investigation of episodic oil migration along fault systems. Journal of Geochemical Exploration, 78–79, 493–8.Google Scholar
Zhang, Y. J. (2002). Petroleum migration and formation in fault-controlled Junggar Basin (in Chinese). PhD thesis, University of Petroleum, Beijing, p. 232.Google Scholar
Zhao, G. and Johnson, A. M. (1991). Sequential and incremental formation of conjugate faults. Journal of Structural Geology, 13, 887896.CrossRefGoogle Scholar
Ziegler, P. A. (1989a). Geodynamic model for Alpine intra-plate compressional deformation in Western and Central Europe. In Inversion Tectonics Meeting, ed. Cooper, M. A. and Williams, G. D.. Geological Society of London Special Publications, 44, 6385.Google Scholar
Ziegler, P. A. (1989b). Evolution of the North Atlantic – an overview. In Extensional Tectonics and Stratigraphy of the North Atlantic Margins, ed. Tankard, A. J. and Balkwill, H. R.. American Association of Petroleum Geologists Memoir, 46, 111–29.Google Scholar
Ziegler, P. A. (1990). Geological Atlas of Western and Central Europe (second edition). Shell Internationale Petroleum Maatschappij and Geological Society, London, p. 239.Google Scholar
Ziegler, P. A. (1990). Tectonic and palaeogeographic development of the North Sea rift system. In Tectonic Evolution of the North Sea rifts, ed. Blundell, D. J. and Gibbs, A. D.. Clarendon Press, Oxford, p. 36.Google Scholar
Ziegler, P. A. (1994). Geodynamic processes governing development of rifted basins. In Geodynamic Evolution of Sedimentary Basins, the International Symposium Held in Moscow, May 18–23, 1992, ed. Roure, F., Ellouz, N., Shein, V. S., and Skvortsov, I. I.. Editions Technip, Paris, pp. 1937.Google Scholar
Ziegler, P. A. (1995). Cenozoic rift system of Western and Central Europe: an overview. Geologie en Mijnbouw, 73, 99127.Google Scholar
Zijerveld, L., Stephenson, R., Cloetingh, S., Duin, E., and van den Berg, M. W. (1992). Subsidence analysis and modelling of the Roer Valley Graben (SE Netherlands): Geodynamics of rifting: Volume I, Case history studies on rifts; Europe and Asia. Geodynamics of rifting symposium, Glion-sur-Montreux, Switzerland, 208, 159–71.Google Scholar
Zimmerman, R. W., Somerton, W. H., and King, M. S. (1986). Compressibility of porous rocks. Journal of Geophysical Research, 91, 12765–77.CrossRefGoogle Scholar
Zoback, M. L. (1983). Structure and Cenozoic tectonism along the Wasatch fault zone, Utah. In Tectonics and Stratigraphy of the Eastern Great Basin, ed. Miller, D. M., Todd, V. R., and Howard, K. A.. Geological Society of America Memoir, 157, 337.CrossRefGoogle Scholar
Zoetemeijer, R. (1993). Tectonic modeling of foreland basins: thin-skinned thrusting, syntectonic sedimentation and lithospheric flexure. PhD thesis, Vrije University, Amsterdam, p. 148.Google Scholar
Zoth, G. (1979). Temperaturmessungen in der Bohrung Hanigsen sowie Warmeleitfahigeitbestimmungen von Gesteinproben. Report BGR/NLfB Hannover, 81, 828.Google Scholar
Zoth, G., Haenel, R., Rybach, L and Stegena, L. (1988). Appendix. In Handbook of Terrestrial Heat-flow Density Determination; with Guidelines and Recommendations of the International Heat Flow Commission, ed. R. Haenel. Kluwer Academic Publishers, Dordrecht.Google Scholar
Zuber, M. T. and Parmentier, E. M. (1986). Lithospheric necking: a dynamic model for rift morphology. Earth and Planetary Science Letters, 77, 373–83.CrossRefGoogle Scholar
Zwach, C., Poelchau, H. S., Hantschel, T., and Welte, D. H. (1994). Simulation with contrasting pore fluids: Can we afford to neglect hydrocarbon saturation in basin modeling? In Basin modeling – What Have We learned?, ed. S. Dueppenbecker and J. Iliffe. Proceedings of Basin Modeling Conference, British Geological Survey, p. 4.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Michal Nemčok, University of Utah
  • Book: Rifts and Passive Margins
  • Online publication: 05 March 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139198844.022
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Michal Nemčok, University of Utah
  • Book: Rifts and Passive Margins
  • Online publication: 05 March 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139198844.022
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Michal Nemčok, University of Utah
  • Book: Rifts and Passive Margins
  • Online publication: 05 March 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781139198844.022
Available formats
×