Skip to main content Accessibility help
×
Hostname: page-component-f554764f5-246sw Total loading time: 0 Render date: 2025-04-21T15:02:54.292Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  23 January 2025

Michal Nemčok
Affiliation:
RM Geology
Anthony G. Doré
Affiliation:
Statoil (UK) Ltd.
Andreas Henk
Affiliation:
Technische Universität Darmstadt
Helen Doran
Affiliation:
Ola Geoscience
Get access
Type
Chapter
Information
Strike-Slip Terrains and Transform Margins
Structural Architecture, Thermal Regimes and Petroleum Systems
, pp. 661 - 734
Publisher: Cambridge University Press
Print publication year: 2025

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.)

Book purchase

Temporarily unavailable

References

Aajdour, M., Hssaïda, T., Fedan, B., et al., 1999. Contribution à l’étude du Jurassique de la marge marocaine. In: Apports de l’analyse palynologique dans l’évolution géodynamique: 1st Colloquium on Moroccan Jurassic, Morocco, pp. 12.Google Scholar
Aamir, M., and Siddiqui, M. M., 2006. Interpretation and visualization of thrust sheets in a triangle zone in eastern Potwar, Pakistan. The Leading Edge, January, 24–37.CrossRefGoogle Scholar
Abercombrie, R. E., and Ekstrom, G. A., 2001. Earthquake slip on oceanic transform faults. Nature, 410: 7477.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.Google Scholar
Accocella, V., Salvini, F., Funiciello, R., and Facenna, C., 1999. The role of transfer structures on volcanic activity at Campi Flegrei (Southern Italy). Journal of Volcanology and Geothermal Research, 91: 123139.CrossRefGoogle Scholar
Accocella, V., Facenna, C., Funiciello, R., and Rossetti, F., 2000. Analogue modeling of extensional transfer zones. Bolletino della Societa Geologica Italiana, 119: 8596.Google Scholar
Adams, E. W., and Schlager, W., 2000 Basic types of submarine slope curvature. Journal of Sedimentary Research, 70(4): 814828.CrossRefGoogle Scholar
Adams, J. T., and Dart, C., 1998. The appearance of potential sealing faults on borehole images. In: Jones, G., Fisher, Q. J., and Knipe, R. J. (Eds.), Faulting, Fault Sealing and Fluid Flow in Hydrocarbon Reservoirs. Geological Society, London, pp. 7186.Google Scholar
Addis, D., Brown, A., Grant, J., et al., 2013a. The tectonic evolution of the Deep Ivorian Basin and its influence on Upper Cretaceous deepwater depositional systems. Sheared Margins Workshop, Snowbird, Utah, USA, September 30–October 1, abstracts.Google Scholar
Addis, D., Brown, A., Grant, J., 2013b. Influences of tectonics, drainage and sediment supply on Upper Cretaceous deepwater deposits in the Deep Ivorian Basin of Western Ghana and Cote d’Ivoire. The 12th PESGB/HGS Conference on African E&P, Wembley Stadium, London, September 10–11. Extended abstracts.Google Scholar
Advocate, D. M., Young, S. W., Ross, A.H., et al., 1998. Post-rift hydrocarbon systems, Greater Amazon Mouth, Brazil: transition from shelf to basin and source distribution controls. In: Mello, M.R., and Yilmaz, P.O. (Eds.), Petroleum Geology in a Changing World. AAPG International Conference, Extended Abstracts Volume, Rio de Janiero, pp. 602603.Google Scholar
African Petroleum, 2015. Corporate presentation. January.Google Scholar
African Petroleum, 2017. Corporate presentation. January.Google Scholar
Aftalion, M., Bowes, D. R., Dash, B., and Fallick, A.E., 1998. Pan-African thermal history of the Mid-Proterozoic Khariar alkali syenite in the Eastern Ghats, Orissa, India: A U–Pb and K–Ar isotopic study. International Seminar on Precambrian Crust in Eastern and Central India, Geological Survey of India, Bhubaneswar. Abstract, pp. 10–12.Google Scholar
Agirrezabala, L. M., 2009. Mid-Cretaceous hydrothermal vents and authigenic carbonates on a transform margin, Basque-Cantabrian Basin (western Pyrenees): a multidisciplinary study. Sedimentology, 59: 969996.CrossRefGoogle Scholar
Agustsson, K., Linde, A., Stefansson, R., and Sacks, S., 1999. Strain changes for the 1987 Vatnafjöll earthquake South Iceland and possible magmatic triggering, Journal of Geophysical Research, 104: 11511161.CrossRefGoogle Scholar
Ahnert, F., 1970. Functional relationships between denudation, relief, and uplift in large, mid-latitude drainage basins. American Journal of Science, 268: 243263.CrossRefGoogle Scholar
Aiello, I. W., 2005. Fossil seep structures of the Monterey Bay region and tectonic/structural controls on fluid flow in an active transform margin. Palaeogeography, Palaeoclimatology, Palaeoecology, 227: 124142.CrossRefGoogle Scholar
Aiello, I. W., Garrison, R. E., Moore, J. C., Kastner, M., and Stakes, D. D., 2001. Anatomy and origin of carbonate structures in a Miocene cold seep field. Geology, 29: 11111114.2.0.CO;2>CrossRefGoogle Scholar
Al Haj Enterprises Limited, 2016. Baska North Block (3169-4) farm-out presentation. Islamabad, May 19.Google 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. AAPG Bulletin, 82: 387411.Google Scholar
Allan, U. S., 1989. Model for hydrocarbon migration and entrapment within faulted structures. AAPG Bulletin, 73(7): 803811.Google Scholar
Allemand, P., and Brun, J.-P., 1991. Width of continental rifts and rheological layering of the lithosphere. Tectonophysics, 188(1–2): 6369.CrossRefGoogle Scholar
Allen, P., 2005. Striking a chord. Nature 434: 961.CrossRefGoogle ScholarPubMed
Allen, P. A., and Allen, J. R., 1990. Basin Analysis: Principles and Applications. Blackwell Scientific, Oxford.Google Scholar
Allen, P. A., and Densmore, A. L., 2000. Sediment flux from an uplifting fault block. Basin Research, 12: 367380.CrossRefGoogle Scholar
Allison, L. E., 1965. Organic carbon. In: Black, C.A., et al., (Eds). Methods of Soil Analysis: Part 2. Wiley, Chichester, pp. 13671378.Google Scholar
Allken, V., Huismans, R. S., and Thieulot, C., 2012. Factors controlling the mode of rift interaction in brittle–ductile coupled systems: a 3D numerical study. Geochemistry, Geophysics, Geosystems, 13: Q05010.CrossRefGoogle Scholar
Almalki, K. A., Betts, P. G., and Ailleres, L., 2014. Episodic sea-floor spreading in the Southern Red Sea. Tectonophysics, 617: 140149.CrossRefGoogle Scholar
Al-Tabaa, A., and Wood, D. M., 1987. Some measurements of permeability of kaolin. Geotechnique, 37: 499503.CrossRefGoogle Scholar
Alvarez, F., Virieux, J., and Le Pichon, X., 1984. Thermal consequences of lithosphere extension over continental margins: the initial stretching phase. Geophysical Journal of the Royal Astronomical Society, 78(2): 389411.CrossRefGoogle Scholar
Amjad, A., 2010. Structural analysis of the Trans-Indus Ranges: Implications for the hydrocarbon potential of the NW Himalayas, Pakistan. PhD thesis. National Center of Excellence in Geology University of Peshawar, Pakistan.Google Scholar
Anderson, E. M., 1951. The Dynamics of Faulting and Dyke Formation With Application to Britain. Edinburgh: Oliver & Boyd.Google Scholar
Anderson, R., and Hobart, M., 1976. The relation between heat flow, sediment thickness, and age in the eastern Pacific. Journal of Geophysical Research, 81: 29682989.CrossRefGoogle 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., et al., 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, 1CD-ROM.Google Scholar
Andreani, M., Mevel, C., Boullier, A. M., and Escartín, J., 2007. Dynamic control on serpentinite crystallization in veins: constraints on hydration processes in oceanic peridotites. Geochemistry, Geophysics, Geosystems, 8: Q02012.CrossRefGoogle Scholar
Angelier, J., 1985. Extension and rifting; the Zeit region, Gulf of Suez. Journal of Structural Geology, 7(5): 605612.CrossRefGoogle Scholar
Angelier, J., 1989. From orientation to magnitudes in paleostress determinations using fault slip data. Journal of Structural Geology, 11: 3750.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: 151160.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(11): 21752200.CrossRefGoogle Scholar
ANP, 1994. Carta estratigrafica da Sub-Bacia de Mundau.Google Scholar
Antobreh, A. A., Faleide, J. I., Tsikalas, F., and Planke, S., 2009. Rift-shear architecture and tectonic development of the Ghana margin deduced from multichannel seismic reflection and potential field data. Marine and Petroleum Geology, 26: 345368.CrossRefGoogle Scholar
Antonellini, M., and Aydin, A., 1994. Effect of faulting on fluid flow in porous sandstones: petrophysical properties. AAPG Bulletin, 78: 355377.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. AAPG Bulletin, 84: 314344.Google Scholar
Antonellini, M., Aydin, A., and Pollard, D. D., 1994. Microstructure of deformation bands in porous sandstones at Arches National Park, Utah. Journal of Structural Geology, 16: 941959.CrossRefGoogle Scholar
AOA Geophysics, 2001. Boujdour Permit Area, Morocco. Phase 1 Investigation. AOA Geophysics, Houston.Google Scholar
Appelgate, B., and Shor, A. N., 1994. The northern Mid-Atlantic and Reykjanes Ridges: spreading centre morphology between 55°50′N and 63°00′N. Journal of Geophysical Research, 99: 1793517956.CrossRefGoogle Scholar
Arabasz, W. J., 1971. Geological and geophysical studies of the Atacama Fault System in northern Chile. PhD thesis. California Institute of Technology, Pasadena.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: 90359047.CrossRefGoogle Scholar
Archie, G. E., 1942. The electrical resistivity log as an aid in determining some reservoir characteristics. Petroleum Transactions of AIME, 146: 5462.CrossRefGoogle Scholar
Arens, G., Delteil, J. R., Valery, P., et al., 1971. The continental margin of the Ivory Coast and Ghana. In: Delany, F. M. (Ed.), The Geology of the East Atlantic Continental Margin; 4, Africa. Report. British Geological Survey, London, pp. 6178.Google Scholar
Armijo, R., Flerit, F., King, G., and Meyer, B., 2004. Linear elastic fracture mechanics explains the past and present evolution of the Aegean. Earth and Planetary Science Letters, 217: 8595.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(B3). doi: 10.1029/2001JB001708.CrossRefGoogle 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: 400411.CrossRefGoogle Scholar
Árnadóttir, T., Jiang, W., Feigl, K. L., Geirsson, H., and Sturkell, E., 2006. Kinematic models of plate boundary deformation in southwest Iceland derived from GPS observations. Journal of Geophysical Research, 111: B07402.CrossRefGoogle Scholar
Árnadóttir, T., Geirsson, H., and Jiang, W., 2008. Crustal deformation in Iceland: plate spreading and earthquake deformation. Jökull, 58: 5974.CrossRefGoogle Scholar
Arnaud, N., Tapponnier, P., Roger, F., et al., 2003. Evidence for Mesozoic shear along the western Kunlun and Altyn-Tagh faults, northern Tibet (China). Journal of Geophysical Research, Solid Earth 108. DOI: 10.1029/2001JB000904.Google Scholar
Arrowsmith, J. R., 1995. Coupled Tectonic Deformation and Geomorphic Degradation Along the San Andreas Fault System, Stanford University, Stanford.Google Scholar
Artemieva, I. M., 2009. The continental lithosphere: reconciling thermal, seismic, and petrologic data. Lithos, 109: 2346.CrossRefGoogle Scholar
Artemieva, I. M., 2011. Lithosphere: An Interdisciplinary Approach. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Artemieva, I. M., and Mooney, W. D., 2001. Thermal thickness and evolution of Precambrian lithosphere: a global study. Journal of Geophysical Research, 106(B8): 1638716414.CrossRefGoogle Scholar
Artemieva, I. M., and Mooney, W. D., 2002. On the relations between cratonic lithosphere thickness, plate motions, and basal drag. Tectonophysics, 358(1–4): 211231.CrossRefGoogle Scholar
Arthur, M. A., Dean, W. E., Schlanger, S. O., 1985. Variations in the global carbon cycle during the Cretaceous related to climate, volcanism, and changes in atmospheric CO2. In Sundquist, E.T., and Broecker, W.S. (Eds.), The Carbon Cycle and Atmospheric CO2: Natural Variations from Archean to Present. American Geophysical Union, Washington, DC, pp. 504529.Google Scholar
Arthur, M. A., Dean, W. E., and Pratt, L. M., 1988. Geochemical and climatic effects of increased marine organic carbon burial at the Cenomanian/Turonian boundary. Nature, 335: 714717.CrossRefGoogle Scholar
Arthur, M. A., Jenkyns, H. C., Brumsack, H. J., and Schlanger, S. O., 1990. Stratigraphy, geochemistry, and paleooceanography of organic carbon-rich Cretaceous sequences. In Ginsburg, R.N., and Beaudoin, B. (Eds.), Cretaceous Resources, Events and Rhythms: Background and Plans for Research. Springer, New York, 75119.Google Scholar
Arzmüller, G., Buchta, S., Ralbovský, E., and Wessely, G., 2006. The Vienna Basin. In: Golonka, J., and Pícha, F. (Eds), The Carpathians and Their Foreland. Geology and Hydrocarbon Resources. AAPG, Washington, DC, pp. 191204.Google Scholar
Asim, S., Qureshi, N. S., Mirza, Q., et al., 2014. Structural and stratigraphical interpretation of seismic profiles along Safed Koh Trend (Eastern Part of Sulaiman Fold Belt), Pakistan. Universal Journal of Engineering Science, 2(4): 7795.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 the Azores. Geochemistry, Geophysics, Geosystems, 5: Q01E16.CrossRefGoogle Scholar
Aswathanarayana, U., 1964. Isotopic ages from Eastern Ghat and Cuddapah of India. Journal of Geophysical Research, 69: 34793486.CrossRefGoogle Scholar
Athy, L. F., 1930. Density, porosity, and compaction of sedimentary rocks. AAPG, 14: 124.Google Scholar
Atwater, T. M., 1970. Implications of plate tectonics for the Cenozoic Tectonic evolution of western North America. Geological Society of America Bulletin, 81 (12): 35133536.CrossRefGoogle Scholar
Atwater, T. M., and Stock, J., 1998. Pacific North America plate tectonics of the Neogene southwestern United States: an update. International Geology Review, 40: 375402.CrossRefGoogle Scholar
Audemard, F. A., Romero, G., Rendon, H., and Cano, V., 2005. Quaternary fault kinematics and stress tensors along the southern Caribbean from fault-slip data and focal mechanism solutions. Earth-Science Reviews, 69: 181233.CrossRefGoogle Scholar
Auzende, J. M., Bideau, D., Bonatti, E., et al., 1989. Direct observation of a section through slow-spreading oceanic crust. Nature, 337: 726729.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
Avni, Y., Segev, A., and Ginat, H., 2012. Oligocene regional denudation of the northern Afar dome: pre- and syn-breakup stages of the Afro-Arabian plate. Geological Society of America Bulletin, 124: 18711897.CrossRefGoogle Scholar
Axen, G. J., 1995. Extensional segmentation of the Main Gulf Escarpment, Mexico and United States. Geology, 23: 515518.2.3.CO;2>CrossRefGoogle Scholar
Aydin, A., 1977. Small faults formed as deformation bands in sandstone. In: Evernden, J. F. (Ed.), Proceedings of Conference II: Experimental Studies of Rock Friction with Application to Earthquake Prediction. US Geological Survey, Menlo Park, pp. 617653.Google Scholar
Aydin, A., 1978a. Faulting in sandstone. Doctoral dissertation. Stanford University, Stanford.Google Scholar
Aydin, A., 1978b. Small faults formed as deformation bands in sandstone. Pure and Applied Geophysics, 116: 913930.CrossRefGoogle Scholar
Aydin, A., and Berryman, J. 2010. Analysis of the growth of strike-slip faults using effective medium theory. Journal of Structural Geology, 32(11): 16291642.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: 931942.CrossRefGoogle 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: Biddle, K.T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation, and Sedimentation. Society of Economic Paleontologists and Mineralogists, Denver, pp. 3544.CrossRefGoogle Scholar
Aydin, A., Borja, R. I., and Eichhubl, P., 2003. Geological and Mathematical Framework for Failure Modes in Granular Rock. Geological Society of London, London.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: 257284.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. AAPG Bulletin, 79: 116129.Google Scholar
Bacoccoli, G., 1991, O futuro da exploração de petróleo na Bacia de Sergipe-Alagoas: Relatório Interno Petrobrás, Depex, Rio de Janeiro, 2 vols.Google Scholar
Bakker, R. J., 2003. Package FLUIDS 1: computer programs for analysis of fluid inclusion data and for modeling bulk fluid properties. Chemical Geology, 194: 323.CrossRefGoogle Scholar
Bakun, W. H., and Lindh, A. G., 1985. The Parkfield, California earthquake prediction experiment. Science, 229: 619624.CrossRefGoogle ScholarPubMed
Ballard, J., 2019. Delineation of submarine fan systems in the Guyana/Suriname Basin. Available at: www.researchgate.net/publication/339365381_Delineation_of_Submarine_Fan_Systems_in_the_GuyanaSuriname_Basin, accessed October 19, 2020.Google Scholar
Bally, A. W., 1982. Musings over sedimentary basin evolution. Philosophical Transactions of the Royal Society of London, Series A: Mathematical and Physical Sciences, 305: 325338.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., et al., 2002. Digital data grids for the magnetic anomaly map of North America: Nova Scotia aeromagnetic field map. OF 02-0414, US Geological Survey.Google Scholar
Banks, C. J., and Warburton, J., 1986. “Passive-roof” duplex geometry in the frontal structures of the Kirthar and Sulaiman belts, Pakistan. Journal of Structural Geology, 8: 229237.CrossRefGoogle Scholar
Barka, A. A., 1992. The North Anatolian Fault Zone. Annales Tectonicae, 6: 164195.Google Scholar
Barka, A. A., and Hancock, P. L., 1984. Neotectonic deformation patterns in the convex-northwards arc of the North Anatolian fault zone. In: Dixon, J. E. and Robertson, A. H. F. (Eds.), The Geological Evolution of the Eastern Mediterranean Region. Geological Society, London, pp. 763773.Google Scholar
Barka, A. A., and Kadinsky-Cade, K., 1988. Strike-slip fault geometry in Turkey and its influence on earthquake activity. Tectonics 7(3): 663684.CrossRefGoogle Scholar
Barker, C. E., 1995, Salton Trough Province. In: Gautier, D. L., Dolton, G. L., Takahashi, K. I., and Varnes, K. L. (Eds.), National Assessment of United States Oil and Gas Resources: Results, Methodology and Supporting Data, US Geological Survey, Reston, 30.Google Scholar
Barrier, E., Huchon, P., and Aurelio, M., 1991. Philippine Fault: a key for Philippine kinematics. Geology, 19: 3235.2.3.CO;2>CrossRefGoogle Scholar
Bartlett, W. L., Friedman, M., and Logan, J. M., 1981. Experimental folding and faulting of rocks under confining pressure. Part IX, Wrench faults in limestone layers. Tectonophysics, 79: 255277.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(9): 10131016.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(B4): 51815200.CrossRefGoogle Scholar
Barton, C. A., Zoback, M. D., and Moos, D., 1995. Fluid flow along potentially active faults in crystalline rock. Geology, 23: 683686.2.3.CO;2>CrossRefGoogle Scholar
Basak, P., 1972. Soil structure and its effects on hydraulic conductivity. Soil Science, 114: 417422.CrossRefGoogle Scholar
Basile, C., 2015. Transform continental margins – Part 1: concepts and models. Tectonophysics, 661: 110.CrossRefGoogle Scholar
Basile, C., and Allemand, P., 2002. Erosion and flexural uplift along transform faults. Geophysical Journal International, 151: 646653.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: Proceedings of the Ocean Drilling Program: Scientific Results, Volume 159. Texas A&M University, College Station, pp. 101110.Google Scholar
Basile, C., Mascle, J., Popoff, M., Bouillin, J. P., and Mascle, G., 1993. The Côte d’Ivoire–Ghana transform margin: a marginal ridge structure deduced from seismic data. Tectonophysics, 222: 119.CrossRefGoogle Scholar
Basile, C., Maillard, A., Patriat, M., et al., 2013. Structure and evolution of the Demerara Plateau, offshore French Guiana; rifting, tectonic inversion and post-rift tilting at transform-divergent margins intersection. Tectonophysics, 591: 1629.CrossRefGoogle 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(2): 159160.Google Scholar
Bates, R. L., and Jackson, J. A., 1987. Glossary of Geology, 3rd edition. American Institute, Alexandria.Google Scholar
Bawden, G. W., Michael, A. J., and Kellog, L. H., 1999. Birth of a fault: connecting the Kern County and Walker Pass, California, earthquakes. Geology, 27: 601604.2.3.CO;2>CrossRefGoogle Scholar
Bayona, G., Cardona, A., Jaramillo, C., et al., 2013. Onset of fault reactivation in the Eastern Cordillera of Colombia and proximal Llanos basin; response to Caribbean–South American convergence in early Palaeogene time. In: Nemčok, M., Mora, A., and Cosgrove, J. W. (Eds.), Thick-Skin-Dominated Orogens: From Initial Inversion to Full Accretion. Geological Society of London, London, pp. 285314.Google Scholar
Bear, J., 1972. Dynamics of Fluids in Porous Media. Elsevier, New York.Google Scholar
Beardsmore, G. R., and Cull, J. P., 2001. Crustal Heat Flow. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Beaubouef, R. T., and Friedmann, S. J., 2000, High resolution seismic/sequence stratigraphic framework for the evolution of the Pleistocene intra slope basins, western Gulf of Mexico: depositional models and reservoir analogs. Gulf Coast Section SEPM (Society of Sedimentary Geology) 20th Annual Research Conference Proceedings, pp. 40–60.CrossRefGoogle Scholar
Beaumont, C., Kooi, H., and Willett, S., 2000. Coupled tectonic-surface process models with applications to rifted margins and collisional orogens. In: Summerfield, M. A. (Ed.), Geomorphology and Global Tectonics, Wiley, Chichester, pp. 355.Google Scholar
Becken, M., Ritter, O., Park, S. K., et al., 2008. A deep crustal fluid channel into the San Andreas Fault system near Parkfield, California. Geophysical Journal, 173(2): 718732.CrossRefGoogle Scholar
Becken, M., Ritter, O., Park, S. K., Bedrosian, P. A., and Weckmann, U., 2011. Correlation between deep fluids, tremor and creep along the central San Andreas fault. Nature, 480(7375): 8790.CrossRefGoogle ScholarPubMed
Becker, K., and Fisher, A., 2000. Permeability of upper oceanic basement on the eastern flank of the Endeavor Ridge determined with drill-string packer experiments. Journal of Geophysical Research, 105: 897912.CrossRefGoogle Scholar
Becker, K., Langseth, M., Von Herzen, R. P., and Anderson, R., 1983. Deep crustal geothermal measurements, Hole 504B, Costa Rica Rift. Journal of Geophysical Research, 88: 34473457.CrossRefGoogle Scholar
Bednaříková, J. (Ed.), 1984. Oil industry on Czechoslovak territory (in Czech). Volume 5. Knihovnička Zemního plynu a nafty, Hodonín.Google Scholar
Bedrosian, P. A., Unsworth, M. J., and Wang, F., 2001. Structure of the Altyn Tagh Fault and Daxue Shan from magnetotelluric surveys: implications for faulting associated with the rise of the Tibetan Plateau. Tectonics, 20(4): 474486.CrossRefGoogle 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(11).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: 137158.CrossRefGoogle Scholar
Begnaud, M. L., McCain, J. S., Barth, G. A., Orcutt, J. A., and Harding, A. J., 1997. Velocity structure from forward modeling of the eastern ridge-transform intersection area of the Clipperton Fracture Zone, East Pacific Rise. Journal of Geophysical Research, 102: 78037820.CrossRefGoogle Scholar
Behn, M. D., Boettcher, M. S., and Hirth, G., 2007. Thermal structure of oceanic transform faults. Geology, 35. 307310.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: 1814518158.CrossRefGoogle Scholar
Bellahsen, N., Leroy, S., Autin, J., et al., 2013. Pre-existing oblique transfer zones and transfer/transform relationships in continental margins: new insights from the southeastern Gulf of Aden, Socotra Island, Yemen. Tectonophysics, 607: 3250.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: 109196.Google Scholar
Benada, S., 1986. New results on the extent of the para-autochthonous thrust sheet with Karpatian sediments in the central portion of the Carpathian Neogene foredeep in Moravia (in Czech). Zemní Plyn a Nafta, 4: 485492.Google Scholar
Ben-Avraham, Z., 1985. Structural framework of the Gulf of Elat (Aqaba), northern Red Sea. Journal of Geophysical Research, 90: 703726,CrossRefGoogle Scholar
Ben-Avraham, Z., 2014. Geophysical studies of the crustal structure along the Southern Dead Sea Fault. In: Garfunkel, Z., Ben-Avraham, Z., and Kagan, E. (Eds.), Dead Sea Transform Fault System: Reviews. Springer, New York, pp. 127.Google Scholar
Ben-Avraham, Z., and Garfunkel, Z., 1986. Character of transverse faults in the Elat pull-apart basin. Tectonics, 5: 7, 11611169.CrossRefGoogle Scholar
Ben-Avraham, Z., and Tibor, G., 1993. The northern edge of the Gulf of Elat. Tectonophysics, 226: 319331.CrossRefGoogle Scholar
Ben-Avraham, Z., and Zoback, M. D., 1992. Transform-normal extension and asymmetric basins: an alternative to pull-apart models. Geology, 20(5): 423426.2.3.CO;2>CrossRefGoogle Scholar
Ben-Avraham, Z., Hanel, R., and Villinger, H., 1978. Heat flow through the Dead sea rift. Marine Geology, 28: 253267CrossRefGoogle Scholar
Ben-Avraham, Z., Almagor, G., and Garfunkel, Z., 1979. Sediments and structure of the Gulf of Elat (Aqaba): northern Red Sea. Sedimentary Geology, 23: 239267.CrossRefGoogle Scholar
Ben-Avraham, Z., Ginzburg, A., and Yuval, Z., 1981. Seismic reflection and refraction investigations of Lake Kinneret-central Jordan Valley, Israel. Tectonophysics, 80: 165181.CrossRefGoogle Scholar
Ben-Avraham, Z., Hartnady, C., and Malan, J., 1993. Early tectonic extension between the Agulhas Bank and the Falkland Plateau due to the rotation of the Lafonia microplate. Earth and Planetary Science Letters, 117: 4358.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., Lazar, M., Garfunkel, Z., et al., 2012. Structural styles along the Dead Sea Fault. In: Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps. Elsevier, Amsterdam.Google Scholar
Bender, J. F., Hodges, F. N., and Bence, A. E., 1978. Petrogenesis of basalts from project Famous Area: experimental study from 0-kbar to 15-kbars. Earth and Planetary Science Letters, 41: 277302.CrossRefGoogle Scholar
Bender, J. F., Langmuir, C. H., and Hanson, G. N., 1984. Petrogenesis of basalt glasses from the Tamayo region, East Pacific Rise. Journal of Petrology, 25: 213254.CrossRefGoogle Scholar
Bengtsson, L., and Enell, M., 1986. Chemical analysis. In: Berglund, B. E. (Ed.), Handbook of Holocene Palaeoecology and Palaeohydrology, Wiley, Chichester, pp. 423451.Google 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(1): 29128.Google Scholar
Benkhelil, J., Mascle, J., and Huguen, C., 1998a. Deformation patterns and tectonic regimes of the Côte d’Ivoire–Ghana Transform Margin as deduced from Leg 159 results. In: Proceedings of the Ocean Drilling Program: Scientific Results, Volume 159. Texas A&M University, College Station, pp. 1323.Google Scholar
Benkhelil, J., Mascle, J., and Guiraud, M., 1998b. Sedimentary and structural characteristics of the Cretaceous along the Côte d’Ivoire-Ghana Transform Margin and in the Benue Trough: a comparison. In: Proceedings of the Ocean Drilling Program: Scientific Results, Volume 159. Texas A&M University, College Station, pp. 9399.Google Scholar
Bennett, K., and Rusk, D. C., 2002. Regional 2D seismic interpretation and exploration potential of offshore deepwater Sierra Leone and Liberia, West Africa. The Leading Edge, November 2002: 1118–1124.CrossRefGoogle Scholar
Berg, R. R., 1975. Capillary pressure in stratigraphic traps. AAPG Bulletin, 59: 939956.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. AAPG Bulletin, 83: 727756.Google Scholar
Bergerat, F., Angelier, J., and Homberg, C., 2000. Tectonic analysis of the Husavík-Flatey Fault (northern Iceland) and mechanisms of an oceanic transform zone, the Tjörnes Fracture Zone. Tectonics, 6: 11611177.CrossRefGoogle Scholar
Bergman, E. A., and Solomon, S. C., 1988. Transform fault earthquakes in the North Atlantic: source mechanisms and depth of faulting. Journal of Geophysical research, 93(B8): 90279057.CrossRefGoogle Scholar
Bernard, M., Shen-Tu, B., Holt, W. E., and Davis, D. M., 2000. Kinematics of active deformation in the Sulaiman Lobe and Range, Pakistan. Journal of Geophysical Research, 105(I B): 1325313279.CrossRefGoogle Scholar
Berndt, M. E., Allen, D. E., and Seyfried, W. E., 1996. Reduction of CO2 during serpentinization of olivine at 300° C and 500 bar. Geology, 24: 351354.2.3.CO;2>CrossRefGoogle Scholar
Bertani, R. T., Coast, I. G., and Matos, R. M. D., 1990, Evolução tectono-sedimentar, estilo estrutural e habitat do petróleo na Bacia Potiguar. In: Raja Gabaglia, G. P., and Milani, E. J. (Eds.), Origem e Evolução de Bacias Sedimentares. Petrobras, Rio de Janeiro, pp. 291301.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: Ristedt, H. (Ed.), First Eurocolloquium of the Ocean Drilling Program, Oldenburg, pp. 2122.Google Scholar
Best, P., et al., 1985. Ghana Project: mid term report. Unpublished. Teknica Ltd.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
Beyer, L. A., 1995. Geologic report of the Los Angeles Basin Province (014) for the 1995 National Assessment of Oil and Gas. In: Los Angeles Basin 1995 Assessment. United States Assessments of Undiscovered Oil and Gas Resources. Available at: https://certmapper.cr.usgs.gov/data/noga95/prov14/text/prov14.pdf, accessed November 1, 2020.Google Scholar
Biddle, K. T., 1991. The Los Angeles Basin: an overview. In: Biddle, K. T. (Ed), Active Margin Basins. AAPG, Washington, DC, pp. 524.Google Scholar
Biddle, K.T., and Christie-Blick, N., 1985. Glossary: strike-slip deformation, basin formation, and sedimentation. In: Biddle, K. T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation, and Sedimentation. SEPM, Tulsa, pp. 375385.CrossRefGoogle Scholar
Bienaszewski, M., 2020. Dead Sea Transform. Final essay, internship in structural geology of continental transforms. Unpublished report. EGL Bratislava.Google Scholar
Bigot-Cormier, F., Basile, C., Poupeau, G., Bouillin, J. P., and Labrin, E., 2005. Denudation of the Cote d’Ivoire–Ghana transform margin from apatite fission tracks. Terra Nova, 17: 189195.CrossRefGoogle Scholar
Bílek, K., 1974. Oil and gas fields in the Slovak part of the Vienna Basin (in Czech). Mineralia Slovaca, 6(5): 399498.Google Scholar
Billia, M. A., Timms, N. E., Toy, V. G., Hart, R. D., and Prior, D. J., 2013. Grain boundary dissolution porosity in quartzofeldspathic ultramylonites: implications for permeability enhancement and weakening of mid-crustal shear zones. Journal of Structural Geology, 53: 214.CrossRefGoogle Scholar
Biot, M. A., 1941. General theory of three dimensional consolidation. Journal of Applied Physics, 12: 155164.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: 168178.CrossRefGoogle Scholar
Birch, F., and Clark, H., 1940. The thermal conductivity of rocks and its dependence upon temperature and composition. American Journal of Science, 238: 529558, 613–635.CrossRefGoogle Scholar
Bird, D., 2001. Shear margins: continent–ocean transform and fracture zone boundaries. The Leading Edge, February 2011, 150–159.CrossRefGoogle Scholar
Bird, P., Kagan, Y.Y., and Jackson, D.D., 2002. Plate tectonics and earthquake potential of spreading ridges and oceanic transform faults. In: Stein, S., and Freymueller, J. T. (Eds.), Plate Boundary Zones. Wiley, Chichester, pp. 203218.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. 6987.Google Scholar
Biswal, T. K., and Sinha, S., 2004. Fold-thrust-belt structure of the Proterozoic Eastern Ghats Mobile Belt: a proposed correlation between India and Antarctica in Gondwana. Gondwana Research, 7(1): 4356.CrossRefGoogle Scholar
Biteau, J. J., Choppin de Janvry, G., Perrodon, A., 2003. How the petroleum system relates to the petroleum province. Oil and Gas Journal, 101: 4649.Google Scholar
Bjarnason, I. T., and Einarsson, P., 1991. Source mechanism of the 1987 Vatnafjσll earthquake in south Iceland. Journal of Geophysical Research, 96: 43134324.CrossRefGoogle Scholar
Bjorlykke, K., Ramm, M., and Saigal, G. C., 1989. Sandstone diagenesis and porosity modification during basin evolution. Geologic Modeling: Aspects of Integrated Basin Analysis and Numerical Simulation, Geologische Rundschau, 78: 243268.Google Scholar
Blackman, D. K., and Forsyth, D. W., 1991. Isostatic compensation of tectonic features of the Mid-Atlantic Ridge: 20°27° 30′S. Journal of Geophysical Research, 96: 1174111758.CrossRefGoogle Scholar
Blackwell, D. D., and Richards, M. C., 2004. Geothermal map of North America. Scale 1: 6, 500, 000. AAPG, Washington, DC.Google Scholar
Blackwell, D. D., and Steele, J. L., 1989. Thermal conductivity of sedimentary rocks: measurement and significance. In: Naeser, N. D., and McCulloh, T. H. (Eds.), Thermal History of Sedimentary Basins: Methods and Case Histories. Springer, Heidelberg, pp. 1436.Google Scholar
Blaich, O. A., Faleide, J. I., and Tsikalas, F., 2011. Crustal breakup and continent–ocean transition at South Atlantic conjugate margins. Journal of Geophysical Research 116: B01402.CrossRefGoogle Scholar
Blanc, P., and Connan, J., 1994. Preservation, degradation, and destruction of trapped oil. In: Magoon, L. B., and Dow, W. G. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 237247.Google Scholar
Bloch, R. B., Huene, R. V., Hart, P. E., and Wentworth, C. M., 1993. Style and magnitude of tectonic shortening normal to the San Andreas fault across Pyramid Hills and Kettleman Hills South Dome, California. Geological Society of America Bulletin, 105, 464478.2.3.CO;2>CrossRefGoogle Scholar
Bock, Y., McCaffrey, R., Rais, J., and Murata, I., 1990. Geodetic studies of oblique plate convergence in Sumatra. Eos Transactions, AGU, 71: 857.Google Scholar
Bodnar, R. J., 2003. Introduction to fluid inclusions. In: Samson, I., Anderson, A., and Marshall, D. (Eds), Fluid Inclusions: Analysis and Interpretation. Mineralogical Association of Canada, Quebec City, pp. 18.Google Scholar
Boehm, A., and Moore, J. C., 2002. Fluidized sandstone intrusions as an indicator of paleostress orientation, Santa Cruz, California. Geofluids, 2: 147161.CrossRefGoogle Scholar
Boettcher, M. S., and Jordan, T. H., 2004. Earthquake scaling relations for mid-ocean ridge transform faults. Journal of Geophysical Research, 109(B12): B12302.CrossRefGoogle 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(B2): 16831701.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, I: 6574.Google 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. AAPG Bulletin, 88: 947970.CrossRefGoogle Scholar
Bonatti, E., 1976. Serpentinite protrusions in the oceanic crust. Earth and Planetary Science Letters, 32(2): 107113.CrossRefGoogle Scholar
Bonatti, E., 1978. Vertical tectonism in oceanic fracture zones. Earth and Planetary Science Letters, 37(3): 369379.CrossRefGoogle Scholar
Bonatti, E., and Honnorez, J., 1976. Sections of the Earth’s crust in the equatorial Atlantic. Journal of Geophysical Research, 81(23): 41044116.CrossRefGoogle Scholar
Bonatti, E., Lawrence, J. P., Hamlyn, P. R., and Breger, D., 1980. Aragonite from deep sea peridotites: temperature dependence on mineralogy and boron content. Earth and Planetary Science Letters, 70: 8894.CrossRefGoogle Scholar
Bonatti, E., Ligi, M., Gasperini, L., et al., 1994. Transform migration and vertical tectonics at the Romanche fracture zone, equatorial Atlantic. Journal of Geophysical Research, 99: 2177921802.CrossRefGoogle Scholar
Bonatti, E., Ligi, M., Brunelli, D., et al., 2003. Mantle thermal pulses below the Mid-Atlantic Ridge and temporal variations in the formation of oceanic lithosphere. Nature (London), 423: 499505.CrossRefGoogle ScholarPubMed
Bonatti, E., Brunelli, D., Buck, W. R., et al., 2005. Flexural uplift of a lithospheric slab near the Vema transform (Central Atlantic): timing and mechanisms. Earth and Planetary Science Letters, 240(3): 642655.CrossRefGoogle Scholar
Bonne, K. P. M., 2014. Reconstruction of the evolution of the Niger River and implications for sediment supply to the Equatorial Atlantic margin of Africa during the Cretaceous and the Cenozoic. In: Scott, R. A., Smyth, H. R., Morton, A. C., and Richardson, N. (Eds.), Sediment Provenance Studies in Hydrocarbon Exploration and Production. Geological Society of London, London, pp. 327349.Google Scholar
Booth, J. R., DuVernay, A. E., III, Pfeiffer, D. S., and Styzen, M. J., 2000, Sequence stratigraphic framework, depositional models, and stacking patterns of ponded and slope fan systems in the Auger Basin: Central Gulf of Mexico slope. Gulf Coast Section SEPM (Society of Sedimentary Geology) 20th Annual Research Conference Proceedings, pp. 82–103.CrossRefGoogle Scholar
Boschi, C., Bonatti, E., Ligi, M., et al., 2013. Serpentinization of mantle peridotites along an uplifted lithospheric section, Mid Atlantic Ridge at 11° N. Lithos, 178: 323.CrossRefGoogle Scholar
Bostick, N. H., 1979. Maturation of organic matter and generation of petroleum in Tertiary oil basins. US Geological Survey Professional Paper.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: 4147.CrossRefGoogle Scholar
Bosworth, W., 1985. Geometry of propagating continental rifts. Nature, 316(6029): 625627.CrossRefGoogle Scholar
Bosworth, W., 1986. Comment on detachment faulting and the evolution of passive continental margins. Geology, 14: 890891.2.0.CO;2>CrossRefGoogle Scholar
Bosworth, W., 1995. A high strain rift model for the southern Gulf of Suez. In: Lambiase, J.J. (Ed.), Hydrocarbon Habitat in Rift Basins. Geological Society of London, London, pp. 75102.Google Scholar
Bosworth, W., Huchon, P., and McClay, K., 2005. The Red Sea and Gulf of Aden Basins. Journal of African Earth Sciences, 43: 334378.CrossRefGoogle Scholar
Botz, R., Winckler, G., Bayer, R., et al., 1999. Origin of trace gases in submarine hydrothermal vents of the Kolbeinsey Ridge. Earth and Planetary Science Letters, 171: 8393.CrossRefGoogle 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. In: Proceedings of the Ocean Drilling Program: Scientific Results, Volume 159. Texas A&M University, College Station, pp. 4348.Google Scholar
Boulton, C., Carpenter, B. M., Toy, V. G., and Marone, C., 2012. Physical properties of surface outcrop cataclastic fault rocks, Alpine Fault, New Zealand. Geochemistry, Geophysics, Geosystems, 13(1). DOI: 10.1029/2011GC003872.CrossRefGoogle Scholar
Bouma, A. H., Normark, W. R., and Barnes, N. E., 1985. Submarine Fans and Related Turbidite Systems: Springer, New York.CrossRefGoogle 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. AAPG Bulletin, 91(2): 173190.CrossRefGoogle Scholar
Brace, W. F., Paulding, B. W., and Scholz, C., 1966. Dilatancy in the fracture of crystalline rocks. Journal of Geophysical Research, 71(16): 39393953.CrossRefGoogle Scholar
Bradley, C. H., 2011. The Tano Basin of Western Ghana: a complex, intriguing and prolific deepwater play. Houston Geological Society Bulletin, 53(8): 2729.Google Scholar
Bradley, C. H., and Fernandez, M. N., 1991. Early Cretaceous paleogeography of Gabon/northeastern Brazil: a tectono-stratigraphic model based on propagating rifts. In: Curnelle, R. (Ed.), Géologie Africaine Mémoire 13. Elf Aquitaine, Courbevoie, pp. 1730.Google Scholar
Bralower, T. J., and Thierstein, H. R., 1984. Low productivity and slow deep-water circulation in mid-Cretaceous oceans. Geology, 12: 614618.2.0.CO;2>CrossRefGoogle 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
Brannon, J. C., Cole, S. C., Podosek, F. A., et al., 1996. Th–Pb and U–Pb dating of ore-stage calcite and Paleozoic fluid flow. Science, 271: 491493.CrossRefGoogle Scholar
Braun, J., and Beaumont, C., 1989. A physical explanation of the relation between flank uplifts and the breakup unconformity at rifted continental margins. Geology (Boulder), 17(8): 760764.2.3.CO;2>CrossRefGoogle Scholar
Braunmiller, J., and Nabelek, J., 2008. Segmentation of the Blanco Transform Fault Zone from earthquake analysis: complex tectonics of an oceanic transform fault. Journal of Geophysical research, 113(B07108). DOI: 10.1029/2007JB005213.CrossRefGoogle Scholar
Breddam, K., Kurz, M. D., and Storey, M., 2000. Mapping out the conduit of the Iceland mantle plume with helium isotopes. Earth and Planetary Science Letters, 176: 4555.CrossRefGoogle 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: National Research Council (Ed.), The Role of Fluids in Crustal Processes. National Academy Press, Washington, DC, pp. 158164.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: 325328.CrossRefGoogle Scholar
Bredehoeft, J. D., Djevanshir, R. D., and Belitz, K. R., 1988. Lateral fluid flow in a compacting sand–shale sequence: South Caspian Basin. AAPG Bulletin, 72: 416424.Google Scholar
Breivik, A. J., Verhoef, J., and Faleide, J. I., 1999. Effect of thermal contrasts on gravity modeling at passive margins: results from the western Barents Sea. Journal of Geophysical Research, 104: 1529315311.CrossRefGoogle Scholar
Bremner, J. M., 1949. Use of the Van Slyke-Neil manometric apparatus for the determination of organic and inorganic carbon in soil and of organic carbon in soil extracts. Analyst 74: 492498.CrossRefGoogle Scholar
Brew, G., Barazangi, M., Al-Maleh, A. K., and Sawaf, T., 2001a. Tectonic and geologic evolution of Syria. GeoArabia, 6(4): 573616.CrossRefGoogle Scholar
Brew, G., Lupa, J., Barazangi, M., et al., 2001b. Structure and tectonic development of the Ghab basin and the Dead Sea fault system, Syria. Journal of the Geological Society of London, 158: 665674.CrossRefGoogle 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. AAPG Bulletin, 84: 12051228.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. AAPG Bulletin, 74: 14591477.Google Scholar
Brito Neves, B. B., Campos Neto, M. C., and Fuck, R., 1999. From Rodinia to Western Gondwana: an approach to the Brasiliano-Pan African cycle and orogenic collage. Episodes, 22: 155166.CrossRefGoogle Scholar
Brix, F., and Schultz, O., 1993. Erdöl und Erdgas in Osterreich. 2. Auflage. Veroffentlichungen aus dem Naturhistorischen Museum Wien, N. F. 19, XXIV + 688 S., 200 Abb., 17 Beil., Wien.Google Scholar
Bronner, A., Sauter, D., Manatschal, G., Péron-Pinvidic, G., and Munschy, M., 2011. Magmatic breakup as an explanation for magnetic anomalies at magma-poor rifted margins. Nature Geoscience, 4: 549553.CrossRefGoogle Scholar
Brooks, M., Trayner, P. M., and Trimble, T. J., 1988. Mesozoic reactivation of Variscan thrusting in the Bristol Channel area, U. K. Journal of the Geological Society, London 145: 439444.CrossRefGoogle Scholar
Brooks, M., Hillier, B. V., and Miliorizos, M., 1993. New seismic evidence for a major geological boundary at shallow depth under north Devon. Journal of the Geological Society, London, 150: 131135.CrossRefGoogle Scholar
Brown, A., Birkhead, S., McLean, D., et al., 2015. The Campanian quartz “claystone” conundrum of the African Transform Margin. In: Sabato Ceraldi, T., Hodgkinson, R. A., and Backe, G. (Eds.), Petroleum Geoscience of the West African Margin. Geological Society of London, London, pp. 2748.Google 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: 26.Google Scholar
Brown, K. M., and Moore, J. C., 1993. Comment on “Anisotropic permeability and tortuosity in deformed wet sediments” by J. Arch and A. J. Maltman. Journal of Geophysical Research, 98: 1785917864.CrossRefGoogle Scholar
Brown, M., Diaz, F., and Grocott, J., 1993. Displacement history of the Atacama fault system 25°00′S–27°00′S, northern Chile. GSA Bulletin, 105: 11651174.2.3.CO;2>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 (Boulder), 19(1): 7477.2.3.CO;2>CrossRefGoogle 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(3): 339350.CrossRefGoogle Scholar
Brown, R. W., Gallagher, K., Gleadow, A.JW., and Summerfield, M.A., 2000. Morphotectonic evolution of the South Atlantic margins of Africa and South America. In: Summerfield, M. A. (Ed.), Geomorphology and Global Tectonics. Wiley, Chichester, pp. 255281.Google Scholar
Brown, R. W., Summerfield, M. A., and Gleadow, A. J. W., 2002. Denudational history along a transect across the Drakenberg Escarpment of southern Africa derived from apatite fission track thermochronology. Journal of Geophysical Research, 107(B12): 2350.CrossRefGoogle Scholar
Brownfield, M., and Charpentier, R. R., 2003. Assessment of the undiscovered oil and gas of the Senegal Province, Mauritania, Senegal, the Gambia, and Guinea-Bissau, Northwest Africa. US Geological Survey Bulletin, 2207–A: 125.Google Scholar
Brozena, J. M., and White, R. S., 1990. Ridge jumps and propagations in the South Atlantic Ocean. Nature, 348(6297): 149152.CrossRefGoogle Scholar
Brun, J. P., 1999. Narrow rifts versus wide rifts: inferences for the mechanics of rifting from laboratory experiments. Philosophical Transactions of the Royal Society of Mathematical, Physical and Engineering Sciences, 357(1753): 695712.CrossRefGoogle Scholar
Brun, J. P., and Beslier, M. O., 1996. Mantle exhumation at passive margins. Earth and Planetary Science Letters, 142(1–2): 161173.CrossRefGoogle Scholar
Brun, J. P., Sokoutis, D., and van den Driessche, J., 1994. Analogue modeling of detachment fault systems and core complexes. Geology (Boulder), 22(4): 319322.2.3.CO;2>CrossRefGoogle Scholar
Brune, J. N., 1968. Seismic moment, seismicity, and rate of slip along major fault zones. Journal of Geophysical Research, 73: 777784.CrossRefGoogle Scholar
Brune, S., Popov, A. A., and Sobolev, S. V., 2012. Modeling suggests that oblique extension facilitates rifting and continental break-up. Journal of Geophysical Research, 117: B08402.CrossRefGoogle Scholar
Brune, S., Heine, C., Perez-Gussinye, M., and Sobolev, S. V., 2014. Rift migration explains continental margin asymmetry and crustal hyper-extension. Nature Communication, 5. doi: 10.1038/ncomms5014.CrossRefGoogle ScholarPubMed
Brune, S., Williams, S. E., Butterworth, N. P., and Muller, R. D., 2016. Abrupt plate accelerations shape rifted continental margins. Nature Letter. DOI: 10.1038/nature18319.CrossRefGoogle Scholar
Brunel, M., 1980. Quartz fabrics in shear-zone mylonites: evidence for a major imprint due to late strain increments. Tectonophysics, 64(3): T33T44.CrossRefGoogle 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.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(2): 213234.CrossRefGoogle Scholar
Buck, W. R., Lavier, L. L., and Poliakov, A. N. B., 2005, Modes of faulting at mid-ocean ridges: Nature, 434: 719723.CrossRefGoogle ScholarPubMed
Buday, T., 1956. Report about the geological mapping of the SE part of the Dolnomoravský Úval (in Czech). Zprávy o geologickém výzkumu, 5–7.Google Scholar
Buday, T., 1965. Tectogenesis and architecture of the Neogene basins of the West Carpathians [in German]. In: Buday, T., Cícha, I., and Seneš, J. (Eds.), Miocene West Carpathian Strata [in German]. GÚDŠ, Bratislava, pp. 169250.Google Scholar
Buday, T., Cambel, B., Maheľ, M., et al., 1962. Explanations to the Geological map of the Czechoslovak Socialistic Republic (in Czech), scale 1:200, 000, sheets M-33-XXXV, M33-XXXVI, Wien-Bratislava. Archive Geofond, Bratislava, pp. 122140.Google Scholar
Buiter, S. J. H., and Torsvik, T. H., 2014. A review of Wilson Cycle plate margins: a role for mantle plumes in continental break-up along sutures? Gondwana Research, 26: 627653.CrossRefGoogle Scholar
Buntebarth, G., 1984. Geothermics: An Introduction. Springer, Berlin.CrossRefGoogle Scholar
Buntebarth, G., and Rybach, L., 1981. Linear relationships between petrophysical properties and mineralogical constitution: preliminary results. Tectonophysics, 75: 4146.CrossRefGoogle Scholar
Burbank, D. W., 1983. The chronology of intermontane-basin development in the northwestern Himalaya and the evolution of the Northwest Syntaxis. Earth and Planetary Science Letters, 64: 7792.CrossRefGoogle Scholar
Burbank, D.W., and Beck, R.A., 1989. Early Pliocene uplift of the Salt Range: temporal constraints on thrust wedge development, northwest Himalaya, Pakistan. In: Malinconico, L. L., and Lillie, R. J. (Eds.), Tectonics of the Western Himalayas. Geological Society of America, New York, 113128.CrossRefGoogle Scholar
Burbank, D. W., and Raynolds, R. G. H., 1988. Stratigraphic keys to the timing of deformation: an example from the northwestern Himalayan foredeep. In: Paola, C., and Kleinspehn, K. (Eds.), New Perspectives in Basin Analysis. Springer, New York, 331351.CrossRefGoogle Scholar
Burbank, D. W., Raynolds, R. G., and Johnson, G. D., 1986. Late Cenozoic tectonics and sedimentation in the north-western Himalayan foredeep: II. Eastern limb of the northwest syntaxis and regional synthesis. In: Allen, P. and Homewood, P. (Eds.), Foreland Basins. International Association of Sedimentologists, Algiers, pp. 293306.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(4): 439441.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(B10): 1042210426.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: 12331261.Google Scholar
Burg, J.-P., 2018. Script to tectonics. Educational material, ETH Zurich Research Collection, https://doi.org/10.3929/ethz-b-000279495 (accessed February 19, 2019).CrossRefGoogle 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: 217222.CrossRefGoogle Scholar
Bürgmann, R., and Dresen, G., 2008. Rheology of the lower crust and upper mantle: evidence from rock mechanics, geodesy, and field observations. Annual Reviews in Earth and Planetary Sciences, 36: 531567.CrossRefGoogle Scholar
Bürgman, R., Ergintav, S., Segall, P., et al., 2002. Time-dependent distributed afterslip on and deep below the Izmit earthquake rupture. Bulletin of the Seismological Society of America, 92: 126137.CrossRefGoogle Scholar
Burkhard, M., 1993. Calcite twins, their geometry, appearance and significance as stress–strain markers and indicators of tectonic regime: a review. Journal of Structural Geology, 15: 351368.CrossRefGoogle 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., and Sweeney, J. J., 1989. A chemical kinetic model of vitrinite reflectance maturation. Geochimica et Cosmochimica Acta, 53: 26492657.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: 931939.CrossRefGoogle 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(1–2): 726.CrossRefGoogle Scholar
Burov, E., and Gerya, T., 2014. Asymmetric three-dimensional topography over mantle plumes. Nature, 513: 8589.CrossRefGoogle ScholarPubMed
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: 1646116481.CrossRefGoogle 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. AAPG Bulletin, 78: 1137.Google Scholar
Busby, C. J., 2013, Birth of a plate boundary at ca. 12 Ma in the ancestral Cascades arc, and implications for transtensional rift propagation, Walker Lane. Geosphere, 9: 11471160.CrossRefGoogle Scholar
Busby, C. J., Putirka, K., Melosh, B., et al., 2018. A tale of two Walker Lane pull-apart basins in the ancestral Cascades arc, central Sierra Nevada, California. Geosphere, 14: 20682117.CrossRefGoogle Scholar
Busby, C., Graettinger, A., López Martínez, M., et al., 2020. Volcanic record of the arc-to-rift transition onshore of the Guaymas basin in the Santa Rosalía area, Gulf of California, Baja California. Geosphere, 16: 130.CrossRefGoogle Scholar
Butler, R. W. H., and Bowler, S., 1995. Local displacement rate cycles in the life of a fold-thrust belt. Terra Nova, 7: 408416.CrossRefGoogle Scholar
Caby, R., Buscail, F., Dembele, D, et al., 2008. Neoproterozoic garnet-glaucophanites and eclogites: new insights for subduction metamorphism along the Gourma Fold and Thrust Belt (western Mali). In: Ennih, N., and Liegeois, J.-P. (Eds.) The Boundaries of the West African Craton. Geological Society of London, London, pp. 203216.Google Scholar
Cacas, M. C., and de Marsily, G., 1987. Modelling of flow through fractured rocks. In: Tissot, B. (Ed.), Migration of Hydrocarbons in Sedimentary Basins. Collection Colloques Et Seminaires. Institut Francais Du Petrole, Technip, Paris, pp. 555566.Google Scholar
Caine, J. S., 1999. The Architecture and Permeability Structure of Brittle Fault Zones. University of Utah, Salt Lake City.Google Scholar
Caine, J. S., Evans, J. P., and Forster, C. B., 1996. Fault zone architecture and permeability structure. Geology, 24: 10251028.2.3.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: 206216.CrossRefGoogle Scholar
Cainelli, C. et al., 1986. Análise de Bacia do Pará-Maranhão, 6 vols.Google Scholar
Calais, E., 2018. Seismotectonic map of the Caribbean Plate. www.nsf.gov/news/mmg/media/images/Caribbean-gps-h, accessed April 30, 2018.Google Scholar
Callot, J. P., Grigne, C., Geoffroy, L., and Brun, J-P., 2001. Development of volcanic margins: two-dimensional laboratory models. Tectonics, 20: 148159.CrossRefGoogle Scholar
Callot, J. P., Geoffroy, L., and Brun, J.-P., 2002. Development of volcanic margins: three-dimensional laboratory models. Tectonics, 21(6): 1052.CrossRefGoogle Scholar
Camelbeeck, T., and Iranga, M. D., 1996. Deep crustal earthquakes and active faults along the Rukwa Trough, eastern Africa. Geophysical Journal International, 124: 612630.CrossRefGoogle Scholar
Campbell, K. A., and Bottjer, D. J., 1993. Fossil cold seeps. National Geographic Research and Exploration, 9: 326343.Google Scholar
Campos Neto, M. C., 2000. Oroenic systems from Southwestern Gondwana. In: Cordani, U. G., Milani, E. J., Thomaz Filho, A., and Campos, D. A. (Eds.), Tectonic Evolution of South America, 31st International Geological Congress, Rio de Janeiro, pp. 335365.Google Scholar
Canales, J. P., Collins, J. A., Escartín, J., and Detrick, R. S., 2000a. Seismic structure across the rift valley of the Mid-Atlantic Ridge at 23° 20′ (MARK area): implications for crustal accretion processes at slow spreading ridges. Journal of Geophysical Research 105: 2841128425.CrossRefGoogle Scholar
Canales, J. P., Detrick, R. S., Lin, J., Collins, J. A., and Toomey, D. R., 2000b. 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(B2): 26992720.CrossRefGoogle Scholar
Canales, J. P., Detrick, R., Toomey, D. R., and Wilcock, S. D., 2003. Segment-scale variations in the crustal structure of 150–300 kyr old fast spreading oceanic crust (East Pacific Rise, 8°15′N – 10°5′N) form wide-angle seismic refraction profiles. International Journal of Geophysics, 152: 766794.CrossRefGoogle Scholar
Cande, S. C., and Rabinowicz, P. D., 1979. Magnetic Anomalies of the Continental Margin of Brazil. AAPG, Washington, DC.Google 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(B11): 1347913492.CrossRefGoogle Scholar
Cann, J. R., Blackman, D. K., Smith, D. K., et al., 1997. Corrugated slip surfaces formed at ridge-transform intersections on the Mid-Atlantic Ridge. Nature, 385(6614): 329332.CrossRefGoogle Scholar
Cannat, M., 1993. Emplacement of mantle rocks in the seafloor at mid-ocean ridges. Journal of Geophysical Research, 98: 41634172.CrossRefGoogle Scholar
Cannat, M., Mével, C., Maia, M., et al., 1995. Thin crust, ultramafic exposures, and rugged faulting patterns at the Mid-Atlantic Ridge (22 degrees–24 degrees N). Geology, 23(1): 4952.2.3.CO;2>CrossRefGoogle Scholar
Cannat, M., Lagabrielle, Y., Bougault, H., et al., 1997. Ultramafic and gabbroic exposures at the Mid-Atlantic Ridge; geological mapping in the 15°N region. Tectonophysics 279: 193213.CrossRefGoogle Scholar
Cannat, M., Fontaine, F., and Escartín, J., 2010. Serpentinization and associated hydrogen and methane fluxes at slow spreading ridges. In: Rona, P., Devey, C., Dyment, J., and Murton, B. (Eds.), Diversity of Hydrothermal Systems on Slow Spreading Ocean Ridges. American Geophysical Union, Washington, DC, pp. 241264.CrossRefGoogle Scholar
Cao, W., Lee, C., Yang, J., and Zuza, A. V., 2019, Hydrothermal circulation cools continental crust under exhumation. Earth and Planetary Science Letters, 515: 248259.CrossRefGoogle Scholar
Capuano, R. M., 1993. Evidence of fluid flow in microfractures in geopressured shales. AAPG Bulletin, 77: 13031314.Google Scholar
Carbotte, S. M., and Macdonald, K. C., 1990. Causes of variation in fault-facing direction on the ocean floor. Geology 18: 749752.2.3.CO;2>CrossRefGoogle Scholar
Carey, S. W., 1958. The tectonic approach to continental drift. In: Carey, S. W. (Ed.), Continental Drift : A Symposium. University of Tasmania, Hobart, pp. 177363.Google Scholar
Carlsson, A., and Olsson, T., 1979. Hydraulic conductivity and its stress dependence. Proceedings of the Workshop on Low-Flow, Low Permeability Measurements in Largely Impermeable Rocks, Paris, pp. 249259.Google Scholar
Carpenter, B. M., Marone, C., and Saffer, D. M., 2011. Weakness of the San Andreas Fault revealed by samples from the active fault zone. Nature Geoscience, 4: 251254.CrossRefGoogle Scholar
Carslaw, H. S., and Jaeger, J. C., 1959. Conduction of Heat in Solids. Oxford University Press, New York.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(9): 13191326.CrossRefGoogle Scholar
Carvajal, C., Steel, R., and Petter, A., 2009. Sediment supply: the main driver of shelf margin growths. Earth Science Reviews, 96: 221248.CrossRefGoogle Scholar
Cassignol, C., Cornette, Y., David, B., and Gillot, P.-Y., 1978. Technologie potassium argon. C.E.N. Saclay. Rapp. CEA, R-4802: 37.Google Scholar
Castelltort, S., and Van Den Dreissche, J., 2003, How plausible are high-frequency sediment supply-driven cycles in the stratigraphic record? Sedimentary Geology, 157: 313.CrossRefGoogle Scholar
Catalan, L., Xiaowen, F., Chatzis, I., and Dullien, F. A. I., 1992. An experimental study of secondary oil migration. AAPG Bulletin, 76: 638650.Google Scholar
Catuneanu, O., 2002. Sequence stratigraphy of clastic systems: concepts, merits and pitfalls. Journal of African Earth Sciences, 35(1): 143.CrossRefGoogle Scholar
Čech, F., 1998: Hydrocarbon Fields in the West Carpathians [in Slovak]. Report, Archive VVNP Bratislava.Google Scholar
Cembrano, J., Gonzales, G., Arancibia, G., et al., 2005. Fault zone development and strain partitioning in an extensional strike-slip duplex: a case study from the Mesozoic Atacama fault system, Northern Chile. Tectonophysics, 400: 105125.CrossRefGoogle Scholar
Čermák, V., and Haenel, R., 1988. Geothermal maps. In: Rybach, L. and Stegena, L. (Eds.), Handbook of Terrestrial Heat-Flow Density Determination; with Guidelines and Recommendations of the International Heat Flow Commission. Kluwer Academic: Dordrecht.Google Scholar
Čermák, V., and Rybach, L., 1982. Thermal conductivity and specific heat of minerals and rocks. In: Angenheister, G. (Ed.), Landolt–Bornstein Numerical Data and Functional Relationships in Science and Technology. Springer, Berlin.Google Scholar
Čermák, V., Bodri, L., and Rybach, L., 1991. Radioactive heat production in the continental crust and its depth dependence. In: Čermák, V., and Rybach, L. (Eds.), Terrestrial Heat Flow and the Lithosphere Structure. Springer, Berlin, pp. 2369.CrossRefGoogle Scholar
Cerquiera, J. R., 1995. Geoquímica do Campo de Ubarana e adjacencies (Bacia Potiguar). Master’s thesis. Universidade Federal da Bahia, Salvador.Google Scholar
CGMW, 1990. Carte Géologique Internationale de l’Afrique, 1:5 000 000. Commission for the Geologic Map of the World.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
Chaimov, T. A., Barazangi, M., Al-Saad, D., Sawaf, T., and Gebran, A., 1990. Crustal shortening in the Palmyride Fold Belt, Syria, and implications for movement along the Dead Sea Fault System. Tectonics, 9(6): 13691386.CrossRefGoogle Scholar
Chang, H. K., Kowsmann, R. O., and de Figueiredo, A. M. F., 1988. New concepts on the development of East Brazilian marginal basins. Episodes, 11(3): 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
Chapin, C. E., 1989. Volcanism along the Socorro accommodation zone, Rio Grande rift, New Mexico. In: Chapin, C. E., and Zidek, J. (Eds.), Field Excursions to Volcanic Terranes in the Western United States. Volume I: Southern Rocky Mountain region, New Mexico Bureau of Mines and Mineral Resources, Socorro, pp. 4657.Google Scholar
Chapin, C. E., and Chater, S. M., 1994. Tectonic setting of the axial basins of the northern and central Rio Grande rift. In: Keller, G. R., and Chater, S. M. (Eds.), Basins of the Rio Grande Rift: Structure, Stratigraphy, and Tectonic Setting. Geological Society of America, New York, pp. 525.CrossRefGoogle Scholar
Chapman, D. S., 1986. Thermal gradients in the continental crust. In: Dawson, J. B., Carswell, D. A., Hall, J., and Wedepohl, K. H. (Eds.), The Nature of the Lower Continental Crust. Geological Society of London, London, pp. 6370.Google Scholar
Chapman, T. J., and Meneilly, A. W., 1990. Fault displacement analysis in seismic exploration. First Break, 8(1): 1112.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(B6): 96259642.CrossRefGoogle Scholar
Charlou, J.-L., Bougault, H., Appriou, P., Nelsen, T., and Rona, P., 1991. Different TDM/CH4 hydrothermal plume signatures: TAG site at 26°N and serpentinized ultrabasic diapir at 15°05′N on the Mid-Atlantic Ridge. Geochimica et Cosmochimica Acta, 55: 32093222.CrossRefGoogle Scholar
Chauvel, C., and Hemond, C., 2000. Melting of a complete section of recycled oceanic crust: trace element and Pb isotopic evidence from Iceland. G-cubed, 1. DOI: 10.1029/1999GC000002.Google Scholar
Cherkaoui, A. S. M., Wilcock, W. S. D., Dunn, R. A., and Toomey, D. R., 2003. A numerical model of hydrothermal cooling and crustal accretion at a fast spreading mid‐ocean ridge. Geochemistry, Geophysics, Geosystems, 4(9): 8616.CrossRefGoogle Scholar
Chery, J., Lucazeau, F., Daignieres, M., and Vilotte, J. P., 1992. Large uplift of rift flanks: a genetic link with lithospheric rigidity? Earth and Planetary Science Letters, 112(1–4): 195211.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: 621634.CrossRefGoogle Scholar
Chester, F. M., Evans, J. P., and Biegel, R. L., 1993. Internal structure and weakening mechanisms of the San Andreas Fault. Journal of Geophysical Research: Solid Earth, 98: (B1): 771786.CrossRefGoogle Scholar
Chester, J. S., Logan, J. S., and Spang, J. H., 1991. Influence of layering and boundary conditions on fault-bend and fault-propagation folding. Geological Society of America Bulletin, 103: 10591072.2.3.CO;2>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: 2423924253.CrossRefGoogle Scholar
Chian, D., Louden, K. E., Minshull, T. A., and Whitmarsh, R. B., 1999. Deep structure of the ocean–continent transition in the southern Iberia Abyssal Plain from seismic refraction profiles: 1. Ocean Drilling Program (Legs 149 and 173) transect. Journal of Geophysical Research, 104: 74437462.CrossRefGoogle Scholar
Chilingarian, G. V., 1983. Compactional diagenesis. In: Parker, A., and Sellwood, B. W. (Eds.), Sediment Diagenesis, Reidel Publishing Company, Dordrecht, pp. 57168.CrossRefGoogle Scholar
Chiozzi, P., Pasquale, M., and Verdoya, M., 2006. Seismicity and rheological modelling in an extensional-compressional tectonic realm. Geophysical Research Abstracts, 8.Google Scholar
Chorowicz, J., 1984. Cartografie géologique à partir d’images spatiales de la zone transformante Tanganyika-Rukwa-Malawi (Rift est-africain). 27éme Congrés Géologique International, Moscou. Géochronique, 10: 4849.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(2–4): 203214.CrossRefGoogle Scholar
Chorowicz, J., 2005. The East African rift system. Journal of African Earth Sciences, 43: 379410.CrossRefGoogle Scholar
Chorowicz, J., Le Fournier, J., Le Mut, C., et al., 1983. Observation par télédétection et au sol de mouvements décrochants NW–SE dextres dans le secteur transformant Tanganyika-Rukwa-Malawi du Rift Est-Africain. Comptes Rendus Académie des Sciences Paris 296 (II): 9961002.Google Scholar
Chorowicz, J., Le Fournier, J., and Vidal, G., 1987. A model for rift development in Eastern Africa. Geological Journal 22: 495513.CrossRefGoogle Scholar
Choubert, G. et al., 1988. International geological map of Africa. Commission for the Geological Map of the World and UNESCO.Google Scholar
Choudhuri, M., Nemčok, M., Stuart, C., et al., 2014. 85° E Ridge, India: constraints on its development and architecture. Journal of the Geological Society of India, 84: 513530.CrossRefGoogle Scholar
Choudhuri, M., Nemčok, M., Melichar, R., and Sinha, N., 2017. Propagation of hotspot volcanism driven flexure in oceanic crust-85°E ridge case study. Journal of Marine and Petroleum Geology. DOI: 10.1016/j.marpetgeo.2017.01.021.CrossRefGoogle 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: 623626.Google Scholar
Christensen, N. I., and Mooney, W. D., 1995. Seismic velocity structure and composition of the continental crust: a global view. Journal of Geophysical Research, 100(B6): 97619788.CrossRefGoogle Scholar
Christie-Blick, N., and Biddle, K. T., 1985. Deformation and basin formation along strike-slip faults. In: Biddle, K. T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation, and Sedimentation. Society of Economic Paleontologists and Mineralogists, Denver, pp. 134.Google Scholar
Cladouhos, T. T., and Marrett, R., 1996. Are fault growth and linkage models consistent with power-law distributions of fault lengths? Journal of Structural Geology, 18(2): 281293.CrossRefGoogle Scholar
Clapp, E. M., Bierman, P. R., and Caffee, M., 2002. Using 10Be and 26Al to determine sediment generation rates and identify sediment source areas in an arid drainage basin. Geomorphology, 45: 89104.CrossRefGoogle Scholar
Clark, S. P. Jr., 1966. Handbook of Physical Constants. Geological Society of America, New York.CrossRefGoogle 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: 383386.Google Scholar
Clauser, C., 2009. Heat transport processes in the Earth’s Crust. Surveys in Geophysics, 30: 163191.CrossRefGoogle Scholar
Clauser, C., 2011a. Radiogenic heat production of rocks. In: Gupta, H. (Ed.), Encyclopedia of Solid Earth Geophysics, 2nd edition, Springer, Berlin-Heidelberg, pp. 10181024.CrossRefGoogle Scholar
Clauser, C., 2011b. Thermal storage and transport properties of rocks. In: Gupta, H. (Ed.), Encyclopedia of Solid Earth Geophysics, 2nd edition, Springer, Berlin-Heidelberg, pp. 14231448.CrossRefGoogle 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. 465–470.CrossRefGoogle 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: 2525725274.CrossRefGoogle Scholar
Clift, P. D., Carter, A., and Hurford, A. J., 1998. Apatite fission track analysis of sites 959 and 960 on the transform margin of Ghana, West Africa. In: Proceedings of the Ocean Drilling Program: Scientific Results, Volume 159. Texas A&M University, College Station, pp. 3541.Google Scholar
Cloetingh, S. A. P. L., Kooi, H., and Groenewoud, W., 1989. Intraplate stresses and sedimentary basin evolution. In Price, R. (Ed.), Origin and Evolution of Sedimentary Basins and Their Energy and Mineral Resources. American Geophysical Union, Washington, DC, pp. 116.Google Scholar
Cloetingh, S. A. P. L., 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
Cloos, H., 1928. Experimenten zur Inneren Tektonik. Centralbl. F. Mineralogie, 1928(B): 609621.Google Scholar
Cobbold, P. R., Meisling, K. E., and Mount, V. S., 2001. Reactivation of an obliquely rifted margin, Campos and Santos basins, southeastern Brazil. AAPG Bulletin, 85: 19251944.Google Scholar
Cochran, J. R., 1973. Gravity and magnetic investigations in the Guiana Basin, western Equatorial Atlantic. Geological Society of America Bulletin, 84(10): 32493268.2.0.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(1): 171201.CrossRefGoogle Scholar
Cochran, J. R., 1983. Effects of finite rifting times on the development of sedimentary basins. Earth and Planetary Science Letters, 66: 289302.CrossRefGoogle Scholar
Cockburn, H. A. P., Brown, R. W., Summerfield, M. A., and Seidl, M. A., 2000. Quantifying passive margin denudation and landscape development using a combined fission-track thermochronology and cosmogenic isotope approach. Earth and Planetary Science Letters, 179: 429435.CrossRefGoogle Scholar
Colella, A., 1988. Pliocene–Holocene fan deltas and braid deltas in the Crati Basin, southern Italy; a consequence of varying tectonic conditions. In: Nemec, W., and Steel, R. J. (Eds.), Fan Deltas, Blackie, London, pp. 5074.Google Scholar
Colletta, B., Letouzey, J., Pinedo, R., Ballard, J. F., and Balle, P., 1991. Computerized X-ray tomography analysis of sandbox models: Examples of thin-skinned thrust systems. Geology, 9: 10631067.2.3.CO;2>CrossRefGoogle Scholar
Collettini, C., Niemeijer, A., Viti, C., and Marone, C., 2009. Fault zone fabric and fault weakness. Nature, 462: 907910.CrossRefGoogle ScholarPubMed
Colwell, J. B., Graham, T. L., Bradshaw, M., et al., 1990. Stratigraphy of Australia’s NW Continental Margin (Project 121-26). Post-cruise report for BMR Survey 96. Bureau of Mineral Resources, Geology & Geophysics, Division of Marine Geosciences & Petroleum Geology.Google Scholar
Commission for the Geological Map of the World, 1985–1990. International Geological Map of Africa.Google Scholar
Commission for the Geological Map of the World, 2000. Geologic map of South America.Google Scholar
Commission for the Geologic Map of the World, UNESCO-ASGU 1990. Carte Géologique Internationale de l’Afrique, 1:5 000 000, 6 sheets.Google Scholar
Condie, K. C., 2001. Mantle Plumes and Their Record in Earth History. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Connan, J., 1984. Biodegradation of crude oils in reservoirs. In Brooks, J., and Welte, D. (Eds.), Advances in Petroleum Geochemistry. Academic Press, London, pp. 299335.CrossRefGoogle Scholar
Connolly, P., and Cosgrove, J., 1999. Prediction of fracture-induced permeability and fluid flow in the crust using experimental stress data. AAPG Bulletin, 83: 757777.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
Coogan, L. A., Parrish, R. R., and Roberts, N. M. W., 2016. Early hydrothermal carbon uptake by the upper oceanic crust: insight from in situ U–Pb dating. Geology, 44(2): 147150.CrossRefGoogle 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. Available at: www.unr.edu/Geothermal.Google Scholar
Cooper, H. H. Jr., 1966. The equation of groundwater flow in fixed and deforming coordinates. Journal of Geophysical Research, 71: 47854790.CrossRefGoogle Scholar
Corey, A. T., 1986. Mechanics of Immiscible Fluids in Porous Media. Water Resources Publications, Littleton.Google Scholar
Corfu, F., Hanchar, J. M., Hoskin, P. W.O., and Kinny, P., 2003. Atlas of zircon textures. Reviews in Mineralogy and Geochemistry, 53, 469500.CrossRefGoogle Scholar
Cortés, M., and Angelier, J., 2005. Current state of stress in the northern Andes as indicated by focal mechanisms of earthquakes. Tectonophysics, 403: 2958.CrossRefGoogle Scholar
Corti, G., Bonini, M., Innocenti, F., Manetti, P., and Mulugeta, G., 2001. Centrifuge models simulating magma emplacement during oblique rifting. Journal of Geodynamics, 31: 557576.CrossRefGoogle Scholar
Corti, G., Bonini, M., Mazzarini, F., et al., 2002. Magma-indiced strain localization in centrifuge models of transfer zones. Tectonophysics, 348: 205218.CrossRefGoogle Scholar
Corti, G., Bonini, M., Conticelli, S., et al., 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(3–4): 169247.CrossRefGoogle 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, November 2002: 1103–1111.CrossRefGoogle Scholar
Covault, J. A., and Graham, S. A., 2010. Submarine fans at sea-level stands: Tectono-morphologic and climatic controls on terrigenous sediment delivery to the deep sea. Geology, 38: 939942.CrossRefGoogle Scholar
Covault, J. A., Normark, W. R., Romans, B. W., and Graham, S. A., 2007. Highstand fans in the California borderland: the overlooked deep-water depositional systems. Geology, 35: 783786.CrossRefGoogle Scholar
Covault, J. A., Romans, B. W., Fildani, A., McGann, M., and Graham, S. A., 2010. Rapid climatic signal propagation from source to sink in a southern California sediment-routing system. The Journal of Geology, 118: 247259.CrossRefGoogle Scholar
Covault, J. A., Fildani, A., Romans, B. W., and McHargue, T., 2011a. The natural range of submarine canyon-and-channel longitudinal profiles. Geosphere, 7(2): 313332.CrossRefGoogle Scholar
Covault, J. A., Romans, B. W., Graham, S. A., Fildani, A., and Hilley, G. E., 2011b. Terrestrial source to deep-sea sink sediment budgets at high and low sea levels: insights from tectonically active Southern California. Geology, 39: 619622.CrossRefGoogle Scholar
Cowie, P. A., Roberts, G. P., and Mortimer, E., 2007. Strain localization within fault arrays over timescales of 100–107 years: observations, explanations and debates. In: Handy, M., Hirth, G., and Hovius, N (Eds.), Tectonic Faults: Agents of Change on a Dynamic Earth. MIT Press, Cambridge, MA, pp. 4778.CrossRefGoogle Scholar
Cox, K. G., 1980. A model for flood basalt volcanism. Journal of Petrology, 21(4): 629650.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. 703–707.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(9): 10291032.CrossRefGoogle Scholar
Craig, H., and Lupton, J. E., 1981. Helium-3 and mantle volatiles in the ocean and oceanic crust. In: Emiliani, C. (Ed.), The Sea. Wiley, New York, pp. 391428.Google Scholar
Crampin, S., Volti, T., Chastin, S., Gudmundsson, A., and Stefansson, R., 2002. Indication of high pore-fluid pressure in a seismically fault zone. International Journal of Geophysics, 151: F1F5.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, Geological Society/PESGB Conference, May 2000.Google Scholar
Cremonini, O. A., 1993. Caracterização estrutural e evolução tectônica da área de Ubarana, porção submerse da Bacia Potiguar, Brasil. Master’s thesis. Universidade Federal de Ouro Preto, Ouro Preto.Google Scholar
Crowell, J. C., 1974. Origin of late Cenozoic basins in southern California. In: Dickinson, W. (Ed.), Tectonics and Sedimentation. Society of Economic Paleontologists and Mineralogists, Denver, pp. 190204.CrossRefGoogle Scholar
Crowell, J. C. 1975. The San Gabriel fault and Ridge Basin, Southern California. California Division of Mines Geology Special Paper (San Francisco), 118: 208–219.Google Scholar
Crowell, J. C., 1982. The tectonics of the Ridge Basin, southern California. In: Crowell, J. C., and Link, M. H. (Eds.), Geologic History of the Ridge Basin. Pacific Section of the Society of Economic Geologists, Paleontologists and Mineralogists, Los Angeles, pp. 2541.Google Scholar
Crowell, J. C., and Link, M. H., 1982. Geologic History of Ridge Basin, Southern California., Pacific Section, Society of Economic Paleontologists and Mineralogists, Los Angeles.Google Scholar
Csontos, L., Nagymarosy, A., Horváth, F., and Kováč, M., 1992. Tertiary evolution of the intracarpathian area, a model. Tectonophysics, 208: 221241.CrossRefGoogle Scholar
Curray, J. R., 2005. Tectonics and history of the Andaman Sea region. Journal of Asian Earth Sciences, 25: 187228.CrossRefGoogle Scholar
Curray, J. R., Moore, D. G., Lawver, L. A., et al., 1979. Tectonics of the Andaman Sea and Burma. In: Watkins, J., Montadert, L., and Dickerson, P. W. (Eds.), Geological and Geophysical Investigations of Continental Margins. AAPG, Washington, DC, pp. 189198.Google Scholar
Czuba, W., Grad, M., Mjelde, R., et al., 2011. Continent–ocean transition across a trans-tensional margin segment off Bear Island, Barents Sea. International Journal of Geophysics 184: 541554.CrossRefGoogle Scholar
d’Alessio, M. A., Blythe, A.E., and Burgmann, R., 2001. Constraints on frictional heating of faults from fission track thermochronology. American Geophysical Union, Fall Meeting 2001, abstract #S52B-0627.Google Scholar
d’Alessio, M. A., Williams, C. F., and Bürgmann, R., 2006. Frictional strength heterogeneity and surface heat flow: implications for the strength of the creeping San Andreas fault. Journal of Geophysical Research: Solid Earth, 111. DOI: 10.1029/2005JB003780Google Scholar
da Silva Pellegrini, B., and Severiano Ribeiro, H. J. P., 2018. Exploratory plays of Pará-Maranhão and Barreirinhas basins in deep and ultra-deep waters, Brazilian Equatorial Margin. Brazilian Journal of Geology, 48(3). DOI: 10.1590/2317-4889201820180146Google Scholar
Dahlberg, E. C., 1995. Applied Hydrodynamics in Petroleum Exploration. Springer, 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. In: Mohriak, W. and Talwani, M. (Eds.), Atlantic Rift and Continental Margins. American Geophysical Union, Washington, DC, pp. 105128.CrossRefGoogle 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, November 2002: 1140–1146.CrossRefGoogle Scholar
Dailly, P., Henderson, T., Hudgens, E., Kanschat, K., and Lowry, P., 2012. Exploration for Cretaceous stratigraphic traps in the Gulf of Guinea, West Africa and the discovery of the Jubilee Field: a play opening discovery in the Tano Basin, Offshore Ghana. In: Mohriak, W., Danforth, A., Post, P. J., et al. (Eds.), Conjugate Divergent Margins. Geological Society, London, pp. 235248.Google Scholar
Dallmann, W. K., Andresen, A., Bergh, S. G., Maher, H. D., Jr., and Ohta, Y., 1993. Tertiary Fold-and-Thrust Belt of Spitsbergen Svalbard. Norsk Polarinstitutt, Meddelelser.Google Scholar
Daniell, J., Jorgensen, D. C., Anderson, T., et al., 2009. Frontier basins of the west Australian continental margin: post-survey report of marine reconnaissance and geological sampling survey GA2476. Geoscience Australia Survey GA2476, post-survey report.Google Scholar
Darcy, J., 1856. Les Fontaines Publiques de la Ville de Dijon. Dalmont, Paris.Google Scholar
Dateuil, O., and Brun, J-P., 1993. Oblique rifting in a slow spreading ridge. Nature, 361: 145148.CrossRefGoogle Scholar
Dateuil, O., and Brun, J-P., 1996. Deformation partitioning in a slow spreading ridge undergoing oblique extension: Mohns Ridge, Norwegian Sea. Tectonics, 15: 870884.CrossRefGoogle Scholar
Dateuil, O., Huchon, P., Quemeneur, F., and Souriot, T., 2001. Propagation of an oblique spreading centre: the western Gulf of Aden. Tectonophysics, 332: 423442.CrossRefGoogle Scholar
Davey, F. J., Henyey, T., Kleffmann, F., et al., 1995. Crustal reflections from the Alpine Fault Zone, South Island, New Zealand. New Zealand Journal of Geology and Geophysics, 38: 601604.CrossRefGoogle Scholar
Davies, G. F., and Brune, J. N., 1971. Regional and global fault slip rates from seismicity. Nature Physical Sciences, 229: 101107.CrossRefGoogle Scholar
Davies, J. H., and Davies, D. R., 2010. Earth’s surface heat flux, Solid Earth, 1(1): 524.CrossRefGoogle Scholar
Davies, T. A., Luyendyk, B. P., Rodolfo, K. S., et al., 1974. Site 256. In: Initial Reports of the Deep Sea Drilling Project, Volume 26. Texas A&M University, College Station, pp. 295325.Google Scholar
Davis, D. M., and Lillie, R. J., 1994. Changing mechanical response during continental collision: active examples from the foreland thrust belts of Pakistan. Journal of Structural Geology, 16: 2134.CrossRefGoogle Scholar
Davis, E. E., and Becker, K., 2004. Observations of temperature and pressure: constraints on ocean hydrologic state, properties, and flow. In: Davis, E. E., and Elderfield, H. (Eds.), Hydrogeology of Oceanic Lithosphere. Cambridge University Press, Cambridge, pp. 225271.Google Scholar
Davis, E. E., and Villinger, H., 1992. Tectonic and thermal structure of the Middle Valley sedimented rift, northern Juan de Fuca Ridge. In: Proceedings of the Ocean Drilling Program: Scientific Results, Volume 139. Texas A&M University, College Station, pp. 941.Google Scholar
Davis, E. E., Chapman, D. S., Forster, C., and Villinger, H., 1989. Heat-flow variations correlated with buried basement topography on the Juan de Fuca Ridge flank. Nature, 342: 533537.CrossRefGoogle Scholar
Davis, E. E., Chapman, D. S., Mottl, M. J., et al., 1992. FlankFlux: an experiment to study the nature of hydrothermal circulation in young oceanic crust. Canadian Journal of Earth Sciences, 29(5): 925952.CrossRefGoogle Scholar
Davis, E. E., Chapman, D. S., and Forster, C., 1996. Observations concerning the vigor of hydrothermal circulation in young volcanic crust. Journal of Geophysical Research, 101: 29272942.CrossRefGoogle Scholar
Davis, E. E., Fisher, A. T., and Firth, J., 1997. Proceedings of the Oceanic Drilling Program, Initial Reports, Volume 168. Texas A&M University, College Station.Google Scholar
Davis, E. E., Chapman, D. S., Wang, K., et al., 1999. Regional heat flow variations across the sedimented Juan de Fuca Ridge eastern flank: constraints on lithospheric cooling and lateral hydrothermal heat transport. Journal of Geophysical Research, 104: 1767517688.CrossRefGoogle Scholar
Davis, M., and Kusznir, N., 2004. Depth-dependent lithospheric stretching at rifted continental margins. In Karner, G. D., Taylor, B., Driscoll, N. W., and Kohlstedt, D. L. (Eds.), Rheology and Deformation of the Lithosphere at Continental Margins. Columbia University Press, New York, pp. 92137.CrossRefGoogle Scholar
Davison, I., 1999. Tectonics and hydrocarbon distribution along the Brazilian South Atlantic margin. In: Cameron, N. R., Bate, R. H., and Clure, V. S. (Eds.), The Oil and Gas Habitats of the South Atlantic. Geological Society, London, pp. 133151.Google Scholar
Davison, I., 2005. Central Atlantic margin basins of North West Africa: geology and hydrocarbon potential (Morocco to Guinea). Journal of African Earth Sciences, 43: 254274.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
Davison, I., Faull, T., Greenhalgh, J., Beirne, E. O., and Steel, I., 2016. Transpressional structures along the Romanche Fracture Zone. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L. (Eds), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London.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, 188: 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, pp. 537542.Google Scholar
Dayem, K. E., Houseman, G. A., and Molnar, P., 2009. Localization of shear along a lithospheric strength discontinuity: application of a continuous deformation model to the boundary between Tibet and the Tarim Basin. Tectonics, 28(3). DOI: 10.1029/2008TC002264CrossRefGoogle Scholar
de Azevedo, R. P., 1991. Tectonic evolution of Brazilian Equatorial continental margin basins. PhD thesis. Imperial College, London.Google Scholar
De Bremaeker, J. C., 1983. Temperature, subsidence and hydrocarbon maturation in extensional basins: a finite element model. AAPG Bulletin, 67: 14101414.Google Scholar
De Matos, R. M. D., 1987. Sistema de rifts Cretáceos do Nordeste Brasiliero, Anais, Tectos. Petrobrás-Depex, Rio de Janeiro.Google Scholar
de Matos, R. M. D., 1992. The northwest Brazilian rift system. Tectonics, 11(4): 766791.CrossRefGoogle Scholar
de Matos, R. M.D., 2000. Tectonic evolution of the Equatorial South Atlantic. In: Mohriak, W. U., and Talwani, M. (Eds.), Atlantic Rifts and Continental Margins. Wiley, Chichester, pp. 331354.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.Google Scholar
De Mets, C., Gordon, R. G., Argus, D. F., and Stein, S., 1994. Effect of recent revisions to the geomagnetic reversal time scale on current plate motions. Geophysics Research Letters, 21: 21912194.CrossRefGoogle Scholar
De Mets, C., Gordon, R. G., and Argus, D. F., 2010. Geologically current plate motions. Geophysical Journal, 181(1): 180.CrossRefGoogle Scholar
De Wit, M. J., Brito Neves, B. B., Trouw, R. A. J., and Pankhurst, R. J., 2018. Pre-Cenozoic correlations across the South Atlantic region: “the ties that bind.” In: Pankhurst, R. J., Trouw, R. A. J., Brito Neves, B. B., and De Wit, M. J. (Eds), Pre-Cenozoic Correlations Across the South Atlantic Region. Geological Society of London, London, pp. 18.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: 58595885.CrossRefGoogle Scholar
Dean, W. E., 1974. Determination of carbonate and organic matter in calcareous sediments and sedimentary rocks by loss on ignition: comparison with other methods. Journal of Sedimentary Petrology, 44: 242248.Google Scholar
Dearman, W. R., 1963. Wrench-faulting in Cornwall and South Devon. Proceedings of the Geologists’ Association, London, 74: 265287.CrossRefGoogle Scholar
Decker, K., 1996. Miocene tectonics at the Alpine–Carpathian junction and the evolution of the Vienna Basin. Mitteilungen der Gesellschaft für Geologie und Bergbaustudenten Österreichs, 41: 3344.Google Scholar
De-Hua, H., and Batzle, M., 2006. Velocities of deepwater reservoir sands. The Leading Edge, April 2006: 260–266.Google Scholar
Delacour, A., Fruh-Green, G. L., Bernasconi, S. M., Schaeffer, P., and Kelley, D. S., 2008. Carbon geochemistry of serpentinites in the Lost City Hydrothermal Systems (30° N, MAR). Geochimica et Cosmochimica Acta, 72: 36813702.CrossRefGoogle Scholar
Delescluse, M., Chamot-Rooke, N., Cattin, R., et al., 2012. April 2012 intra-oceanic seismicity off Sumatra boosted by the Banda-Aceh megathrust. Nature, 490: 240244.CrossRefGoogle ScholarPubMed
Demaison, G., and Huizinga, B. J., 1991. Genetic classification of petroleum systems. AAPG Bulletin, 75: 16261643.Google Scholar
Demaison, G. J., and Moore, G.T, 1980. Anoxic environments and oil source bed genesis. AAPG Bulletin, 64: 11791209.Google Scholar
Dembicki, H. Jr., and Anderson, M. L., 1989. Secondary migration of oil experiments supporting efficient movement of separate, buoyant oil phase along limited conduits. AAPG Bulletin, 73: 10181021.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
Deming, D., 1994a. Fluid flow and heat transport in the upper continental crust. In: Parnell, J. (ed.), Geofluids; Origin, Migration and Evolution of Fluids in Sedimentary Basins. Geological Society of London, London, pp. 2742.Google Scholar
Deming, D., 1994b. Overburden rock, temperature, and heat flow. In: Magoon, L. B., and Dow, W. G. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 165186.Google Scholar
Deming, D., 1994c. Factors necessary to define a pressure seal. AAPG Bulletin, 78: 10051009.Google Scholar
Deming, D., and Nunn, J. A., 1991. Numerical simulations of brine migration by topographically driven recharge. Journal of Geophysical Research, 96: 24852499.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: 66696683.CrossRefGoogle Scholar
Deming, D., Nunn, J. A., Jones, S., and Chapman, D. S., 1990b. Some problems in thermal history studies. In: Nuccio, F., and Barker, C. E. (Eds.), Applications of Thermal Maturity Studies to Energy Exploration. Society of Economic Paleontologists and Mineralogists, Los Angeles, 6180.Google Scholar
Deng, Q., Wu, D., Zhang, P., and Chen, S., 1986. Structure and deformational character of strike-slip fault zones. Pure and Applied Geophysics, 124 (1–2): 203223.CrossRefGoogle 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: 1520315219.CrossRefGoogle Scholar
Deptuck, M. E., 2011. Proximal to distal postrift structural provinces of the western Scotian Margin, offshore Eastern Canada: geological context and parcel prospectivity for call for bids NS11–11. CNSOPB Geoscience Open File Report, 2011-001MF.Google Scholar
Desa, M., Ramana, M. V., and Ramprasad, T., 2006. Seafloor spreading magnetic anomalies south off Sri Lanka. Marine Geology, 229: 227240.CrossRefGoogle Scholar
DESERT Group, Weber, M., Abu-Ayyash, K., et al., 2004. The crustal structure of the Dead Sea Transform. Geophysical Journal International, 156 (3): 655681.Google Scholar
Destro, N., Szatmari, P., Alkmim, F. F., and Magnavita, L. P., 2003. Release faults, associated structures, and their control on petroleum trends in the Recõncavo rift, northeast Brazil. AAPG Bulletin, 87(7): 11231144.CrossRefGoogle Scholar
Detrick, R. S., White, R. S., and Purdy, G. M., 1993. Crustal structure of North Atlantic fracture zones. Reviews of Geophysics, 31 (4): 439458.CrossRefGoogle Scholar
Detrick, R. S., Needham, H. D., and Renard, V., 1995. Gravity-anomalies and crustal thickness variations along the Mid-Atlantic Ridge between 33°N and 40°N. Journal of Geophysical Research, 100: 37673787.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
Dewey, J. F., 1982. Plate tectonics and evolution of British Isles. Journal of the Geological Society of London, 139: 371412.CrossRefGoogle Scholar
Dewey, J. F., and Burke, K., 1974. Hot spots and continental break-up: implications for collisional orogeny. Geology, 2: 5760.2.0.CO;2>CrossRefGoogle Scholar
Dewey, J. F., and Şengör, A. M. C., 1979. Aegean and surrounding regions complex multiplate and continuum tectonics in a convergent zone. Geological Society of America Bulletin, 90: 8492.2.0.CO;2>CrossRefGoogle Scholar
Dewey, J. F., Hempton, M. R., Kidd, W. S. F., Şaroğlu, F., and Şengör, A. M.C., 1986. Shortening of continental lithosphere: the neotectonics of eastern Anatolia, a young collision zone. In: Coward, M.P., and Ries, A.C. (Eds.), Collision Tectonics. Geological Society of London, London, p. 336Google 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: 253267.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: 761768.CrossRefGoogle 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. In: Carneiro, C.d.R. (Ed.), Revista Brasileira de Geociencias, pp. 162–167.CrossRefGoogle Scholar
Dibblee, T. R., 1980. Geology Along the San Andreas Fault from Gilroy to Parkfield. California Division of Mines and Geology, Sacramento.Google Scholar
Dick, H. J., Lin, J., and Schouten, H., 2003. An ultraslow spreading class of ocean ridge. Nature, 426: 405412.CrossRefGoogle ScholarPubMed
Dickinson, G., 1953. Geological aspects of abnormal reservoir pressures in Gulf Coast, Louisiana. AAPG Bulletin, 37: 410432.Google Scholar
Diegel, F. A., Karlo, J. F., Schuster, D. C., Shoup, R.C., and Tauvers, P. R., 1995. Cenozoic structural evolution and tectonostratigraphic framework of the northern Gulf Coast continental margin. In: Jackson, M. P. A., Roberts, D. G., and Snelson, S. (Eds.), Salt Tectonics: a Global Perspective. AAPG, Washington, DC, pp. 109151.Google Scholar
Diehl, T., Waldhauser, F., Cochran, J. R., et al., 2013. Back-arc extension in the Andaman Sea: tectonic and magmatic processes imaged by high-precision teleseismic double-difference earthquake relocation. Journal of Geophysical Research, 118: 22062224.CrossRefGoogle Scholar
Dodson, M. H., 1973. Closure temperature in cooling geochronological and petrological systems. Contributions to Mineralogy and Petrology, 40(3): 259274.CrossRefGoogle 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: Burrus, J. (Ed.), Thermal Modeling in Sedimentary Basins. Technip, Paris, pp. 173195.Google Scholar
Donelick, R. A., 1994. Selected Poland samples: apatite fission track data. Donelick Analytical Report 76: 1114.Google Scholar
Dooley, T., and McClay, K., 1997. Analog modeling of pull-apart basins. AAPG Bulletin, 81(11): 18041826.Google Scholar
Doré, A. G., Lundin, E. R., Jensen, L. N., et al., 1999. Principal tectonic events in the evolution of the northwest European Atlantic margin. In: Fleet, A. J., and Boldy, S. A. R. (Eds.), Petroleum Geology of Northwest Europe: Proceedings of the 5th Conference. Geological Society, London, pp. 4161.Google Scholar
Doré, A. G., Lundin, E. R., Gibbons, A., Sømme, T., and Tørudbakken, B. O., 2016. Transform margins of the Arctic: a synthesis and re-evaluation. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L. (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society, London, pp. 6394.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: 177199.2.0.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
Doust, H., and Omatsola, E., 1990. Niger Delta. In: Edwards, J. D., and Santogrossi, P. A. (Eds.), Divergent/Passive Margin Basins, AAPG, Washington, DC, pp. 201238.Google Scholar
Dow, W. G., 1977. Kerogen studies and geological interpretations. Journal of Geochemical Exploration, 7(2): 7779.CrossRefGoogle Scholar
Downey, M. W., 1994. Hydrocarbon seal rocks. In: Magoon, L. B. and Dow, W. G. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 159164.Google 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, 40–41: 101116.CrossRefGoogle Scholar
Du Rouchet, J., 1981. Stress fields: a key to oil migration. AAPG 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: 12091216.CrossRefGoogle Scholar
Dubessy, J., Poty, B., and Ramboz, C., 1989. Advances in C-O-H-N-S fluid geochemistry based on micro-Raman spectrometric analysis of fluid inclusions. European Journal of Mineralogy, 1: 517534.CrossRefGoogle Scholar
Ducassou, E., Migeon, S., Mulder, T., et al., 2009. Evolution of the Nile deep-sea turbidite system during the Late Quaternary: influence of climate change on fan sedimentation: Sedimentology, 56: 20612090.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
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: Parnell, J. (Ed.), Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins. Geological Society of London, London, pp. 325345.Google Scholar
Dula, W. F., 1981. Correlation between deformation lamellae, microfractures, macrofractures, and in situ stress measurements, White River Uplift, Colorado. Geological Society of America Bulletin, Part I, 92: 3746.2.0.CO;2>CrossRefGoogle Scholar
Dullo, W. C., 1983. Diagenesis of fossils of the Miocene Leitha Limestone of the Paratethys, Austria: an example for faunal modifications due to changing diagenetic environments. Facies, 8: 1112.CrossRefGoogle Scholar
Dumitru, T. A., Hill, K., Coyle, D., et al, 1991. Fission track thermochronology: application to continental rifting of south-eastern Australia. Australian Petroleum Explorers Association Journal, 10: 131142.Google Scholar
Dumont, J. F., 1987. Etude structurale des bordures nord et sud du plateau de l’Adamaoua: influence du contexte atlantique. Geodynamique, 2(1): 5568.Google Scholar
Durand, B., 1980. Sedimentary organic matter and kerogen: definition and quantitative importance of kerogen. In: Durand, B. (Ed.), Kerogen: Insoluble Organic Matter from Sedimentary Rocks. Technip, Paris, pp. 1334.Google Scholar
Durand, B., 1988. Understanding of HC migration in sedimentary basins (present state of knowledge). Organic Geochemistry, 13: 445459.CrossRefGoogle Scholar
Ďurica, D., Namestnikov, J. G., Pagáč, I., and Roth, Z., 1986. Oil and Gas Fields in the Central Europe [in Czech]. Alfa, Bratislava.Google Scholar
Durrheim, R. J., and Mooney, W. D., 1991. Archean and Proterozoic crustal evolution: evidence from crustal seismology. Geology, 19(6): 606609.2.3.CO;2>CrossRefGoogle Scholar
Durrheim, R. J., and Mooney, W. D., 1992. Archean and Proterozoic crustal evolution: evidence from crustal seismology: Reply [modified]. Geology (Boulder), 20(7): 665666.Google Scholar
Duval, B., Cramez, C., and Jackson, M. P. A., 1992, Raft tectonics in the Kwanza Basin, Angola: Marine and Petroleum Geology, 9: 389404.CrossRefGoogle Scholar
Dziak, R. P., Fox, C. G., Embley, R. W., et al., 1996. Detection of and response to a probable volcanogenic T-wave event swarm on the western Blanco Transform Fault Zone. Geophysical Research Letters 23(8): 873876.CrossRefGoogle Scholar
Dziak, R. P., Chadwick, W. W., Jr., Fox, C. G., and Embley, R. W., 2003. Hydrothermal temperature changes at the Southern Juan de Fuca Ridge associated with Mw 6.2 Blanco transform earthquakes. Geology, 31: 119122.2.0.CO;2>CrossRefGoogle Scholar
Ebinger, C. J., 1989a. Tectonic development of the western branch of the East African Rift System. Geological Society of America Bulletin, 101: 885903.2.3.CO;2>CrossRefGoogle Scholar
Ebinger, C. J., 1989b. Geometric and kinematic development of border faults and accommodation zones, Kivu-Rusizi rift, Africa. Tectonics, 8: 117133.CrossRefGoogle Scholar
Ebinger, C. J., and Casey, M., 2001. Continental breakup in magmatic provinces; an Ethiopian example. Geology (Boulder), 29(6): 527530.2.0.CO;2>CrossRefGoogle Scholar
Ebinger, C. J., Deino, A. L., Drake, R. E., and Tesha, A. L., 1989. Chronology of volcanism and rift basin propagation: Rungwe Volcanic Province, East Africa. Journal of Geophysical Research, 94: 1578515803.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, 1992.CrossRefGoogle 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(6): 371377.CrossRefGoogle Scholar
Eckstein, Y., and Simmons, G., 1978. Measurements and interpretation of terrestrial heat flow in Israel. Geothermics, 6: 117142.CrossRefGoogle Scholar
Edgar, N. T., Saunders, J. B., Bolli, H. M., et al., 1973. Site 153. In: Initial Reports of the Deep Sea Drilling Project 15. US Government Printing Office, Washington, DC, pp. 367407.CrossRefGoogle Scholar
Edwards, M. H., Kurras, G. J., Tolstoy, M., et al., 2001, Evidence of recent volcanic activity on the ultraslow-spreading Gakkel Ridge: Nature, 409: 808812.CrossRefGoogle ScholarPubMed
Edwards, R. A., Whitmarsh, R. B., and Scrutton, R. A., 1997a. The crustal structure across the transform continental margin off Ghana, eastern equatorial Atlantic. Journal of Geophysical Research, 102: 747772.CrossRefGoogle Scholar
Edwards, R. A., Whitmarsh, R. B., and Scrutton, R. A., 1997b. Synthesis of the crustal structure of the transform continental margin off Ghana, northern Gulf of Guinea. Geo-Marine Letters, 17: 1220.CrossRefGoogle Scholar
Egbobawaye, E. I., 2017. Petroleum source rock evaluation and hydrocarbon potential in Montney Formation unconventional reservoir, Northeastern British Columbia, Canada. Natural Resources, 8: 716756.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: 217234.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(1–2): 114.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: 179194.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 (B3): 2173.CrossRefGoogle Scholar
Eichhubl, P., and Boles, J. R., 2000. Focused fluid flow along faults in the Monterey Formation, coastal California. GSA Bulletin, 112: 16671679.2.0.CO;2>CrossRefGoogle Scholar
Einarsson, P., 1991. Earthquakes and present-day tectonism in Iceland. Tectonophysics, 189 (1–4): 261279.CrossRefGoogle Scholar
Einarsson, P., and Brandsdóttir, B., 1980. Seismological evidence for lateral magma intrusion during the July 1978 deflation of the Krafla volcano in NE-Iceland. Journal of Geophysics, 47: 160165.Google Scholar
Einarsson, P., and Eiriksson, J., 1982. Earthquake fractures in the districts Land and Rangarvellir in the South Iceland Seismic Zone. Jökull, 32: 113120.CrossRefGoogle Scholar
Einsele, G., 1992. Sedimentary Basins: Evolution, Facies and Sediment Budget. Springer, Berlin.CrossRefGoogle Scholar
Elders, W. A., 1979, The geological background of the geothermal fields of the Salton Trough. In Elders, W. A., (Ed.), Geology and Geothermics of the Salton Trough: University of California, Riverside. Campus Museum Contributions, Riverside, pp. 119.Google Scholar
Elders, W. A., and Sass, J. H., 1988. The Salton Sea Scientific Drilling Project. Journal of Geophysical Research, B, Solid Earth and Planets, 93(11): 1295312968.CrossRefGoogle Scholar
Elders, W. A., Rex, R. W., Meidav, T., Robinson, P. T., and Biehler, S., 1972. Crustal spreading in Southern California. Science, 178(4056): 1524.CrossRefGoogle ScholarPubMed
Elliott, D., 1976. The energy balance and deformation mechanisms of thrust sheets. Philosophical Transactions of the Royal Society of London, A283: 183192.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(10): 17111741.CrossRefGoogle Scholar
Ellis, S., and Stockhert, B., 2004. Imposed strain localization in the lower crust on seismic timescales. Earth, Planets and Space, 56: 11031109.CrossRefGoogle Scholar
Emmanuel, S., and Ague, J. J., 2007. Implications of present-day abiogenic methane fluxes for the early Archean atmosphere. Geophysics Research Letters, 34: L15810.CrossRefGoogle Scholar
Emmons, R. C., 1969. Strike-slip rupture patterns in sand models. Tectonophysics, 7: 7187.CrossRefGoogle 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. Stud. Cercet. Geol. Geofiz. Geogr. Ser. Geofiz., 30: 319.Google Scholar
Engelder, T., 1993. Stress Regimes in the Lithosphere. Princeton University Press, Princeton.Google Scholar
Engeln, J. F., Wiens, D. A., and Stein, S., 1986. Mechanisms and depths of Atlantic transform earthquakes. Journal of Geophysical Research, 91: 548577.CrossRefGoogle Scholar
England, P. C., and Houseman, G. A., 1988. The mechanics of the Tibetan Plateau, Philosophical Transactions of the Royal Society, London, Series A, 326: 301320.Google Scholar
England, P., and Molnar, P., 1990. Surface uplift, uplift of rocks, and exhumation of rocks. Geology, 18: 11731177.2.3.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: 201213.CrossRefGoogle Scholar
England, W. A., 1994. Secondary migration and accumulation of hydrocarbons. In: Magoon, L. B. and Dow, W. G. (Ed.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 211217.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: 327347.CrossRefGoogle Scholar
ENI, 2015. Ghana offshore Cape Three Points oil block development Phase 2. Final environmental impact statement.Google Scholar
Ensley, R. A., and Verosub, K. L., 1982. A magnetostratigraphic study of the sediments of the Ridge Basin, southern California and its tectonic and sedimentologic implications. Earth and Planetary Science Letters, 59(1): 192207.CrossRefGoogle Scholar
Eppelbaum, L., Kutasov, I., and Pilchin, A., 2014. Thermal properties of rocks and density of fluids. In: Applied Geothermics: Lecture Notes in Earth System Sciences. Springer, New York, pp. 99149.CrossRefGoogle Scholar
Erbacher, J., and Thurow, J., 1997. Influence of oceanic anoxic events on the evolution of Mid-Cretaceous Radiolaria in the North Atlantic and western Tethys. Marine Micropaleontology, 30(103): 139158.CrossRefGoogle Scholar
Ercan, T., Turkecan, A., Guillou, H., et al., 1998. Features of the Tertiary volcanism around Sea of Marmara. Mineral Research and Exploration Bulletin, 120: 97118.Google Scholar
Escalona, A., and Mann, P., 2006. Tectonic controls of the right-lateral Burro Negro tear fault on Paleogene structure and stratigraphy, northeastern Maracaibo Basin. AAPG Bulletin, 90(4): 479504.CrossRefGoogle Scholar
Escartín, J., Hirth, G., and Evans, B., 1997a. Effects of serpentinization on the lithospheric strength and the style of normal faulting at slow-spreading ridges. Earth and Planetary Science Letters, 151: 181189.CrossRefGoogle Scholar
Escartín, J., Hirth, G., and Evans, B., 1997b. Nondilatant brittle deformation of serpentinites: implications for Mohr–Coulomb theory and the strength of faults. Journal of Geophysical Research, 102(B2): 28972913.CrossRefGoogle Scholar
Escartín, J., Hirth, G., and Evans, B., 2001. Strength of slightly serpentinized peridotites: implications for the tectonics of oceanic lithosphere. Geology, 29: 10231026.2.0.CO;2>CrossRefGoogle Scholar
Escartín, J., Mével, 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°45′N. Geochemistry, Geophysics, Geosystems, 4. DOI: 10.1029/2002GC000472CrossRefGoogle Scholar
Espitalié, J., Laporte, J. L., Madec, M., et al., 1977a. 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(1): 2342.Google Scholar
Espitalié, J., Madec, M., Tissot, B., Menning, J. J., and Leplat, P., 1977b. Source rock characterization method for petroleum exploration. Ninth Annual Offshore Technology Conference, pp. 439–448.CrossRefGoogle Scholar
Etchecopar, A., 1977. A plane kinematic model of progressive deformation in a polycrystalline aggregate. Tectonophysics, 39(1): 121139.CrossRefGoogle 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: 37213735.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(2): 129152.CrossRefGoogle Scholar
Exon, N. F., and Ramsay, D. C., 1990. BMR CRUISE 95 Triassic and Jurassic sequences of the northern Exmouth Plateau and offshore Canning Basin. Record. Bureau of Mineral Resources, Geology and Geophysics, Geoscience Australia, Canberra.Google Scholar
Exon, N. F., Williamson, P. E., Boyd, R., et al., 1988: BMR Record 1988/30, Rig Seismic Research Cruise 7&8: Sedimentary basin framework of the Northern and Western Exmouth Plateau. Department of Primary Industries and Energy, Bureau of Mineral Resources, Geology and Geophysics, Geoscience Australia, Canberra.Google Scholar
Exon, N. F., Berry, R. F., Crawford, A. J., and Hill, P. J., 1997. Geological evolution of the East Tasman Plateau, a continental fragment southeast of Tasmania. Australian Journal of Earth Sciences, 44: 597608.CrossRefGoogle Scholar
Eyal, M., Eyal, Y., Bartov, Y., and Steinitz, G., 1981. The tectonic development of the Western margin of the Gulf of Elat (Aqaba) Rift. Tectonophysics, 80: 3966.CrossRefGoogle Scholar
Fachmann, S., 2001. Geologische Entwicklung im Umfeld des Mahanadi-Riftes (Indien), Universitat Freiberg.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(3): 198211.Google Scholar
Faleide, J. I., Bjørlykke, K., and Gabrielsen, R. H., 2010. Geology of the Norwegian continental shelf. In: Bjørlykke, K. (Ed.), Petroleum Geoscience: From Sedimentary Environments to Rock Physics. Springer, Berlin.Google Scholar
Fall, A., Tattitch, B., and Bodnar, R., 2011. Combined microthermometric and Raman spectroscopic technique to determine the salinity of H2O–CO2–NaCl fluid inclusions based on clathrate melting. Geochimica et Cosmochimica Acta, 75: 951964.CrossRefGoogle Scholar
Farley, K. A., 2000. Helium diffusion from apatite: general behavior as illustrated by Durango fluorapatite. Journal of Geophysical Research, 105(B2): 2000.CrossRefGoogle Scholar
Farley, K. A., 2002. (U–Th)/He dating: techniques, calibrations, and applications. Reviews in Mineralogy and Geochemistry, 47: 819843.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(21): 42234229.CrossRefGoogle Scholar
Faugére, E., and Brun, J. P., 1984. Modelisation experimentale de la distention continentale. Experimental models of a stretched continental crust. Comptes-Rendus des Seances de l’Academie des Sciences, Serie 2: Mecanique-Physique, Chimie, Sciences de l’Univers, Sciences de la Terre, 299(7): 365370.Google Scholar
Faugére, E., Brun, J. P., van den Driessche, J., et al., 1986. Bassins asymetriques en extension pure et en decrochement; modeles experimentaux. Bulletin des Centres de Recherches Exploration-Production Elf-Aquitaine, 10: 1321.Google Scholar
Faulds, J. E., and Varga, R. J., 1998. The role of accommodation zones and transfer zones in the regional segmentation of extended terranes. Geological Society of America Special Paper 323.CrossRefGoogle Scholar
Fechtig, H., and Kalbitzer, S., 1966. The diffusion of argon in potassium-bearing solids. Federal Republic of Germany (DEU).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: 41214142.CrossRefGoogle Scholar
Fernàndez, M., Ayala, C., Torne, M., et al. 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
Feybesse, J. L., Johan, V., Triboulet, C., et al., 1998. The West Central African belt, a model of 2.5–2.0 Ga accretion and two-phase orogenic evolution. Precambrian Research, 87: 161216.CrossRefGoogle 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) : 409417.CrossRefGoogle Scholar
Field, E. H., Dawson, T. E., Felzer, K. R., et al., 2009. Uniform California earthquake forecast, Version 2 (UCERF 2). Bulletin of the Seismological Society of America, 99: 20532107.CrossRefGoogle Scholar
Figueiredo, A. M.F., Braga, J. A. E., Zabalaga, J. C., et al., 1994. Recôncavo Basin, Brazil: a prolific intracontinental rift basin. In: Landon, S. M. (Ed.), Interior Rift Basins. AAPG, Washington, DC, pp. 157203.Google Scholar
Finlayson, D. M., Collins, C. D., 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: 177189.CrossRefGoogle Scholar
Finzi, Y., Hearn, E. H., Ben-Zion, Y., and Lyakhovsky, V., 2009. Structural properties and deformation patterns of evolving strike-slip faults: numerical simulations incorporating damage rheology. Pure and Applied Geophysics, 166(10–11): 15371573.CrossRefGoogle Scholar
Fisher, A. T., 1998. Permeability within basaltic oceanic crust. Reviews in Geophysics, 36(2): 143182.CrossRefGoogle Scholar
Fisher, A. T., 2004. Rates and patterns of fluid circulation. In: Davis, E. E., and Elderfield, H. (Eds.), Hydrogeology of Oceanic Lithosphere. Cambridge University Press, Cambridge, pp. 339377.Google 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 (B7): 1264112657.CrossRefGoogle Scholar
Fisher, A. T., and Becker, K., 2000. Channelized fluid flow in oceanic crust reconciles heat-flow and permeability data. Nature, 403: 7174.CrossRefGoogle ScholarPubMed
Fisher, A. T., Becker, K., Narasimhan, T. N., Langseth, M. G., and Mottl, M. J., 1990. Passive, off-axis convection on the southern flank of the Costa Rica Rift. Journal of Geophysical Research, 95: 93439370.CrossRefGoogle Scholar
Fisher, A. T., Zwart, G., Shipley, T., et al., 1996. Relation between permeability and effective stress along a plate-boundary fault, Barbados accretionary complex. Geology, 24: 307310.2.3.CO;2>CrossRefGoogle Scholar
Fisher, A. T., Davis, E. E., Hutnak, M., et al., 2003a. Hydrothermal recharge and discharge across 50 km guided by seamounts on a young ridge flank. Nature, 421: 618621.CrossRefGoogle Scholar
Fisher, A. T., Stein, C. A., Harris, R. N., et al., 2003b. Abrupt thermal transition reveals hydrothermal boundary and role of seamounts within the Cocos Plate. Geophysics Research Letters, 30(11): 1550.CrossRefGoogle Scholar
Fisher, Q. J., and Knipe, R. J., 1998. Fault sealing processes in siliciclastic rocks. In: Jones, G., Fisher, Q. J., and Knipe, R. J. (Eds.), Faulting, Fault Sealing and Fluid Flow in Hydrocarbon Reservoirs. Geological Society of London, London, pp. 117134.Google Scholar
Fitch, T. J., 1972. Plate convergence, transcurrent faults, and internal deformation adjacent to southeast Asia and the western Pacific. Journal of Geophysical Research, 84: 44324460.CrossRefGoogle Scholar
Flemings, P. B., and Jordan, T. E., 1990. Stratigraphic modeling of foreland basins: interpreting thrust deformation and lithosphere rheology. Geology, 18: 430434.2.3.CO;2>CrossRefGoogle Scholar
Fliervoert, T. F., White, S. H., and Drury, M. R., 1997. Evidence for dominant grain-boundary sliding deformation in greenschist- and amphibolite-grade polymineralic ultramylonites from the Redbank Deformed Zone, Central Australia. Journal of Structural Geology, 19(12): 14951520.CrossRefGoogle Scholar
Flores, M. R., 2014. Coal and Coalbed Gas, Fueling the Future. Elsevier Science, Waltham, MA.Google 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: 341345.Google Scholar
Flovenz, O. G., and Saemundsson, K., 1993. Heat flow and geothermal processes in Iceland. Tectonophysics, 225: 123138.CrossRefGoogle Scholar
Fodor, L., 1995. From transpression to transtension: Oligocene–Miocene structural evolution of the Vienna Basin and the East Alpine–Western Carpathian junction. Tectonophysics, 242(1–2): 151182.CrossRefGoogle Scholar
Fodor, L., Marko, F., and Nemčok, M., 1990. Microtectonic evolution and paleo-stress field in the Vienna Basin (in French). Geodinamica Acta, 4(3): 147158.CrossRefGoogle Scholar
Forman, D. J., and Wales, D. W., 1981. Geological evolution of the Canning Basin, Western Australia. Bulletin 210 from the Bureau of Mineral Resources, Geology and Geophysics, Australia.Google Scholar
Fornari, D. J., Gallo, D. G., Edwards, M. H., et al., 1989. Structure and topography of the Siqueros transform-fault system: evidence for the development of intra-transform spreading centers. Marine Geophysics Research, 11: 263299.CrossRefGoogle Scholar
Forster, A., Schouten, S., Moriya, K., Wilson, P. A., and Sinninghe Damsté, J. S., 2007. Tropical warming and intermittent cooling during the Cenomanian/Turonian oceanic anoxic event 2: sea surface temperature records from the equatorial Atlantic. Paleooceanography 22: PA1219.CrossRefGoogle Scholar
Foster, J., 2010. The construction and development of SHRIMP I: an historical outline. Precambrian Research, 183: 18.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: 2549925509.CrossRefGoogle Scholar
Foustoukos, D. L., and Seyfried, W. E., 2004. Hydrocarbons in hydrothermal vent fluids: the role of chromium-bearing catalysts. Science, 304: 10021005.CrossRefGoogle ScholarPubMed
Fox, P. J., and Gallo, D. G., 1984. A tectonic model for ridge-transform–ridge plate boundaries: implications for the structure of oceanic lithosphere. Tectonophysics, 104: 205242.CrossRefGoogle Scholar
Francis, T. J., 1981. Serpentinization faults and their role in the tectonics of slow spreading ridges. Journal of Geophysical Research, 86(B12): 1161611622.CrossRefGoogle Scholar
Franců, J., Radke, M., Schaefer, R. G., et al., 1996. Oil–oil and oil–source correlations in the northern Vienna Basin and adjacent Carpathian Flysch Zone (Czech and Slovak area). In: Wessely, G., and Liebl, W. (Eds.), Oil and Gas in Alpidic Thrustbelts and Basins of Central and Eastern Europe. EAGE, Bunnik, pp. 343353.Google Scholar
Frape, S. K., and Fritz, P., 1987. Geochemical trends for groundwaters from the Canadian Shield. In: Fritz, P., and Frape, S. K. (Eds.), Saline Water and Gases in Crystalline Rocks. Geological Association of Canada, Newfoundland, pp. 1938.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
Freund, R., and Merzer, A. M., 1976. Anisotropic origin of transform faults. Science, 192: 137138.CrossRefGoogle ScholarPubMed
Frohlich, C., and Apperson, K. D., 1992. Earthquake focal mechanisms, moment tensors, and the consistency of seismic activity near plate boundaries. Tectonics, 11: 279296.CrossRefGoogle Scholar
Frost, B. R., and Bucher, K., 1994. Is water responsible for geophysical anomalies in the deep continental crust? A petrological perspective. Tectonophysics, 231(4): 293309.CrossRefGoogle Scholar
Frost, B., and Nemčok, M., 2017. Crustal architecture of the proto-Caribbean oceanic crust. William Smith Meeting, Plate tectonics at 50, October 3–5, 2017, Burlington House London, Abstracts, Geological Society of London.Google Scholar
Früh-Green, G. L., Kelley, D. S., Bernasconi, S. M., et al., 2003. 30,000 years of hydrothermal activity at the Lost City Vent Field. Science, 301: 495498.CrossRefGoogle ScholarPubMed
Früh-Green, G. L., Connolly, J. A. D., Kelley, D. S., Plas, A., and Grobery, B., 2004. Serpentinization of oceanic peridotites: implications for geochemical cycles and biological activity. In: Wilcock, W. D., Kelley, D. S., DeLong, E., and Cary, C. (Eds.), The Subseafloor Biosphere at Mid-Ocean Ridges. American Geophysical Union, Washington, DC, pp. 119136.CrossRefGoogle Scholar
Fuchs, W., 1965. Geology of Ruster Hills (Burgenland) [in German]. Jahrbuch der Geologischen Bundesanstalt, 108: 155194.Google Scholar
Fuis, G. S., and Clowes, R. M., 1993. Comparison of deep structure along three transects of the western North American Continental Margin. Tectonics, 12(6): 14201435.CrossRefGoogle Scholar
Fuis, G. S., Ryberg, T., Godfrey, N. J., Okaya, D. A., and Murphy, J. M., 2001. Crustal structure and tectonics from the Los Angeles basin to the Mojave Desert, southern California. Geology, 29(1): 1518.2.0.CO;2>CrossRefGoogle Scholar
Fuis, G. S., Kohler, M. D., Scherwath, M., et al., 2007. A comparison between transpressional plate boundaries of South Island, New Zealand, and Southern California, USA: the Alpine and San Andreas Fault systems. In: Okaya, D., Stern, T., and Davey, F. (Eds.), A Continental Plate Boundary: Tectonics at South Island, New Zealand. American Geophysical Union, Washington, DC, pp. 309330.Google Scholar
Fuis, G. S., Scheirer, D. S., Langenheim, V. E., and Kohler, M. D., 2012. A new perspective on the geometry of the San Andreas Fault in southern California and its relationship to lithospheric structure. Bulletin of the Seismological Society of America, 102: 236251.CrossRefGoogle Scholar
Fullerton, L. G., Sager, W. W., and Handschumacher, D. W., 1989. Late Jurassic – Early Cretaceous evolution of the eastern Indian Ocean adjacent to northwest Australia. Journal of Geophysical Research, 94: 29372953.CrossRefGoogle Scholar
Fulton, P. M., and Saffer, D. M., 2009. Potential role of mantle-derived fluids in weakening the San Andreas Fault. Journal of Geophysical Research: Solid Earth, 114( B7). DOI: 10.1029/2008JB006087Google Scholar
Funck, T., Hopper, J. R., Larsen, H. C., et al., 2003. Crustal structure of the ocean–continent transition at Flemish Cap: seismic refraction results. Journal of Geophysical Research, 108: 20.CrossRefGoogle Scholar
Furlong, K. P., and Chapman, D. S., 2013. Heat flow, heat generation, and the thermal state of the lithosphere. Annual Review of Earth and Planetary Sciences, 41: 385410.CrossRefGoogle Scholar
Fusion Oil and Gas, 2002. Deepwater Northwest Africa: a geological tour. Presentation in Cape Town.Google Scholar
Fyock, T., 1968. Penetration chart: west side oil fields. In: Guidebook, Geology and Oilfields, West Side Southern San Joaquin Valley, 43rd Annual Meeting of Pacific Sections of AAPG, SEG, SEPM, Tulsa, pp. 54–56.Google Scholar
Gabrielsen, R. H., and Færseth, R. B., 1989. The inner shelf of North Cape, Norway and its implications for the Barents Shelf–Finnmark Caledonide boundary. A comment. Norsk Geologisk Tidsskrift, 69: 5762.Google Scholar
Gabrielsen, R. H., Færseth, R. B., Jensen, L. N., Kalheim, J. E., and Riis, F., 1990. Structural elements of the Norwegian continental shelf, Part 1: the Barents Sea region. Norwegian Petroleum Directorate Bulletin, 6: 47.Google Scholar
Gadd, S. A., and Scrutton, R. A., 1997. An integrated thermomechanical model for transform continental margin evolution. Geo-Marine Letters, 17: 2130.CrossRefGoogle Scholar
Gaina, C., Müller, R. D., Brown, B., and Ishihara, T., 2007. Breakup and early sea floor spreading between India and Antarctica. Geophysical Journal International, 170: 151169.CrossRefGoogle 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 86-631.CrossRefGoogle Scholar
Galbraith, R.F., 1992. Statistical models for mixed ages. International Workshop on Fission Track Thermochronology, University if Pennsylvania.Google Scholar
Gallagher, K., 1995. Evolving thermal histories from fission track data. Earth Planetary Science Letters, 136: 421435.CrossRefGoogle Scholar
Gallagher, K., 2012. Uplift, denudation, and their causes and constraints over geological timescales. In: Roberts, D. G., and Bally, A. W. (Eds.), Regional Geology and Tectonics: Principles of Geologic Analysis. Elsevier, New York, pp. 609644.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: Cameron, N. R., Bate, R. H., and Clure, V. S. (Eds.), The Oil and Gas Habitats of the South Atlantic. Geological Society, London, pp. 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(B9): 1811718144.CrossRefGoogle Scholar
Galloway, W. E., 1989. Depositional framework and hydrocarbon resources of the early Miocene (Fleming) episode, Northwest Gulf Coast basin. Marine Geology, 90(1–2): 1929.CrossRefGoogle Scholar
Gamond, J. F., 1987. Bridge structures as sense of displacement criteria in brittle fault zones. Journal of Structural Geology 9: 609620.CrossRefGoogle Scholar
Garfunkel, Z., 1986. Review of oceanic transform activity and development. Journal of the Geological Society, London. 143: 775784.CrossRefGoogle Scholar
Garfunkel, Z., 2014. Lateral motion and deformation along the Dead Sea Transform. In: ed. Garfunkel, Z., Ben-Avraham, Z., and Kagan, E. (Eds.), Dead Sea Transform Fault System: Reviews, Springer, New York, pp. 109150.CrossRefGoogle Scholar
Garfunkel, Z., and Ben-Avraham, Z., 1996. The structure of the Dead Sea basin. Tectonophysics, 266: 155176.CrossRefGoogle Scholar
Gartman, A., and Hein, J. R., 2019. Mineralization at oceanic transform faults and fracture zones. In: Duarte, J. (Ed.), Transform Plate Boundaries and Fracture Zones. Elsevier, New York, pp. 105118.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: 10851124.CrossRefGoogle 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 result. American Journal of Science, 284: 11251174.CrossRefGoogle Scholar
Gasperini, L., Bonatti, E., Ligi, M., et al., 1997. Stratigraphic and paleoenvironmental analysis of a carbonate platform on the Romanche transverse ridge, equatorial Atlantic. Marine Geology, 136: 245257.CrossRefGoogle Scholar
Gasperini, L., Bernoulli, D., Bonatti, E., et al., 2001. Lower Cretaceous to Eocene sedimentary transverse ridge at the Romanche Fracture Zone and the opening of the equatorial Atlantic. Marine Geology, 176: 101119.CrossRefGoogle Scholar
Gayer, R. A., Hathaway, T. M., and Nemčok, M., 1998. Transpressionally driven rotation in the external orogenic zones of the Western Carpathians and the SW British Variscides. In: Holdsworth, R. E., Strachan, R. A., and Dewey, J. F. (Eds.), Continental Transpressional and Transtensional Tectonics. Geological Society, London, pp. 253266.Google Scholar
Gaža, B., 1973. Fault classification for the Neogene fill of the Danube Basin (in Slovak). Report 1627, Archive Nafta Gbely, 3 enclosures.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: 91199144.CrossRefGoogle Scholar
GEBCO Compilation Group, 2019. GEBCO 2019 Grid. DOI: 10.5285/836f016a-33be-6ddc-e053-6c86abc0788eCrossRefGoogle Scholar
GEBCO Compilation Group, 2020. GEBCO 2020 Grid. DOI: 10.5285/a29c5465-b138-234d-e053-6c86abc040b9CrossRefGoogle Scholar
Geiser, P., 1974. Cleavage in some sedimentary rocks of the central Valley and Ridge province, Maryland. Geological Society of America Bulletin, 85: 13991412.2.0.CO;2>CrossRefGoogle Scholar
Géli, L., Piau, J.-M., Dziak, R., et al., 2014. Seismic precursors linked to highly compressible fluids at oceanic transform faults. Nature Geoscience, Letters, 7: 757761.CrossRefGoogle Scholar
Gensterblum, Y., Ghanizadeh, A., Cuss, R. J., et al., 2015. Gas transport and storage capacity in shale gas reservoirs: a review. Part A: transport processes. Journal of Unconventional Oil and Gas Resources, 12: 87122.CrossRefGoogle Scholar
Geoffroy, L., 2005. Volcanic passive margins. Comptes Rendus Geosciences, 337)(16): 13951408.CrossRefGoogle Scholar
Geoffroy, L., Burov, E. B., and Werner, P., 2015. Volcanic passive margins: another way to break up continents. Nature Scientific Reports, 5: 14828.CrossRefGoogle ScholarPubMed
Georgen, J. E., Lin, J., and Dick, H. J. B., 2001. Evidence from gravity anomalies for interactions of the Marion and Bouvet hotspots with the Southwest Indian Ridge: effects of transform offsets. Earth and Planetary Science Letters, 187: 283300.CrossRefGoogle Scholar
Geoscience Australia, 2013. Regional geology of the Perth Basin. Australia 2013. Offshore Petroleum Exploration Acreage Release, Australian Government, Department of Resources, Energy and Tourism. Available at: www.petroleum-acreage.gov.au/files/files/2013/documents/regionalgeology/Regional_Geology-North_Perth.pdf (accessed June 2, 2015).Google Scholar
Geotrack International, 1999a. (U–Th)/He dating of apatite. Improved resolution at low temperatures. Geotrack Information Sheet 99/8.Google Scholar
Geotrack International, 1999b. Thermal history interpretation of AFTA data. Geotrack Information Sheet 99/2.Google Scholar
Gerard, J., 2009. Stratigraphic traps: quantitative approach based upon a producing field database. Search and Discovery Article #40436. Posted September 22, 2009. Adapted from poster presentation at AAPG Annual Convention, Denver, Colorado, June 7–10, 2009.Google Scholar
Gerber, T. P., Amblas, D., Wolinsky, M. A., Pratson, L. F., and Canals, M., 2009, A model for the long-profile shape of submarine canyons: Journal of Geophysical Research, 114: F03002.CrossRefGoogle Scholar
Geršlová, J., Opletal, V., Sýkorová, I., Sedláková, I., and Geršl, M., 2015. A geochemical and petrographical characterization of organic matter in the Jurassic Mikulov Marls from the Czech Republic. International Journal of Coal Geology, 141–142: 4250.CrossRefGoogle Scholar
Gerya, T., 2010. Dynamic instability produces transform faults at mid-ocean ridges. Science, 329: 10471050.CrossRefGoogle ScholarPubMed
Gerya, T., 2012. Origin and models of oceanic transform faults. Tectonophysics, 522–523: 3454.CrossRefGoogle Scholar
Gerya, T., 2013a. Three-dimensional thermomechanical modeling of oceanic spreading initiation and evolution. Physics of the Earth and Planetary Interiors, 214: 3552.CrossRefGoogle Scholar
Gerya, T., 2013b. Initiation of transform faults at rifted continental margins: 3D petrological–thermomechanical modeling and comparison to the Woodlark Basin. Petrology, 21(6): 550560.CrossRefGoogle Scholar
Gerya, T., 2016. Origin, evolution, seismicity, and models of oceanic and continental transform boundaries. In: Duarte, J. C., and Schellart, W. (Eds.), Plate Boundaries and Natural Hazards. American Geophysical Union, Washington, DC, pp. 3976.CrossRefGoogle Scholar
Geyh, M. A., and Schleicher, H., 1990. Absolute Age Determination: Physical and Chemical Dating Methods and Their Application. Springer, Berlin.CrossRefGoogle Scholar
Gholamrezaie, E., Scheck-Wenderoth, M., Bott, J., Heidbach, O., and Strecker, M. R., 2019. 3-D crustal density model of the Sea of Marmara. Solid Earth, 10: 785807.CrossRefGoogle Scholar
Giannenas, P. A., 2018. The structural development of the Vestbakken volcanic province, Western Barents Sea: relation between faults and folds. Master’s thesis. University of Oslo.Google Scholar
Gibbons, A. D., Whittaker, J. M., and Müller, R. D., 2013. The breakup of East Gondwana: assimilating constraints from Cretaceous ocean basins around India into a best-fit tectonic model. Journal of Geophysical Research: Solid Earth, 118: 808822.CrossRefGoogle Scholar
Gibbs, A. D., 1984. Structural evolution of extensional basin margins. Journal of the Geological Society of London, 141(4): 609620.CrossRefGoogle Scholar
Gibson, G. M., Totterdell, J. M., White, L. T., et al., 2013. Pre-existing basement structure and its influence on continental rifting and fracture development along Australia’s southern rifted margin. Journal of the Geological Society of London, 170(2): 365377.CrossRefGoogle Scholar
Gilchrist, A. R., and Summerfield, M. A., 1990. Differential denudation and flexural isostasy, and long-lived escarpments: a numerical modeling study. Journal of Geophysical Research, 99: 1222912243.Google Scholar
Gilchrist, A. R., and Summerfield, M. A., 1991. Denudation, isostasy and landscape evolution. Earth Surface Processes and Landforms, 16: 555562.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(2–3): 205222.Google Scholar
Glantschnig, J., and Kroell, E., 1997. Injection of nitrogen for improved oil recovery: a successful case history. Presented at the 15th World Petroleum Congress, October, 12–17, 1997, Beijing, China.CrossRefGoogle Scholar
Godfrey, N. J., Fuis, G. S., Langenheim, V., Okaya, D. A., and Brocher, T. M., 2002. Lower crustal deformation beneath the central Transverse Ranges, southern California: results from the Los Angeles Region Seismic Experiment. Journal of Geophysical Research: Solid Earth, 107(B7). DOI 10.1029/2001JB000354CrossRefGoogle Scholar
Goldsby, D. L., and Tullis, T. E., 2011. Flash heating leads to low frictional strength of crustal rocks at earthquake slip rates. Science, 334: 216218.CrossRefGoogle ScholarPubMed
Goldstein, R.H., 2003. Petrographic analysis of fluid inclusions. In: Samson, I., Anderson, A., and Marshall, D. (Eds.), Fluid Inclusions: Analysis and Interpretation. Mineralogical Association of Canada, Quebec City, pp. 953.CrossRefGoogle Scholar
Golonka, J., 2000. Cambrian–Neogene Plate Tectonic Maps. Wydawnictwo Uniwersytetu Jagiellonskiego, Krakow.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 off NE Brazil: results of LEPLAC Project. In: Mohriak, W., and Talwani, M. (Eds.), Atlantic Rifts and Continental Margins. American Geophysical Union, Washington, DC, pp. 261292.CrossRefGoogle Scholar
Gonzáles, G. L., Dunai, T., Carrizo, D., and Allmendinger, R., 2006. Young displacements on the Atacama Fault System, northern Chile from field observations and cosmogenic 21Ne concentrations. Tectonics, 25. DOI: 10.1029/2005TC001846Google Scholar
Gopalakrishnan, K., 1998. Extensions of Eastern Ghats mobile Belt, India: a geological enigma. Proceedings of Workshop on the Eastern Ghats Mobile Belt.Google Scholar
Gosnold, W. D., and Panda, B., 2002. The global heat flow database of the International Heat Flow Commission. Available at: www.und.edu/org/ihfc/index2.html.Google Scholar
Gosse, J. C., and Phillips, F. M., 2001. Terrestrial in situ cosmogenic nuclides: theory and application. Quaternary Science Reviews, 20: 14751560.CrossRefGoogle Scholar
Goutorbe, B., Poort, J., Lucazeau, F., and Raillard, S., 2011. Global heat flow trends resolved from multiple geological and geophysical proxies. Geophysical Journal International, 187: 14051419.CrossRefGoogle Scholar
Gouyet, S., 1988. Tectono-sedimentary evolution of the Guyana and Northern Brazilian margins during the South Atlantic evolution [in French]. PhD thesis. University of Pau.Google Scholar
Govers, R., and Wortel, M. Jr., 2005. Lithosphere tearing at STEP faults: response to edge of subduction zones. Earth and Planetary Science Letters, 236(1–2): 505523.CrossRefGoogle Scholar
Gradstein, F. M., Ludden, J. N., Adamson, A. C., et al., 1990. Site 765. In: Proceedings of the Ocean Drilling Program, Part A: Initial Reports, Volume 123. Texas A&M University, College Station, pp. 63267.Google Scholar
Gradstein, F. M., Ludden, J. N., Adamson, A. C., et al.,1992. Leg 122–123, northwestern Australian margin; a stratigraphic and paleogeographic summary. In: Proceeding of the Ocean Drilling Program, Scientific Results, Volume 123. Texas A&M University, College Station, pp. 801816.Google Scholar
Graham, S. A., 1976. Tertiary sedimentary tectonics of the central Salinian Block of California. PhD thesis. Stanford University, California.Google Scholar
Granger, D. A., Kirchner, J. W., and Finkel, R. C., 1996. Spatially averaged long-term erosion rates measured from in situ-produced cosmogenic nuclides in alluvial sediments. Journal of Geology, 104: 249257.CrossRefGoogle Scholar
Grant, C. J., 2019. Predicting lateral seal development in deepwater stratigraphic traps of the Atlantic Basin. HGS PESGB 19th African Conference, October 1–2, 2019.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: 155165.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: 845856.CrossRefGoogle Scholar
Gray, N. H., 1988. The origin of copper occurrences in the Hartford basin. In: Froelich, A. J., and Robinson, G. R., Jr. (Eds.), Studies of the Early Mesozoic Basins of the Eastern U.S. US Geological Survey, Reston, pp. 341349.Google Scholar
Green, P. F., 1981. A new look at statistics in fission-track dating. Nuclear Tracks, 5: 7786.CrossRefGoogle Scholar
Green, P. F., Duddy, I. R., Laslett, G. M., et al., 1989. Thermal annealing of fission tracks in apatite 4: quantitative modeling techniques and extension to geological timescales. Chemical Geology, 79: 155182.Google 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: Dore, A. G., Cartwright, J., Stoker, M. S., Turner, J. P., and White, N. (Eds.), Exhumation of the North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum Exploration. Geological Society of London, London, pp. 331354.Google 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. 117–131.Google Scholar
Greene, H. G., and Hicks, K. R., 1990. Ascension–Monterey canyon system: history and development. In: Garrison, R. E., Greene, H. G., Hicks, K. R., Weber, G. E., and Wright, T. L. (Eds.), Geology and Tectonics of the Central California Coastal Region, San Francisco to Monterey. AAPG, Washington, DC, pp. 229250.Google Scholar
Greenfield, S., Cox, P., Staffurth, N., and McAfee, A., 2019. Evolution of sandstone composition through the Cretaceous and its impact on reservoir quality along the West African Transform Margin. Abstract. P94 in Abstract Volume. HGS PESGB 18th African Conference, October 1–2, 2019.Google Scholar
Greenroyd, C. J., Peirce, C., Rodger, M., Watts, A. B., and Hobbs, R. W., 2007. Crustal structure of the French Guiana margin, West Equatorial Atlantic. Geophysical Journal International, 169: 964987.CrossRefGoogle Scholar
Gregg, P. M., Lin, J., and Smith, D. K., 2006. Segmentation of transform systems on the East Pacific Rise: implications for earthquake processes at fast-slipping oceanic transform faults. Geology, 34: 289292.CrossRefGoogle Scholar
Gregg, P. M., Lin, J., Behn, M. D., and Montesi, L. G. J., 2007. Spreading rate dependence of gravity anomalies along oceanic transform faults. Nature Letters, 448(12): 183187.CrossRefGoogle ScholarPubMed
Gregg, P. M., Behn, M. D., Lin, J., and Grove, T. L., 2009. Melt generation, crystallization, and extraction beneath segmented oceanic transform faults. Journal of Geophysical Research, 114(B11102). DOI: 10.1029/2008JB006100CrossRefGoogle Scholar
Gregory, A. R., 1977. Aspects of rock physics from laboratory and log data that are important to seismic interpretation. In: Payton, C. E. (Ed.), Seismic Stratigraphy, Applications to Hydrocarbon Exploration. AAPG, Washington, DC, pp. 1546.Google Scholar
Grew, E. S., and Manton, W. I., 1986. A new correlation of sapphirine granulites in the India–Antarctic metamorphic terrain: late Proterozoic dates from the Eastern Ghats province of India. Precambrian Research, 33: 123137.CrossRefGoogle Scholar
Griggs, D. T., and Blacic, J. D., 1965. Quartz: anomalous weakness of synthetic crystals. Science, 147(3655): 292295.CrossRefGoogle ScholarPubMed
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. Doctoral thesis. University of Miami (Florida), Coral Gables.Google Scholar
Grunau, H. R., 1981. Worldwide review of seals for major accumulation of natural gas. AAPG Bulletin, 65: 933.Google Scholar
Grunow, A., Hanson, R., and Wilson, T., 1996. Were aspects of Pan-African deformation linked to Iapetus opening? Geology, 24(12): 10631066.2.3.CO;2>CrossRefGoogle Scholar
Gudmundsson, A., 1999. Fluid overpressure and stress drop in fault zones. Geophysical Research Letters, 26: 115118.CrossRefGoogle Scholar
Gudmundsson, A., 2000. Dynamics of volcanic systems in Iceland. Annual Reviews of Earth and Planetary Sciences, 28: 107140.CrossRefGoogle Scholar
Gudmundsson, A., and Homberg, C., 1999. Evolution of stress fields and faulting in seismic zones. Pure and Applied Geophysics, 154: 257280.CrossRefGoogle Scholar
Gudmundsson, A., Brynjolfsson, S., and Jonsson, M. T., 1993. Structural analysis of a transform fault-rift zone junction in North Iceland. Tectonophysics, 220: 205221.CrossRefGoogle Scholar
Guillou, H., Turpin, L., and Garcia, M. O., 1996. Unspiked K–Ar dating of young submarine volcanic rocks from Loihi and Mauna Loa. Eos, Transactions American Geophysical Union, 77(46): 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: 8188.CrossRefGoogle Scholar
Guiraud, R., and Maurin, J. C., 1991. Le Rifting en Afrique au Cretace inferrieur: synthese structurale, mise en evidence de deux etapes dans la genese des basins, relations avec les ouvertures oceaniques pre-africanes. Bulletin de la Societe Geologique de France, 162(5): 811823.CrossRefGoogle Scholar
Guiraud, R., and Maurin, J. C., 1992. Early Cretaceous rifts of Western and Central Africa: an overview. Tectonophysics, 213: 153168.CrossRefGoogle Scholar
Guiraud, R., Bellion, Y., Benkhelil, J., and Moreau, C., 1987. Post-Hercynian tectonics in Northern and Western Africa. African Geology Reviews: Geological Journal, 22, 433466.Google Scholar
Guliyev, I. S., Tagiyev, M. F., and Feyzullayev, A. A., 2001. Geochemical characteristics of organic matter from Maykop rocks of eastern Azerbaijan. Lithology and Mineral Resources, 36(3): 280285.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
Guo, Y., and Morgan, J. K., 2006. The frictional and micromechanical effects of grain comminution in fault gouge from distinct element simulations. Journal of Geophysical Research: Solid Earth, 111(B12). DOI: 10.1029/2005JB004049CrossRefGoogle 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: 167189.CrossRefGoogle Scholar
Gussow, W. C., 1954. Differential entrapment of gas and oil: a fundamental principle. AAPG Bulletin, 38: 816853.Google Scholar
Gustavson Associates, 2012. Resource evaluation report on the Corentyne Petroleum Prospecting License, Guyana, South America. Available at: http://cgxenergy.ca/cmsassets/docs/pdf/51–101gustavsonreport-apr15–11.pdf (accessed October 19, 2020).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(101): 1523.Google Scholar
Guterch, A., Grad, M., Janik, T., et al., 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(12): 14891496.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: 11671176.2.3.CO;2>CrossRefGoogle Scholar
Haberland, C., Agnon, A., El-Kelani, R., et al., 2003. Modeling of seismic guided waves at the Dead Sea Transform. Journal of Geophysical Research Solid Earth, 108(B7): DOI: 10.1029/2002JB002309Google Scholar
Haenel, R., Rybach, L., and Stegena, L. (Eds.), 1988. Handbook of Terrestrial Heat-Flow Density Determination. Kluwer Academic, Dordrecht.CrossRefGoogle Scholar
Hafid, M., Zizi, M., Bally, A. W., and Ait Salem, A., 2006. Structural; styles of the western onshore and offshore termination of the High Atlas, Morocco. Comptes Rendus Geoscience, 338: 5064.CrossRefGoogle Scholar
Hall, L. S., Sanchez, G., Borissova, I., et al., 2017. Crustal structure and tectonic evolution of the northern Perth Basin, Australia. AAPG Search and Discovery Article 11027.Google Scholar
Hall, R., 1997. Cenozoic tectonics of SE Asia and Australasia. In: Howes, J. V. C. and Noble, R. A. (Eds.), Proceedings of an International Conference on Petroleum Systems of SE Asia and Australasia, Jakarta, Indonesia, May 21–23, 1997. Indonesian Petroleum Association, Jakarta, pp. 4762.Google Scholar
Hallam, A., 1984. Pre-Quaternary sea-level changes. Annual Review of Earth and Planetary Science, 12: 205243.CrossRefGoogle Scholar
Hamilton, W., Wagner, L., and Wessely, G., 2000. Oil and gas in Austria. Mitteilungen der Osterreichischen Geologischen Gesellschaft, 9: 235262.Google Scholar
Hammond, D. E., Leslie, B. W., Ku, T.-L., 1988. 222Rn concentrations in deep formation waters and the geohydrology of the Cajon Pass borehole. Geophysical Research Letters, 15(9): 10451048.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., 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: 457484.CrossRefGoogle 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 Hampson, G. J., Steel, R. J., Burgess, P. M. and Dalrymple, R. W. (Eds.), Recent Advances in Models of Siliciclastic Shallow Marine Stratigraphy, SEPM, Tulsa, pp. 117142.CrossRefGoogle Scholar
Hanchar, J. M., and Miller, C. F., 1993. Zircon zonation patterns as revealed by cathodoluminiscence and backscattered electron images: implications for interpretation of complex crustal histories. Chemical Geology, 110: 113.CrossRefGoogle Scholar
Handin, J., Hager, R. V., Friedman, M., and Feather, J. N., 1963. Experimental deformation of sedimentary rocks under confining pressure: pore pressure tests. AAPG Bulletin, 47: 717755.Google Scholar
Handy, M., Wissing, S. B., and Streit, L. E., 1999. Frictional viscous flow in mylonite with varied bimineralic composition and its effect on lithospheric strength. Tectonophysics, 303(1): 175191.CrossRefGoogle Scholar
Hannington, M., Herzig, P., Stoffers, P., et al., 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 Barnes, H. L. (Eds.), Geochemistry of Hydrothermal Ore Deposits. Wiley, New York, pp. 137168.Google 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: 113127.CrossRefGoogle 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. AAPG Bulletin, 93(2): 155179.CrossRefGoogle Scholar
Haq, B. U., 1991. Sequence Stratigraphy, Sea-Level Change and Significance for the Deep Sea. International Association of Sedimentologists, Gent, pp. 339.Google Scholar
Haq, B. U., Hardenbol, J., and Vail, P.R., 1987. Chronology of fluctuating sea levels since the Triassic. Science, 235: 11561167.CrossRefGoogle ScholarPubMed
Haq, B. U., Hardenbol, J., and Vail, P.R., 1988. Mesozoic and Cenozoic chronostratigraphy and cycles of sea level changes. In: Wilgus, C. K., Hastings, B. S., Kendall, C. G. C., et al. (Eds.), Sea-Level Changes: An Integrated Approach. SEPM, Tulsa, pp. 71108.CrossRefGoogle Scholar
Haq, B. U., von Rad, U., O’Connel, S., et al., 1990. Site 763: Proceeding of the Ocean Drilling Program, Initial Report, 122. Texas A&M University, College Station, pp. 289352.CrossRefGoogle Scholar
Harash, A., and Bar, Y., 1988. Faults, landslides and seismic hazards along the Jordan River Gorge, northern Israel. Engineering Geology, 25: 115.CrossRefGoogle Scholar
Hardcastle, K. C., 1989. Possible paleostress tensor configurations derived from fault-slip data in eastern Vermont and western New Hampshire. Tectonics 8: 265284.CrossRefGoogle Scholar
Hardenbol, J., Thierry, J., Farley, M., et al., 1998. Mesozoic and Cenozoic sequence chronostratigraphic chart. In: Hardenbol, J., Thierry, J., Farley, M. B., et al. (Eds.), Mesozoic and Cenozoic Sequence Chronostratigraphic Framework of European Basins. SEPM, Tulsa, pp. 314.Google Scholar
Harding, T. P., 1974. Petroleum traps associated with wrench faults. AAPG Bulletin, 58: 12901304.Google Scholar
Harding, T. P., 1976. Tectonic significance and hydrocarbon trapping consequences of sequential folding synchronous with San Andreas faulting, San Joaquin Valley, California. AAPG Bulletin, 69: 582600.Google Scholar
Harding, T. P., 1990. Identification of wrench faults using subsurface structural data: criteria and pitfalls. AAPG Bulletin, 74: 15901609.Google Scholar
Harding, T. P., and Lowell, J. D., 1979. Structural styles their plate tectonic habitats, and hydrocarbon traps in petroleum provinces. AAPG Bulletin, 63: 10161058.Google Scholar
Harding, T. P., Vierbuchen, R. C., and Christie-Blick, N., 1985. Structural styles, plate-tectonic settings, and hydrocarbon traps of divergent (transtensional) wrench faults. In: Biddle, K. T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation and Sedimentation. SEPM, Tulsa,, pp. 5177.CrossRefGoogle Scholar
Hardy, S., Poblet, J., McClay, K.R., and Waltham, D., 1996. Mathematical modeling of growth strata associated with fault-related fold structures. In: Buchanan, P. G. and Nieuwland, D. A. (Eds.), Modern Developments in Structural Interpretation, Validation and Modeling, Geological Society of London, London, pp. 265282.Google Scholar
Harland, W. B., 1965. The tectonic evolution of the Arctic–North Atlantic region. Philosophical Transactions Royal Society of London Series A, 258: 5975.Google Scholar
Harland, W.B., 1969. Contribution of Spitsbergen to understanding of tectonic evolution of North Atlantic region. In: May, K. (Ed.), North Atlantic: Geology and Continental Drift. AAPG, Washington, DC, pp. 817851.Google Scholar
Harland, W. B., 1971. Tectonic transpression in Caledonian Spitsbergen. Geological Magazine, 108: 2742CrossRefGoogle Scholar
Harp, E. L., Wells, W. G. II., and Sarmiento, J. G., 1990. Pore pressure response during failure in soils. Geological Society of America Bulletin, 102: 428438.2.3.CO;2>CrossRefGoogle Scholar
Harzhauser, M., Theobalt, D., Strauss, P., Mandic, O., and Piller, W. E., 2018. Seismic-based lower and middle Miocene stratigraphy in the northwestern Vienna Basin (Austria). Newsletters on Stratigraphy. DOI: 10.1127/nos/2018/0490CrossRefGoogle Scholar
Hasterok, D., and Chapman, D. S., 2011. Heat production and geotherms for the continental lithosphere. Earth and Planetary Science Letters, 307: 5970CrossRefGoogle Scholar
Hatcher, R. D., Osberg, P. H., and Drake, A. A. Jr., 1986. Tectonic map of the U.S. Appalachians: DEAG Appalachians – Ouachitas Volume, Plate 1.Google Scholar
Hauksson, E., 2000. Crustal structure and seismicity distribution adjacent to the Pacific and North America plate boundary in southern California. Journal of Geophysical Research Solid Earth, 105: 1387513903.CrossRefGoogle Scholar
Hauksson, E., Jones, L. M., and Hutton, K., 2002. The 1999 Mw 7.1 Hector Mine, California, Earthquake sequence: complex conjugate strike-slip faulting. Bulletin of the Seismological Society of America, 92: 11541170.CrossRefGoogle Scholar
Hayes, D. E., Frakes, L. A., Barrett, P. J., et al., 1975. Site 264. In: Initial Reports of the Deep Sea Drilling Project, Volume 28. Texas A&M University, College Station, pp. 1948.Google Scholar
Hayward, N. J., and Ebinger, C. J., 1996. Variations in the along-axis segmentation of the Afar Rift system. Tectonics, 15: 244257.CrossRefGoogle Scholar
Hearn, E. H., McClusky, S., Ergintav, S., and Reilinger, R. E., 2009. Izmit earthquake postseismic deformation and dynamics of the North Anatolian Fault Zone. Journal of Geophysical Research: Solid Earth, 114(B8). DOI: 10.1029/2008JB006026CrossRefGoogle Scholar
Hedberg, H. D., 1936. Gravitational compaction of clays and shales. American Journal of Science, 31: 241287.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 (3–4): 299322.Google Scholar
Heine, C., Müller, R. D., and Gaina, C., 2004. Reconstructing the lost eastern Tethys ocean basin: convergence history of the SE Asian margin and marine gateways. In: Clift, P., Wang, P., Kuhnt, W., and Hayes, D. (Eds.), Continent-Ocean Interactions Within East Asian Marginal Seas. AGU, Washington, DC, pp. 3754.CrossRefGoogle Scholar
Heiri, O., Lotter, A.F., and Lemcke, G. 2001. Loss-on-ignition as a method for estimating organic and carbonate content in sediments: reproducibility and comparability of results. Journal of Paleolimnology 25: 101110.CrossRefGoogle Scholar
Heirtzler, J. R., Veevers, J. J., Bolli, H. M., et al., 1974a. Site 260. In: Initial Reports of the Deep Sea Drilling Project, Volume 27. Texas A&M University, College Station, pp. 89127.Google Scholar
Heirtzler, J.R., Veevers, J. J., Bolli, H.M., et al., 1974b. Site 259. In: Initial Reports of the Deep Sea Drilling Project, Volume 27. Texas A&M University, College Station, pp. 1587.Google Scholar
Hekinian, R., Bideau, D., Cannat, M., Francheteau, J., and Hebert, R., 1992. Volcanic activity and crust mantle exposure in the ultrafast Garrett transform-fault neat 13°28′S in the Pacific. Earth and Planetary Science Letters, 108: 259275.CrossRefGoogle Scholar
Hekinian, R., Bideau, D., Hebert, R., and Niu, Y. L., 1995. Magmatism in the Garrett transform-fault (East Pacific Rise near 13°27′S). Journal of Geophysical Research, 100: 1016310185.CrossRefGoogle 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, 24: 113128.Google Scholar
Hellinger, S. J., and Sclater, J. G., 1983. Some comments on two-layer extensional models for the evolution of sedimentary basins. Journal of Geophysical Research, 88(B10): 82518269.CrossRefGoogle Scholar
Hempton, M., 1987. Constraints on Arabian Plate motion and extensional history of the Red Sea. Tectonics, 6(6): 687705.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(6): 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: 409438.CrossRefGoogle Scholar
Henk, A., 2006. Stress and strain during fault-controlled lithospheric extension: insights from numerical experiments. Tectonophysics, 415: 3955.CrossRefGoogle Scholar
Henk, A., and Nemčok, M., 2015. Lower crust ductility patterns associated with transform margins. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L. (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London. DOI: 10.1144/SP431.9Google Scholar
Henry, A. A., and Lewan, M. D., 2001. Comparison of kinetic-model prediction of deep gas generation. In Dyman, T. S. and Kuuskraa, V. A. (Eds.), Geologic Studies of Deep Natural Gas Resources USGS, Reston, pp. 125.Google Scholar
Hensen, C., Duarte, J. C., Vannucchi, P., et al., 2019a. Marine transform faults and fracture zones: a joint perspective integrating seismicity, fluid flow and life. Frontiers in Earth Science, 7: 129.CrossRefGoogle Scholar
Henstock, T. J., Levander, A., and Hole, J. A., 1997. Deformation in the Lower Crust of the San Andreas Fault System in Northern California. Science, 278 (5338): 650653.CrossRefGoogle Scholar
Herbert, T. D., and Fischer, A. G., 1986. Milankovich climatic origin of mid-Cretaceous black shale rhythms in central Italy: Nature, 329: 739743.CrossRefGoogle Scholar
Hermeston, S., and Nemčok, M., 2013. Thick-skin orogen–foreland interaction and their controlling factors, Northern Andes of Colombia. In: Nemčok, M., Mora, A., and Cosgrove, J. W. (Eds.), Thick-Skin-Dominated Orogens: From Initial Inversion to Full Accretion. Geological Society of London, London. DOI: 10.1144/SP377.16.Google Scholar
Hermeston, S., Torres, M., Kean, A., Ramos, D., and Brisson, I., 2015. New exploration in the proven petroleum system of the Suriname offshore basin, Suriname. EAGE 67th Conference & Exhibition – Madrid, Spain, June 13–16, 2005.Google Scholar
Herquel, G., Tapponnier, P., Wittlinger, G., Mei, J., and Danian, S., 1999. Teleseismic shear wave splitting and lithospheric anisotropy beneath and across the Altyn Tagh Fault. Geophysical Research Letters, 26(21): 32253228.CrossRefGoogle 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(B11): 1296912980.CrossRefGoogle Scholar
Hess, 2018b. Hess announces eighth oil discovery offshore Guyana, 06/20/2018. www.hess.com/newsroom/news-article/2018-06-20-hess-announces-eighth-oil-discovery-offshore-guyanaGoogle Scholar
Hess, 2018c. Hess announces 10th discovery and increased Stabroek Block resource estimate offshore Guyana, 12/03/2018. https://investors.hess.com/news-releases/news-release-details/hess-announces-10th-discovery-and-increased-stabroek-block.Google Scholar
Hess, 2020. Timothy Chisholm. Hess’s journey into an emerging superbasin and ultra-high impact exploration for an independent E&P Company. HGS Annual University of Houston Sheriff Lecture, November 9.Google Scholar
Hilairet, N., Reynard, B., Wang, Y., et al., 2007. High-pressure creep of serpentine, interseismic deformation, and initiation of subduction. Science, 318: 19101913.CrossRefGoogle ScholarPubMed
Hill, D. P., Eaton, J. P., and Jones, L. M., 1990. Seismicity 1980–86. USGS, Reston.Google Scholar
Hills, E. S., 1963. Elements of Structural Geology. Methuen and Co. Ltd., London.Google Scholar
Hindle, A. D., 1997. Petroleum migration pathways and charge concentration: a three-dimensional model. AAPG Bulletin, 81: 14511481.Google Scholar
Hladecek, K., 1965. Sequence stratigraphy and deposition of the Helvetian strata in Matzen [in German]. PhD Thesis, University of Vienna.Google Scholar
Hochella, M. F., Jr., Eggleston, C. M., Elings, V. B., and Thompson, M. S., 1990. Atomic structure and morphology of the albite {010} surface: an atomic-force microscope and electron diffraction study. American Mineralogist, 75: 723730.Google Scholar
Hochstein, M. P., and Browne, P. R. L., 2000. Surface manifestations of geothermal systems with volcanic heat sources. In: Sigurdsson, H., Houghton, B. F., McNutt, S. R., Rymer, H., and Stix, J. (Eds.), Encyclopedia of Volcanoes, Academic Press, San Diego, pp. 835855.Google Scholar
Hofstetter, A., Rybakov, M., and ten Brink, U., 1998. Comment on “New evidence of magmatic diapirs in the intermediate crust under the Dead Sea, Israel” by Rabinowitz, N., Steinberg, J., and Mart, Y. Tectonics, 17(5): 819820.CrossRefGoogle Scholar
Hohenegger, J., Ćorić, S., and Wagreich, M., 2014. Timing of the middle Miocene Badenian stage of the central Paratethys. Geologica Carpathica, 65: 5566.CrossRefGoogle 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. AAPG Bulletin, 68: 753764.Google Scholar
Hók, J., Kováč, M., Pelech, O., et al., 2016. The Alpine tectonic evolution of the Danube Basin and its northern periphery (southwestern Slovakia). Geologica Carpathica, 67(5): 495505.CrossRefGoogle Scholar
Holbrook, W. S., Brocher, T. M., ten Brink, U. S., and Hole, J. A., 1996. Crustal structure of a transform plate boundary: San Francisco Bay and the central California continental margin. Journal of Geophysical Research, Solid Earth, 101: 2231122334.CrossRefGoogle Scholar
Holloway, J. R., 1977. Fugacity and activity of molecular species in supercritical fluids. In: Fraser, D. G. (Ed.), Thermodynamics in Geology. D. Reidel Publishing Company, Dordrecht, pp. 161181.CrossRefGoogle Scholar
Holm, D. K., Norris, R. J., and Craw, D., 1989. Brittle and ductile deformation in a zone of rapid uplift: Central Southern Alps, New Zealand. Tectonics, 8(2): 153168.CrossRefGoogle Scholar
Holyoke, C. W., III and Tullis, J., 2006. Formation and maintenance of shear zones. Geology, 34(2): 105108.CrossRefGoogle Scholar
Hölzel, M., Wagreich, M., Faber, R., and Strauss, P., 2008. Regional subsidence analysis in the Vienna Basin (Austria). Austrian Journal of Earth Science, 97: 3850.Google 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: 703718.CrossRefGoogle Scholar
Honnorez, J., Villeneuve, M., and Mascle, J., 1994. Old continent-derived metasedimentary rocks in the equatorial Atlantic: an acoustic basement outcrop along the fossil trace of the Romanche transform fault at 6°30′W. Marine Geology, 117: 237251.CrossRefGoogle Scholar
Hopper, J. R., Mutter, J. C., Larson, R. L., et al. 1992. Magmatism and rift margin evolution: evidence from Northwest Australia. Geology, 20: 853857.2.3.CO;2>CrossRefGoogle Scholar
Hopper, J. R., Funck, T., Tucholke, B. E., et al. 2004. Continental breakup and the onset of ultraslow seafloor spreading off Flemish Cap on the Newfoundland rifted margin. Geology, 32: 9396.CrossRefGoogle Scholar
Horita, J., and Berndt, M. E., 1999. Abiogenic methane formation and isotopic fractionation under hydrothermal conditions. Science, 285: 10551057.CrossRefGoogle ScholarPubMed
Horsfield, B., and Rullkötter, J., 1994. Diagenesis, catagenesis, and metagenesis of organic matter. In: Magoon, L. B., and Dow, W. G. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 189199.Google Scholar
Horváth, F., 1993. Towards a mechanical model for the formation of the Pannonian basin. Tectonophysics 226: 333357.CrossRefGoogle Scholar
Horváth, F., Musitz, B., Balazs, A., and Uhrin, A., 2015. Evolution of the Pannonian basin and its geothermal resources. Geothermics. DOI: 10.1016/j.geothermics.2014.07.009.CrossRefGoogle 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(3–4): 167179.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: 167171.Google Scholar
Hovland, M., and Judd, A. G., 1988. Seabed Pockmarks and Seepages. Graham and Trotman, London.Google 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
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. AAPG Bulletin, 37: 19542026.Google Scholar
Hubbert, M. K., 1969. The Theory of Ground-Water Motion and Related Papers. London: Hafner Publ. Co.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.Google Scholar
Huismans, R. S., and Beaumont, C., 2005. Effect of lithospheric stratification on extensional styles and rift basin geometry. In: Post, P. J., Rosen, N., Olson, D.L., et al. (Eds.), Petroleum Systems of Divergent Continental Margin Basins, 25th Annual GCSSEPM Foundation Bob F. Perkins Research Conference, Session I, Crustal Architecture of Divergent Margins. GCSSEPM, Houston, pp. 1255.CrossRefGoogle Scholar
Huismans, R. S., and Beaumont, C., 2008. Complex rifted continental margins explained by dynamical models of depth-dependent lithospheric extension. Geology, 36(2): 163166.CrossRefGoogle Scholar
Huismans, R. S., and Beaumont, C., 2011. Depth-dependent extension, two-stage breakup and cratonic underplating at rifted margins. Nature, 473: 7479.CrossRefGoogle ScholarPubMed
Hulen, J. B., 2001. Characterization and conceptual modeling of magmatically-heated and deep-circulation, high-temperature hydrothermal systems in the Basin and Range and Cordilleran USA, In: Creed, B. (Ed.), Geothermal Technologies Program, Geoscience and Supporting Technologies, 2001 University Research Summaries. Idaho National Engineering and Environmental Laboratory, Idaho Falls, 19.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.Google Scholar
Hulen, J. B., Norton, D., Kaspereit, D., et al., 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: 227240.Google Scholar
Humayon, M., Lillie, R. J., and Lawrence, R. D., 1991. Structural interpretation of the eastern Sulaiman fold belt and foredeep, Pakistan. Tectonics, 10: 299324.CrossRefGoogle Scholar
Humayon, M., Akram, M., and Malik, R. M., 1998. Dhodak Field: a case history of first and the largest condensate discovery of Pakistan. PAPG-PPEPCA Pakistan Petroleum Convention Proceedings, pp. 134–174.Google Scholar
Humphreys, E., Clayton, R. W., and Hager, B. H., 1984. A tomographic image of mantle structure beneath Southern California. Geophysical Research Letters, 11(7): 625627.CrossRefGoogle Scholar
Hunt, C. B., and Mabey, D. R., 1966. General geology of Death Valley, California: stratigraphy and structure. USGS Professional Paper 0494-A.CrossRefGoogle Scholar
Hunt, J. M., 1979. Petroleum Geochemistry and Geology. W. H. Freeman, San Francisco.Google Scholar
Hunt, J. M., 1990. Generation and migration of petroleum from abnormally pressured fluid compartments. AAPG Bulletin, 74: 112.Google Scholar
Hunt, J. M., 1991. Generation of gas and oil from coal and other terrestrial organic matter. Organic Geochemistry, 17: 673680.CrossRefGoogle Scholar
Hurai, V., 2013. Hydrothermal overprint of sedimentary rocks: mineralogy, Raman spectroscopy, electron probe microanalysis and fluid inclusion microthermometry of well cores from north-west Atlantic continental margin. In: Kotulová, J., and Nemčok, M. (Eds.), NW African Margins Project. EGI Technical Report I 01013. EGI, Salt Lake City, pp. 334393.Google Scholar
Hurai, V., Marko, F., Tokarski, A. K., et al., 2006. Fluid inclusion evidence for deep burial of the Tertiary accretionary wedge of the Carpathians. Terra Nova, 18: 440446.CrossRefGoogle Scholar
Hurford, A. J., and Green, P. F., 1983. The zeta age calibration of fission-track dating. Chemical Geology, 41(4): 285317.CrossRefGoogle Scholar
Hurst, N., McDermott, K., and Bellingham, P., 2019. Insights into the development of the Guyana Transform Margin and its future play potential from newly acquired long-offset seismic reflection data. Abstract. P67 in Abstract Volume. HGS PESGB 18th African Conference, October 1–2, 2019, London, Olympia.Google Scholar
Hurwitz, S., Garfunkel, Z., Ben-Gai, Y., et al., 2002. The tectonic framework of a complex pull-apart basin: seismic reflection observations in the Sea of Galilee, Dead Sea transform. Tectonophysics, 359: 289306.CrossRefGoogle Scholar
Hutchison, I., 1985. The effects of sedimentation and compaction on oceanic heat flow. Geophysical Journal of the Royal Astronomic Society, 82: 439459.CrossRefGoogle Scholar
Hutnak, M., and Fisher, A. T., 2007. Influence of sedimentation, local and regional hydrothermal circulation, and thermal rebound on measurements of seafloor heat flux. Journal of Geophysical Research, 112: B121010. DOI: 10.1029/2007JB005022CrossRefGoogle Scholar
Hutnak, M., Fisher, A. T., Zuehlsdorf, L., et al., 2006. Hydrothermal recharge and discharge guided by basement outcrops on 0.7–3.6 Ma seafloor east of the Juan de Fuca Ridge: observations and numerical models. Geochemistry, Geophysics, Geosystems, 7: Q07O02. DOI: 10.1029/2006GC001242CrossRefGoogle 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: 711714.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(1): 4467.Google Scholar
Ibrmajer, J. and Suk, M., 1989. Geophysical Image of CSSR [in Czech]. ÚÚG, Prague.Google Scholar
Illies, J. H., 1981. Mechanism of graben formation. Tectonophysics, 73: 249266.CrossRefGoogle Scholar
Imperato, D. P., 1994. Structure, tectonics, and sedimentation, Elk Hills and vicinity, southwestern San Joaquin Valley, California. Report. Archive of OXY at Elk Hills.Google Scholar
INA, 2000a. Exploration area Sava, Blocks 1–4. Bid round data packages. INA-Industrija Nafte d.d. NAFTAPLIN, Zagreb.Google Scholar
INA, 2000b. Exploration area Drava, Blocks 1–7. Bid round data packages. INA-Industrija Nafte d.d. NAFTAPLIN, Zagreb.Google Scholar
ION/GX Technology, 2007. IndiaSpan East, Phase 1. Interpretation report. Archive of Reliance, Mumbai.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: 713716.2.0.CO;2>CrossRefGoogle Scholar
Irwin, W. P., and Barnes, I., 1980. Tectonic relations of carbon dioxide discharges and earthquakes. Journal of Geophysical Research, 85: 31153121.CrossRefGoogle Scholar
Issler, D. R., McQueen, H., and Beaumont, C., 1989. Thermal and isostatic consequences of simple shear extension of the continental lithosphere. Earth and Planetary Science Letters, 91(3–4): 341358.CrossRefGoogle Scholar
Itoh, Y., and Nagasaki, Y., 1996. Crustal shortening of Southwest Japan in the Late Miocene. Island Arc 5: 337353.CrossRefGoogle Scholar
Itoh, Y., Tsutsumi, H., Yamamoto, H., and Arato, H., 2002. Active right-lateral strike-slip fault zone along the southern margin of the Japan Sea. Tectonophysics, 351: 301314.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. 6575.Google Scholar
Iyer, K., Rupke, L. H., and Morgan, J. P., 2010. Feedbacks between mantle hydration and hydrothermal convection at ocean spreading centers. Earth and Planetary Science Letters, 296: 3444.CrossRefGoogle Scholar
Jackson, J., and McKenzie, D., 1988. The relationship between plate motions and seismic moment tensors, and the rates of active deformation in the Mediterranean and Middle East. Geophysical Journal International, 93: 4573.CrossRefGoogle Scholar
Jadoon, I. A. K., Lawrence, R. D., and Lille, R. J., 1993. Evolution of foreland structures: an example from the Sulaiman thrust lobe of Pakistan, southwest of the Himalayas. In: Treloar, P. J., and Searle, M. P., (Eds.), Himalayan Tectonics. Geological Society of London, London, pp. 589602.Google Scholar
Jagoutz, O., Muntener, O., Manatschal, G., et al., 2007. The rift-to-drift transition in the North Atlantic: a stuttering start of the MORB machine? Geology, 35: 10871090.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(9): 973976.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(7): 13571378.CrossRefGoogle Scholar
Jardim de Sá, E. F., 1984. A evolução Proterozoica da Província Borborema, paper presented at 11st Simp. de. Geol. do Nordeste, Soc. Bras. De Geologica, Núcleo Nordeste, Natal, Brazil.Google Scholar
Jarrad, R. D., 1986. Relations among subduction parameters. Reviews in Geophysics, 24: 217284.CrossRefGoogle Scholar
Jarriage, J.-J., Ott dÉstevou, P., Burollet, P. F., et al., 1990. The multistage tectonic evolution of the Gulf of Suez and northern Red Sea continental rift from field observations. Tectonics, 9(3): 441465.CrossRefGoogle Scholar
Jarvie, D. M., 2012. Extended abstract: shale resource systems for oil and gas. In: Breyer, J. A. (Ed.), Shale Reservoirs: Giant Resources for the 21st Century. AAPG, Washington, DC, pp. 13.Google Scholar
Jaswal, T., Lillie, R., and Lawrence, R, 1997. Structure and evolution of the northern Potwar deformed zone, Pakistan. AAPG Bulletin, 81: 308329.Google Scholar
Jaupart, C., and Mareschal, J.-C., 2011. Heat Generation and Transport in the Earth, Cambridge University Press, Cambridge.Google Scholar
Jenkyns, H. C., 1980. Cretaceous anoxic events: from continents to oceans. Journal of the Geological Society (London, U.K.), 137: 171188.CrossRefGoogle Scholar
Jenkyns, H. C., 1994. Carbon-isotope stratigraphy and paleooceanographic significance of the Lower Cretaceous shallow water carbonates of Resolution Guyot, Mid-Pacific Mountains. In: Proceedings of the Ocean Drilling Program, Scientific Results, Volume 143. Texas A&M University, College Station, pp. 99104Google Scholar
Jerolmack, D. J., and Paola, C., 2010, Shredding of environmental signals by sediment transport. Geophysical Research Letters, 37: L19401. DOI: 10.1029/2010GL044638CrossRefGoogle Scholar
Jessop, A. M., and Majorowicz, J. A., 1994. Fluid flow and heat transfer in sedimentary basins. In: Parnell, J. (Ed.), Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins. Geological Society of London, London, pp. 4354.Google Scholar
Jiang, X., Jin, Y., and McNutt, M. K., 2004. Lithospheric deformation beneath the Altyn Tagh and West Kunlun faults from recent gravity surveys. Journal of Geophysical Research: Solid Earth, 109(B5). DOI: 10.1029/2003JB002444CrossRefGoogle Scholar
Jimenez-Munt, I., and Sabadini, R., 2002. The block-like behavior of Anatolia envisaged in the modeled and geodetic strain rates. Geophysical Research Letters, 29(20). DOI: 10.1029/2002GL015995CrossRefGoogle Scholar
Jin, Z., Cao, J., Hu, W., et al., 2008. Episodic petroleum fluid migration in fault zones of the northwestern Junggar Basin (northwest China): evidence from hydrocarbon-bearing zoned calcite cement. AAPG Bulletin, 92(9): 12251243.CrossRefGoogle Scholar
Jiracek, G. R., Gonzalez, V. M., Caldwell, G., Wannamaker, P. E., and Kilb, D., 2007. Seismogenic, electrically conductive, and fluid zones at continental plate boundary in New Zealand, Himalaya, and California, U.S.A. In: Okaya, D., Stern, T., and Davey, F. (Eds.), A Continental Plate Boundary: Tectonics at South Island, New Zealand. AGU, Washington, DC, pp. 347369.CrossRefGoogle Scholar
Jiříček, R., 1979. Tectonogenetic evolution of the Carpathian arc during Oligocene and Neogene [in Czech]. In: Maheľ, M. (Ed.), Tectonic profiles of West Carpathians [in Slovak]. GÚDŠ, Bratislava, pp. 205215.Google Scholar
Jiříček, R., and Seifert, P., 1990. Paleogeography of the Neogene in Vienna Basin and adjacent part of the foredeep. In: Minaříková, D., and Lobitzer, H. (Eds.), Thirty Years of Geological Cooperation Between Austria and Czechoslovakia. Ústřední Úřad Geologický, Praha, pp. 89104.Google Scholar
Jiříček, R., and Tomek, Č., 1981. Sedimentary and structural evolution of the Vienna Basin. Earth Evolution Sciences, 3–4: 195204.Google Scholar
Johnson, G. D., Raynolds, R. G., and Burbank, D. W., 1986. Late Cenozoic tectonics and sedimentation in the north-western Himalayan foredeep: I. Thrust ramping and associated deformation in the Potwar region. In: Allen, P., and Homewood, P. (Eds.), Foreland Basins. International Association of Sedimentologists, Gent, pp. 273291.CrossRefGoogle Scholar
Johnson, K. M., Bürgmann, R., Freymueller, J. T., 2009. Coupled afterslip and visco-elastic flow following the 2002 Denali Fault, Alaska earthquake. International Geophysics Journal, 176: 670-682.CrossRefGoogle Scholar
Johnson, K. M., Shelly, D. R., and Bradley, A. M., 2013. Simulations of tremor-related creep reveal a weak crustal root of the San Andreas Fault. Geophysical Research Letters, 40(7): 13001305.CrossRefGoogle Scholar
Johnson, N. M., Officer, C. B., Opdyke, N. D., et al., 1983. Rates of late Cenozoic tectonism in the Vallecito-Fish Creek basin, western Imperial Valley, California. Geology, 11: 664667.2.0.CO;2>CrossRefGoogle Scholar
Johnson, P., 1991. Magnetic surveys of Iceland. Tectonophysics, 189: 246263.Google 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
Jones, D. S., Snoke, A. W., Premo, W.R., and Chamberlain, K. R., 2010. New models for Paleoproterozoic orogenesis in the Cheyenne belt region: evidence from the geology and U–Pb geochronology of the Big Creek Gneiss, southeastern Wyoming. GSA Bulletin, 122 (11–12), 18771898.CrossRefGoogle 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: 18051821.CrossRefGoogle Scholar
Jonk, R., Hurst, A., Duranti, D., et al., 2005. Origin and timing of sand injection, petroleum migration, and diagenesis in Tertiary reservoirs, south Viking Graben, North Sea. AAPG Bulletin, 89(3): 329357.CrossRefGoogle Scholar
Jordan, T. E., 1995. Retroarc foreland and related basins. In: Busby, C. J., and Ingersoll, R. V. (Eds.), Tectonics of Sedimentary Basins. Blackwell Science, Cambridge, pp. 331362.Google 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: 66816699.CrossRefGoogle Scholar
Jowett, E. C., Cathles, L. M. III., and Davis, B. W., 1993. Predicting depths of gypsum dehydration in evaporitic sedimentary basins. AAPG Bulletin, 77: 402413.Google Scholar
Junger, A., 1976. Tectonics of the Southern California Borderland, Aspects of the Geological History of the California Continental Borderland. AAPG, Washington, DC, pp. 486498.Google Scholar
Kadko, D., Rosenberg, N. D., Lupton, J. E., Collier, R. W., and Lilley, M. D., 1990. Chemical reaction rates and entrainment within the Endeavour Ridge hydrothermal plume. Earth and Planetary Science Letters, 120: 361374.CrossRefGoogle Scholar
Kahle, H.-G., Straub, C., Reilinger, R., et al., 1998. The strain rate field in the eastern Mediterranean region, estimated by repeated GPS measurements. Tectonophysics, 294(3): 237252.CrossRefGoogle Scholar
Kamesh Raju, K. A., Ramprasad, T., Rao, P. S., Rao, B. R., and Varghese, J., 2004. New insights into the tectonic evolution of the Andaman basin, northeast Indian Ocean. Earth Planetary Science Letters, 221: 145162.CrossRefGoogle Scholar
Kanamori, H., and Stewart, G. S., 1976. Mode of the strain release along the Gibbs fracture zone, Mid-Atlantic Ridge. Physics of the Earth and Planetary Interiors, 11: 312332.CrossRefGoogle Scholar
Kanungo, S., Thusu, B., Platon, E., and Apaalse, L., 2012. High Resolution Mesozoic Chronostratigraphy and Depositional Facies of the Offshore Ghana Margin (Gulf of Guinea). EGI Report I00979. EGI Archive, Salt Lake City.Google Scholar
Kappelmeyer, O., and Haenel, R., 1974. Geothermics With Special Reference to Application. Gebrueder Borntraeger, Berlin.Google Scholar
Karalliyadda, S. C., and Savage, M. K., 2013. Seismic anisotropy and lithospheric deformation of the plate-boundary zone in South Island, New Zealand: inferences from local S-wave splitting. Geophysical Journal, 193(2): 507530.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(6): 1112.Google Scholar
Karner, G. D., 2000. Rifts of the Campos and Santos basins, southeastern Brazil; distribution and timing. In: Mello, M. R., and Katz, B. J. (Eds.), Petroleum Systems of South Atlantic Margins. AAPG, Washington, DC, pp. 301315.Google Scholar
Karner, G. D., and Dewey, J. F., 1986. Rifting: lithospheric versus crustal extension as applied to the Ridge Basin of southern California. AAPG Memoir, A131: 317337.Google 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 Special Publications, 153: 1140.CrossRefGoogle Scholar
Karnkowski, P., 1999. Oil and Gas Deposits in Poland. The Geosynoptics Society “GEOS,” University of Mining and Metallurgy, Cracow.Google Scholar
Karrech, A., Regenauer-Lieb, K., and Poulet, T., 2011. Continuum damage mechanics for the lithosphere. Journal of Geophysical Research: Solid Earth, 116(B4). DOI: 10.1029/2010JB007501CrossRefGoogle Scholar
Karson, J. A., 1991. Seafloor spreading on the Mid-Atlantic Ridge: implications for the structure of ophiolites and oceanic lithosphere produced in slow-spreading environments. In: Malpas, J., Moores, E. M., Panayiotou, A., and Xenophontos, C. (Eds.), Proceedings of Symposium “Troodos 1987.” Geological Survey Department, Nicosia, pp. 547555.Google Scholar
Karson, J. A., 2017. The Iceland plate boundary zone: propagating rifts, migrating transforms, and rift-parallel strike-slip faults. Geochemistry, Geophysics, Geosystems, 18: 40434054.CrossRefGoogle Scholar
Karson, J. A., Tivey, M. A., and Delaney, J. R., 2002. Internal structure of uppermost oceanic crust along the Western Blanco Transform Scarp: implications for subaxial accretion and deformation at the Juan de Fuca Ridge. Journal of Geophysical; research, 107: 2181. DOI: 10.1029/2000JB000051CrossRefGoogle Scholar
Katz, B. J., and Mello, M. R., 2000. Petroleum systems of South Atlantic marginal basins: an overview. In: Mello, M. R., and Katz, B. J. (Eds.), Petroleum Systems of South Atlantic Margins. AAPG, Washington, DC, pp. 113.Google Scholar
Kaus, B. J. P., and Podladchikov, Y. Y., 2006. Initiation of localized shear zones in viscoelastoplastic rocks. Journal of Geophysical Research: Solid Earth, 111(B4). DOI: 10.1029/2005JB003652CrossRefGoogle Scholar
Kaymakci, N., Inceöz, M., Ertepinar, P., and Koç, A., 2010. Late Cretaceous to Recent kinematics of SE Anatolia (Turkey). In: Sosson, M., Kaymakci, N., Stephenson, R. A., Bergerat, F., and Starostenko, V. (Eds.), Sedimentary Basin Tectonics from the Black Sea and Caucasus to the Arabian Platform. Geological Society of London, London, pp 409435.Google Scholar
Keen, C. E., 1985. Evolution of rifted continental margins. Paper from the Geological Survey of Canada 85-8.Google Scholar
Keen, C. E., Kay, W. A., and Roest, W. R., 1990. Crustal anatomy of a transform continental margin. Tectonophysics, 173: 527544.CrossRefGoogle Scholar
Keen, C. E., Kay, W. A., Keppie, D., et al., 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 = Journal Canadien des Sciences de la Terre, 28(7): 10961111.CrossRefGoogle Scholar
Keith, J. F., Jr., Vass, D., Kanes, W. H., et al., 1989. Sedimentary Basins of Slovakia, Part II: Final Report on the Hydrocarbon Potential of the Danube Lowland Basin. ESRI, University of South Carolina, Columbia.Google Scholar
Keith, J. F., Jr. Vass, D., and Kováč, M., 1995. The Danube Lowland Basin. ESRI Occasional Publication, 11A: Slovakian Geology, pp. 63–88.Google Scholar
Kelley, D. S., and Gillis, K. M., 2002. Petrologic constraints upon hydrothermal circulation. Paper presented at InterRidge Theoretical Institute: Thermal Structure of Ocean Crust and Dynamics of Hydrothermal Circulation, Pavia, Italy.Google Scholar
Kelley, D. S., Karson, J. A., Früh-Green, G. L., et al., 2005. A serpentinite-hosted ecosystem: the Lost City hydrothermal field. Science, 307: 14281434,CrossRefGoogle ScholarPubMed
Kempe, D. R. C., 1974. The petrology of the basalts, Leg 26. In: Initial Reports of the Deep Sea Drilling Project, suppl. Durban, South Africa to Freemantle, Australia; Sept–Oct. 1972, Volume 26. Texas A&M University, College Station, pp. 465503.Google Scholar
Kennedy, B. M., Kharaka, Y. K., Evans, W. C., et al., 1997. Mantle fluids in the San Andreas fault system, California. Science, 278: 12781281.CrossRefGoogle Scholar
Keppie, J. D., 1982. Tectonic map of Nova Scotia. Abstracts with Programs – Geological Society of America, 14(1–2): 30.Google 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
Ketcham, R., Donelick, R. A., and Carlson, W., 1999. Variability of apatite fission-track annealing kinetics: III. Extrapolation to geological timescales. American Mineralogist, 84: 12351255.CrossRefGoogle Scholar
Ketcham, R., Donelick, R. A., and Donelick, M., 2000. AFTsolve: a program for multikinetic modeling of apatite fission track data. Geological Material Research, 2: 132.Google Scholar
Key, R. M., and Watkins, R. T., 1988. Geology of the Sabarei area; degree sheets 3 and 4 with coloured 1:250,000 geological map and results of geochemical exploration.Google Scholar
Khalil, S. M., 1998. Tectonic and structural development of the eastern Gulf of Suez margin. PhD Thesis. Royal Holloway University of London, London.Google Scholar
Khan, M. A., Ahmed, R., Raza, H., and Arif, K., 1986. Geology of petroleum in Kohat-Potwar Depression, Pakistan. AAPG Bulletin, 70: 396414.Google Scholar
Khan, M. R., Bhatti, M. A., Baitu, A. H., and Sarwar, M. Z., 2009. Effect of Mega-Shear Fractures / Strike Slip Faults on Entrapment Mechanism in Sulaiman Fold Belt, Pakistan. Search and Discovery Article #30229.Google Scholar
Khan, S. H., Lawrence, R. D., and Nakata, T., 1991. Chaman fault, Pakistan, Afghanistan. Report of the Geological Survey of Pakistan. Geological Survey of Pakistan, Islamabad.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(9): 10491052.CrossRefGoogle 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(9): 10371040.CrossRefGoogle 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: Haneberg, W.C.M., Moore, P. S., and Goodwin, J. C. (Eds.), Faults and Subsurface Fluid Flow in the Shallow Crust. AGU, Washington, DC, pp. 129148.CrossRefGoogle Scholar
Khodayar, M., and Einarsson, P., 2002. Strike-slip faulting, normal faulting, and lateral dike injections along a single fault: field example of the Gljúfurá fault near a Tertiary oblique rift-transform zone, Borgarfjodrur, west Iceland. Journal of Geophysical Research, 107(B5).CrossRefGoogle Scholar
Khodayar, M., Einarsson, P., Franzson, H., and Bjornsson, S., 2008. Tectonic settings of low temperature geothermal activity in Iceland: relation to plate boundaries, earthquakes, and rift jumps. International Geological Congress, Abstracts, 33: 1343956.Google Scholar
Khodayar, M., Bjornsson, S., Guðnason, E. A., et al., 2018. Tectonic control of the Reykjanes geothermal field in the oblique rift of SW Iceland: from regional to reservoir scales. Open Journal of Geology, 8: 333382.CrossRefGoogle Scholar
Khripounoff, A., Vangriesheim, A., Babonneau, N., et al., 2003, Direct observation of intense turbidity current activity in the Zaire submarine valley at 4000 m water depth: Marine Geology, 194: 151158.CrossRefGoogle Scholar
Kilembe, E. A., and Rosendahl, B. R., 1992. Structure and stratigraphy of the Rukwa rift. Tectonophysics, 209: 143158.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: 24452457.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, April 2–7, 2006, Vienna.Google Scholar
Klemperer, S. L., Kennedy, B. M., Sastry, S. R., et al., 2013. Mantle fluids in the Karakoram fault: helium isotope evidence. Earth and Planetary Science Letters, 366: 5970.CrossRefGoogle Scholar
Klitgord, K. D., and Schouten, H., 1986. Plate Kinematics of the Central Atlantic: The Geology of North America. Geological Society of America, Boulder.CrossRefGoogle Scholar
Klitgord, K. D., Hutchinson, D. R., and Schouten, H., 1988. U.S. Atlantic continental margin: structural and tectonic framework. In: Sheridan, R. E., and Grow, J. A. (Eds.), The Atlantic Continental Margin. The Geology of North America. GSA, Boulder, pp. 1955.CrossRefGoogle Scholar
Kluft, I., 2020. Aerial photo of the San Andreas Fault in the Carrizo Plain. Available at: www.usgs.gov/media/images/san-andreas-faultGoogle Scholar
Knipe, R. J., 1986. Faulting mechanisms in slope sediments: Examples from deep sea drilling project cores. In: Moore, J. C. (Ed.), Structural Fabric in Deep Sea Drilling Project Cores from Forearcs. Geological Society of America, Boulder, pp. 4554.CrossRefGoogle Scholar
Knipe, R. J., 1989. Deformation mechanisms: recognition from natural tectonites. Journal of Structural Geology, 11: 127146.CrossRefGoogle Scholar
Knipe, R. J., 1992. Faulting processes and fault seal. In: Larsen, R. M., Brekke, H., Larsen, B. T., and Talleraas, E. (Eds.), Structural and Tectonic Modeling and Its Application to Petroleum Geology. Norwegian Petroleum Society, Oslo, pp. 325342.CrossRefGoogle Scholar
Knipe, R. J., 1993. The influence of fault zone processes and diagenesis on fluid flow. In: Horbury, A. D., and Robinson, A. (Eds.), Diagenesis and Basin Development. AAPG, Washington, DC, pp. 135151.Google Scholar
Knipe, R. J., 1997. Juxtaposition and seal diagrams to help analyze fault seals in hydrocarbon reservoirs. AAPG Bulletin, 81: 187195.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 complexes. Philosophical Transactions of the Royal Society of London 335: 261273.Google Scholar
Kocák, A., Mayer, S., and Pšeničková, M., 1981. Structural Scheme of the Vienna Basin on Ore-Neogene Base [in Czech]. Geofond, Prague.Google Scholar
Koçyiğit, A., 1989. Suşehri basin: an active fault-wedge basin on the North Anatolian Fault Zone, Turkey. Tectonophysics, 167(1): 1329.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. In: Proceedings of the World Geothermal Congress 2000, Kyushu-Tohoku, May 28–June 10, pp. 26712676.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
Kohli, A. H., Goldsby, D. L., Hirth, J. G., and Tullis, T. E., 2011. Flash weakening of serpentinite at near-seismic slip rates. Journal of Geophysical Research, 116. DOI: 10.1029/2010JB007833CrossRefGoogle Scholar
Kohlstedt, D. L., Evans, B., and Mackwell, S. J., 1995. Strength of the lithosphere: constraints imposed by laboratory experiments. Journal of Geophysical Research, 100: 1758717602.CrossRefGoogle Scholar
Kolandaivelu, K. P., Harris, R. N., Lowell, R. P., et al., 2017. Analysis of a conductive heat flow profile in the Ecuador Fracture Zone. Earth and Planetary Science Letters, 467: 120127.CrossRefGoogle Scholar
Kooi, H., and Beaumont, C., 1994. Escarpment retreat on high-elevation rifted continental margins: insights derived from a surface processes model that combines diffusion, advection and reaction. Journal of Geophysical Research, 99(B6): 1219112209.CrossRefGoogle Scholar
Kooijman, E., Mezger, K., and Berndt, J., 2010. Constraints on the U–Pb systematics of metamorphic rutile from in situ LA-ICP-MS analysis. Earth and Planetary Science Letters, 293(3–4): 321330.CrossRefGoogle Scholar
Koons, P. O., 1987. Some thermal and mechanical consequences of rapid uplift: an example from the Southern Alps, New Zealand. Earth and Planetary Science Letters, 86(2–4): 307319.CrossRefGoogle Scholar
Koons, P. O., Norris, R. J., Craw, D., and Cooper, A. F., 2003. Influence of exhumation on the structural evolution of transpressional plate boundaries: an example from the Southern Alps, New Zealand. Geology, 31(1): 36.2.0.CO;2>CrossRefGoogle Scholar
Kopietz, J., and Jung, R., 1978. Geothermal in situ experiments in the Asse salt-mine. In: Seminar on in Situ Heating Experiments in Geological Formations. OECD Publications, Paris.Google Scholar
Korenaga, J., 2007. Thermal cracking and the deep hydration of oceanic lithosphere: a key to the generation of plate tectonics? Journal of Geophysical Research, 112(B5): B05408.CrossRefGoogle Scholar
Kosmos Energy, 2017. Competent Persons Report. Available at: https://sec.report/CIK/0001509991.Google Scholar
Kotulová, J., and Rybár, S., 2013. Offshore Equatorial Guinea: Evaluation of Block-H Aleta Sub-Basin Hydrocarbon Potential. EGI Report I 01157. Archive EGI, Salt Lake City.Google Scholar
Koutsoukos, E. A. M., 1992. Late Aptian to Maastrichtian foraminiferal biogeography and palaeooceanography of the Sergipe basin, Brazil. Palaeogeography, Palaeoclimatology, Palaeoecology, 92: 295324.CrossRefGoogle Scholar
Koutsoukos, E. A. M., Mello, M. R., Azambuja, , et al., 1991. The Albian succession of Sergipe basin, Brazil: an integrated paleoenvironmental assessment. AAPG Bulletin, 75: 479498.Google Scholar
Kováč, M., 2000. Geodynamic, Paleogeographic and Structural Evolution of the Carpathian–Pannonian Region During Miocene: a New Insight at Neogene Basins of Slovakia. Comenius University, Bratislava.Google Scholar
Kováč, M., Baráth, I., Holický, I., Marko, F., and Túnyi, I., 1988. Stratigraphic and paleogeographic correlation of the evolution of Eggenburgian sediments in the SE part of the Malé Karpaty Mts., Trnava plain and Považie [in Slovak]. Geological Institute of the Slovak Academy of Sciences, Bratislava.Google Scholar
Kováč, M., Baráth, I., Kováčová-Slamková, M., et al., 1989. Late Miocene paleoenvironments and sequence stratigraphy: northern Vienna Basin. Geologica Carpathica, 49(6): 445458.Google Scholar
Kováč, M., Baráth, I., Šútovská, K., and Uher, P., 1991. Lower Miocene events in the sedimentary record of the Dobrá Voda Depression. Mineralia Slovaca, 23: 201213.Google Scholar
Kováč, M., Marko, F., and Nemčok, M., 1993. Neogene structural evolution and basin opening in the Western Carpathians. Geophysical Transactions, 37(4): 297309.Google Scholar
Kováč, M., Hudáčková, N., Rudinec, R., and Lankreijer, A., 1996. Basin evolution in the foreland and hinterland of the Carpathian accretionary prism during the Neogene: evidence from the Western to Eastern Carpathians Junction. Annales Tectonicae, 10(1–2): 319.Google Scholar
Kováč, M., Baráth, I., Harzhauser, M., Hlavatý, I., and Hudáčková, N., 2004. Miocene depositional systems and sequence stratigraphy of the Vienna Basin. Cour. Forsch.-Inst. Senckenberg, 246: 187212.Google Scholar
Kranz, R. L., 1983. Microcracks in rocks: a review. Tectonophysics, 100: 449480.CrossRefGoogle 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: 48934898.CrossRefGoogle Scholar
Krasnov, S. G., Cherkasov, G. A., Stepanova, T. V., et al., 1995. Detailed geological studies of hydrothermal fields in the North Atlantic. In: Parson, L. M., Walker, C. L., and Dixon, D. R. (Eds.), Hydrothermal Vents and Processes. Geological Society of London, London, pp. 4364.Google Scholar
Kreutzer, N., 1986. Die Ablagerungssequenzen der miozanen Badener Serie in Feld Matzen und in zentralen Wiener Becken. Erdol-Erdgas-Kohle, 102: 492503.Google Scholar
Kreutzer, N., 1992. Matzen field – Austria, Vienna Basin. AAPG Treatise-Atlas, Structural Traps, 7: 5793.Google Scholar
Kreutzer, N., 1993a. The Neogene of the Vienna Basin [in German]. In: Brix, F., and Schultz, O. (Eds), Oil and Gas in Austria. Naturhistorischen Museum, Wien und F. Berger, Horn.Google Scholar
Kreutzer, N., 1993b. The reservoirs of the Vienna Basin and underlying units [in German]. In: Brix, F., and Schultz, O. (Eds), Oil and Gas in Austria. Naturhistorischen Museum, Wien und F. Berger, Horn.Google Scholar
Kreutzer, N., and Hlavatý, V., 1990. Sediments of the Miocene (mainly Badenian) in the Matzen area in Austria and in the southern part of the Vienna Basin in Czechoslovakia. In: Minaříková, D., and Lobitzer, H. (Eds.), Thirty Years of Geological Cooperation Between Austria and Czechoslovakia. Ústřední Úřad Geologický, Praha, pp. 110123.Google Scholar
Kreuzer, H., Harre, W., and Kuersten, M., 1977. K/Ar dates of two glauconites from the Chandrapur series (Chhattishgarh India): on the stratigraphic status of the late Precambrian basin in central India. Geologisches Jahrbuch, 28: 2336.Google Scholar
Kriner, K. A., Pockalny, R. A., and Larson, R. L., 2006. Bathymetric gradients of lineated abyssal hills: inferring seafloor spreading vectors and a new model for hills formed at ultra-fast rates. Earth and Planetary Science Letters, 242: 98110.CrossRefGoogle Scholar
Kröll, A. W., and Wessely, G., 1993. Strukturkarte Basis der tertiaren Beckenfullung. Geologische Themenkarte der Republik Osterreich. Wiener Becken und angrenzende Gebiete 1:200, 000. Geological Survey of Austria, Vienna.Google Scholar
Krooss, B. M., Brothers, L., and Engel, M. H., 1991. Geochromatography in petroleum migration: a review. In: England, W. A. and Fleet, A. J. (Eds.), Petroleum Migration. Geological Society of London, London, pp. 149163.Google Scholar
Krumbein, W. C., and Monk, G. D., 1943. Permeability as a function of the size parameters of unconsolidated sand. Transactions of AIME, 151: 153.CrossRefGoogle 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. Geothermal regime, petroleum generation and migration. In: Evans, D., Graham, C., Armour, A., and Bathurst, P. (Eds.), The Millennium Atlas; petroleum geology of the central and northern North Sea. Geological Society of London, London, pp. 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
Kukal, Z., 1990. The Rate of Geological Processes. Elsevier, Amsterdam.Google Scholar
Kulm, L. D., and Suess, E., 1990. Relationship between carbonate deposits and fluid venting: oregon accretionary prism. Special section on the Role of fluids in sediment accretion, deformation, diagenesis, and metamorphism in subduction zones. Journal of Geophysical Research, 95: 88998915.CrossRefGoogle Scholar
Kuo, B.-Y., and Forsyth, D. W., 1988. Gravity anomalies of the ridge-transform system in the South Atlantic between 31 and 34.5°S: upwelling centers and variations in crustal thickness. Marine Geophysics Research, 10: 205232.CrossRefGoogle Scholar
Labails, C., Olivet, J., Aslanian, D., and Roest, W., 2010. An alternative early opening scenario for the Central Atlantic Ocean. Earth and Planetary Science Letters, 297: 355368.CrossRefGoogle Scholar
Lachenbruch, A. H., 1968. Preliminary geothermal model of the Sierra Nevada. Journal of Geophysical Research, 73: 69776989.CrossRefGoogle Scholar
Lachenbruch, A. H., and Sass, J. H., 1977. Heat flow in the United States and the thermal regime of the crust. In: Heacock, J. G. (Eds.), The Earth’s Crust, Its Nature and Physical Properties. AGU, Washington, DC, pp. 626675.Google Scholar
Lachenbruch, A. H., and Sass, J. H., 1980. Heat flow and energetics of the San Andreas Fault Zone. Journal of Geophysical Research: Solid Earth, 85(B11): 61856222.CrossRefGoogle Scholar
Lachenbruch, A. H., and Sass, J. H., 1988. The stress heat-flow paradox and thermal results from Cajon Pass. Geophysical Research Letters, 15(9): 981984.CrossRefGoogle Scholar
Ladwein, H. W., 1988. Organic geochemistry of Vienna basin: model for hydrocarbon generation in overthrust belts. AAPG Bulletin, 72: 586599.Google Scholar
LaFemina, P. C., Dixon, T. H., Malsevisi, R., et al., 2005. Geodetic GPS measurements in south Iceland: strain accumulation and partitioning in a propagating ridge system. Journal of Geophysical Research, 111: B11405.Google Scholar
Lal, R. K., Ackermand, D., and Upadhyay, H., 1987. P–T–X relationships deduced from corona textures in sapphirine–spinel–quartz assemblages from Paderu, S. India. Journal of Petrology, 28: 11391168.CrossRefGoogle Scholar
Lallemand, S., and Chemenda, A., 1999. Strain partitioning and coupling between plates: field observations and critical input from experimental modeling. Paper presented at Penrose Conference on Subduction to Strike-Slip Transitions on Plate Boundaries. GSA, Puerta Plata, Dominican Republic.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. Wiley, New York.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: 361375.CrossRefGoogle 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. Zbl. Geol. Palaont. Teil I, 10–12: 13751389.Google Scholar
Lana, M. C., 1993, Potencial petrolífero e exploração em agues profundas na Bacia de Sergipe- Alagoas: Relatório Interno Petrobrás, DEPEX/DINORD/SESEA.Google Scholar
Landuyt, W., and Bercovici, D., 2009. Formation and structure of lithospheric shear zones with damage. Physics of the Earth and Planetary Interiors, 175(3–4): 115126.CrossRefGoogle Scholar
Laney, S. E., and Gates, A. E., 1996. Three-dimensional shuffling of horses in a strike-slip duplex: an example from the Lambertville Sill, New Jersey. Tectonophysics, 258: 5370.CrossRefGoogle Scholar
Langmuir, C. H., and Bender, J. F., 1984. The geochemistry of oceanic basalts in the vicinity of transform faults: observations and implications. Earth and Planetary Science Letters, 69: 107127.CrossRefGoogle Scholar
Langseth, M. G., Mottl, M. J., Hobart, M. A., and Fisher, A. T., 1988. The distribution of geothermal and geochemical gradients near Site 501/504: Implications for hydrothermal circulation in the oceanic crust. In: Proceedings of the Ocean Drilling Program, Initial Reports, Ocean Drilling Program, Volume 111. Texas A&M University, College Station, pp. 2332.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.Google Scholar
Lankreijer, A. C., Kováč, M., Cloetingh, S. A. P. L., et al., 1995. Quantitative subsidence analysis and forward modeling of the Vienna and Danube basins: thin-skinned versus thick-skinned extension. Tectonophysics, 252: 433451.CrossRefGoogle Scholar
Lankreijer, A. C., Bielik, M., Cloetingh, S. A. P. L., and Majcin, D., 1998. Rheology predictions across Western Carpathians, Bohemian Massif and Pannonian basin: implications for tectonic scenarios. Tectonics, 18(6): 11391153.CrossRefGoogle Scholar
Lantz, R. J., 1968. A review of the Elk Hills oil field, Kern County, California. In: Guidebook, Geology and Oilfields, West Side Southern San Joaquin Valley, 43rd Annual Meeting of Pacific Sections of AAPG, SEG, SEPM. AAPG, SEG, SEPM, Tulsa, pp. 4954.Google 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(2–3): 7985.CrossRefGoogle 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
Laubscher, H. P., 1972. Some overall aspects of Jura dynamics. American Journal of Science, 272: 293304.CrossRefGoogle Scholar
Lavier, L., and Manatschal, G. A., 2006. A mechanism to thin the continental lithosphere at magma-poor margins. Nature 440: 324328.CrossRefGoogle ScholarPubMed
Laville, E., and Petit, J.-P., 1984. Role of synsedimentary strike-slip faults in the formation of Moroccan Triassic basins. Geology, 12(7): 424427.2.0.CO;2>CrossRefGoogle Scholar
Lawrence, S. R., and Cornford, C., 1995. Basin geofluids. Basin Research, 7: 17.CrossRefGoogle Scholar
Lawrence, S., and Coster, P., 1985. Petroleum potential of offshore Guyana. Oil and Gas Journal 83(49): 6774.Google Scholar
Lawver, L., and Williams, D., 1979. Heat flow in the central Gulf of California. Journal of Geophysical Research, 84: 34653478.CrossRefGoogle Scholar
Lay, T., 2019. Reactivation of oceanic fracture zones in large intraplate earthquakes? In: Duarte, J. C. (Ed.), Transform Plate Boundaries and Fracture Zones, Elsevier, Amsterdam, pp. 89104.CrossRefGoogle Scholar
Lay, T. and Wallace, T. C. 1995. Modern Global Seismology. Academic Press, San Diego.Google Scholar
Le Pichon, X., and Hayes, D. E., 1971. Marginal offsets, fracture zones, and the early opening of the South Atlantic. Journal of Geophysical Research, 76(26): 62836293.CrossRefGoogle Scholar
Le Pichon, X., Francheteau, J., and Bonin, J., 1973. Plate Tectonics: Developments in Geotectonics. Elsevier, Amsterdam.Google Scholar
Le Pichon, X., Chamot-Rooke, N., Lallemant, S., Noomen, R., and Veis, G., 1995. Geodetic determination of the kinematics of central Greece with respect to Europe: implications for eastern Mediterranean tectonics, Journal of Geophysical Research, 100: 1267512690.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: 359381.CrossRefGoogle Scholar
Lee, E. Y., and Wagreich, M., 2017. Polyphase tectonic subsidence evolution of the Vienna Basin inferred from quantitative subsidence analysis of the northern and central parts. International Journal of Earth Sciences, 106: 687705.CrossRefGoogle Scholar
Lee, M., and Clechenko, C., 2018. Oil below water: Perched water and high order sealing elements, implications for exploration in stratigraphic traps. Adapted from poster presentation given at 2018 AAPG Annual Convention & Exhibition, Salt Lake City, Utah, May 20–23, 2018.Google Scholar
Lee, M., Cox, P., Cheshire, O., Pluim, S., and Nyamaah, L., 2014. Reservoir characterisation of the Paradise and Hickory discoveries, offshore Ghana: integration of depositional and diagenetic concepts. AAPG Search and Discovery: 50985.Google 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: Nemec, W. and Steel, R. J. (Eds.), Fan Deltas; Sedimentology and Tectonic Settings. Blackie and Son, Glasgow.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 A 83: 339346.Google Scholar
Leitner, B., Eberhart-Phillips, D., Anderson, H., and Nabelek, J. L., 2001. A focused look at the Alpine fault, New Zealand: seismicity, focal mechanisms, and stress observations. Journal of Geophysical Research: Solid Earth, 106(B2): 21932220.CrossRefGoogle 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: 568579.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 the Ocean Drilling Program: Scientific Results, Volume 159. Texas A&M University, College Station, pp. 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: Engel, M. H., and Macko, S. A. (Eds.), Organic Geochemistry Plenum Press, New York, pp 419442.CrossRefGoogle Scholar
Lewan, M. D., 1994. Assessing natural oil expulsion from source rocks by laboratory pyrolysis. In: Magoon, L. B., and Dow, W. G. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 201210.Google Scholar
Lewan, M. D., 1997. Experiments on the role of water in petroleum formation. Geochimica et Cosmochimica Acta, 61: 36913723.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: Dyman, T. S. and Kuuskraa, V. A. (Eds.), Geologic Studies of Deep Natural Gas Resources. USGS, Reston, 19.Google Scholar
Lewis, C. R., and Rose, S. C., 1970. A theory relating high temperatures and overpressures. Journal of Petroleum Technology, 22: 1116.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: England, W. A. and Fleet, A. J. (Eds.), Petroleum Migration. Geological Society of London, London, pp. 3346.Google Scholar
Leythaeuser, D., Littke, 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(1–3): 489502.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(3): 701713.CrossRefGoogle Scholar
Li, X., Faure, M., Lin, W., and Manatschal, G., 2013. New isotopic constraints on age and magma genesis of an embryonic oceanic crust: the Chenaillet Ophiolite in the Western Alps. Lithos (Oslo) 160–161: 283291.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(1): 210221.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.-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(4B): S349S363.CrossRefGoogle Scholar
Libak, A., Mjelde, R., Keers, H., Faleide, J. I., and Murai, Y., 2012. An integrated geophysical study of Vestbakken Volcanic Province, western Barents Sea continental margin, and adjacent oceanic crust. Marine Geophysical Research, 33: 185207.CrossRefGoogle Scholar
Lidmar-Bergstrom, K., and Naslund, J. O., 2002. Landforms and uplift in Scandinavia. In: Dore, A. G., Cartwright, J. A., Stoker, M. S., Turner, J. P., and White, N. (Eds.), Exhumation of the North Atlantic Margins: Timing, Mechanisms, and Implications for Petroleum Exploration. Geological Society of London, London, pp. 103116.Google Scholar
Ligi, M., Bonatti, E., Gasperini, L., and Poliakov, A. N. B., 2002. Oceanic broad multifault transform plate boundaries. Geology, 30: 1114.2.0.CO;2>CrossRefGoogle Scholar
Ligi, M., Bonatti, E., Cipriani, A., and Ottolini, L., 2005. Water-rich basalts at mid-ocean-ridge cold spots. Nature, 434: 6669.CrossRefGoogle ScholarPubMed
Lillie, R. J., 1999. Whole Earth Geophysics: An Introductory Textbook for Geologists and Geophysicists. Prentice Hall, Upper Saddle River.Google Scholar
Lillis, , 1994. Soda Lake-Painted Rock(!) petroleum system in the Cuyama Basin, California, U.S.A. In: Magoon, L. B., and Dow, W. E. (Eds.), The Petroleum System; from Source to Trap. AAPG, Washington, DC, pp. 437451.Google Scholar
Lima, C. C., Bentz, C. M., Fonseca, L. E. N., Lima Neto, F. F., and Gusso, G. L. N., 1993. Correlações entre a direção do campo de tensões neotectônicas, a topografia e estruturas geológicas na Bacia Potiguar. Internal report. Petrobras.Google Scholar
Lima Neto, F. F., 1993. Geologia da Bacia Potiguar e de suas acumulaces de petroles. Internal Report. Petrobras.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
Lin, J., and Phipps Morgan, J., 1992. The spreading rate dependence of three-dimensional mid-ocean ridge gravity signature. Geophysics Research Letters, 19: 1316.CrossRefGoogle Scholar
Lin, J., Purdy, G. M., Schouten, H., Sempere, J.-C., and Zervas, C., 1990. Evidence for focused magmatic accretion along the Mid-Atlantic Ridge. Nature, 344: 627632.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: 109122.CrossRefGoogle 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(1–2): 179191.CrossRefGoogle Scholar
Lisker, F., and Fachmann, S., 2001. Phanerozoic history of the Mahanadi region, India. Journal of Geophysical Research, 106(B10): 2202722050.CrossRefGoogle Scholar
Lister, C. R. B., 1977. Estimators for heat flow and deep rock properties based on boundary layer theory. Tectonophysics, 41: 157171.CrossRefGoogle Scholar
Lister, G. S., Etheridge, M. A., and Symonds, P. A., 1986. Detachment faulting and the evolution of passive continental margins. Geology (Boulder), 14(3): 246250.2.0.CO;2>CrossRefGoogle 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: 549559.CrossRefGoogle Scholar
Little, T. A., Holcombe, R. J., and Ilg, B. R., 2002a. Kinematics of oblique collision and ramping inferred from microstructures and strain in middle crustal rocks, central Southern Alps, New Zealand. Journal of Structural Geology, 24(1): 219239.CrossRefGoogle Scholar
Little, T. A., Savage, M. K., and Tikoff, B., 2002b. Relationship between crustal finite strain and seismic anisotropy in the mantle, Pacific–Australia plate boundary zone, South Island, New Zealand. Geophysical Journal, 151(1): 106116.CrossRefGoogle Scholar
Lobontiu, N., 2018. System Dynamics for Engineering Students, Concepts and Applications, 2nd edition, Academic Press, New York.Google Scholar
Lockner, D. A., Moore, D. E., and Reches, Z., 1992. Microcrack interaction leading to shear fracture. In: 33rd Symposium on Rock Mechanics, pp. 807–816.Google Scholar
Lodolo, E., Coren, F., and Ben-Avraham, Z. 2013. How do long-offset oceanic transforms adapt to plate motion changes? The example of the Western Pacific–Antarctic plate boundary. Journal of Geophysical Research, 118(3): 11951202.CrossRefGoogle Scholar
Logan, J. M., 1992. The influence of fluid flow on the mechanical behaviour of faults. In: Tillerson, J. R., and Wawersik, W. R. (Eds.), Rock Mechanics, Balkema, Rotterdam, pp. 141149.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. In: Proceedings of VIII Conference on Analysis of Actual Fault Zones in Bedrock. USGS, Menlo Park, pp. 305343.Google Scholar
Loncke, L., Maillard, A., Basile, C., et al., 2016. Structure of the Demerara passive transform margin and associated sedimentary processes: initial results from the IGUANES cruise. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London.Google Scholar
Lonsdale, P., 1985. Non-transform offsets of the Pacific–Cocos plate boundary and their trace on the rise flank. Geological Society of America Bulletin, 96: 313327.2.0.CO;2>CrossRefGoogle Scholar
Lonsdale, P., 1989. Geology and tectonic history of the Gulf of California. In: Winterer, E. L., Hussong, D. M., and Decker, R. W. (Eds.), The Eastern Pacific Ocean and Hawaii. Geological Society of America, Denver, pp. 499522.CrossRefGoogle Scholar
Lonsdale, P., 1991. Structural patterns of the pacific floor offshore of peninsular California. In: Dauphin, J. P., and Simoneit, B. (Eds.), The Gulf and Peninsular Province of the Californias. AAPG, Washington, DC, pp. 87125.Google Scholar
Lonsdale, P., and Becker, K., 1985. Hydrothermal plumes, hot springs, and conductive heat flow in the southern trough of Guaymas Basin. Earth and Planetary Science Letters, 73: 211225.CrossRefGoogle 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
Lorenzo, J. M., and Vera, E. E., 1992. Thermal uplift across the continent–ocean boundary of the Southern Exmouth Plateau. Earth and Planetary Science Letters, 108: 7992.CrossRefGoogle Scholar
Lorenzo, J. M., Mutter, J. C., Larson, R. L., et al. 1991. Development of the continent–ocean transform boundary of the southern Exmouth Plateau. Geology, 19: 843846.2.3.CO;2>CrossRefGoogle 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. AAPG Bulletin, 83: 244276.Google Scholar
Losh, S., Walter, L., Meulbroek, P., et al., 2002. Reservoir fluids and their migration into the South Eugene Island Block 330 reservoirs, offshore Louisiana. AAPG Bulletin, 86(8): 14631488.Google Scholar
Louden, K. E., and Forsyth, D. W., 1976. Thermal conduction across fracture zones and the gravitational edge effect. Journal of Geophysical Research, 81(6): 48694874.CrossRefGoogle Scholar
Lowell, J. D., 1985. Structural Styles in Petroleum Exploration. OGCI Publications, Tulsa.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: 351365.CrossRefGoogle 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: 565575.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: 87698783.CrossRefGoogle Scholar
Ludwig, K. R., 2001. Users manual for Isoplot/Ex rev. 2.49. Spec. Publ. 1a. Berkeley Geochronology Center, Berkeley.Google Scholar
Lundin, E. R., and Doré, A. G., 2017. Non-Wilsonian break-up pre-disposed by transforms: examples from the North Atlantic and Arctic. In: Wilson, W., McCaffrey, K., Doré, A. G., Houseman, G., and Royden, L. (Eds.), Tectonic Evolution: 50 years of the Wilson Cycle Concept. Geological Society of London, London.Google Scholar
Lundin, E. R., Doré, A. G., Rønning, K., and Kyrkjebø, R., 2013. Repeated inversion and collapse in the Late Cretaceous–Cenozoic northern Vøring Basin, offshore Norway. Petroleum Geoscience, 19(4): 329341.CrossRefGoogle Scholar
Lundin, E. R., Doré, A. G., and Redfield, T. F., 2018. Magmatism and extension rates at rifted margins. Petroleum Geoscience 24(4). DOI: 10.1144/petgeo2016-158.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
Lupton, J. E., 1995. Hydrothermal plumes: near and far field. In: Humphries, S., Zierenberg, R., Mullineaux, L., and Thompson, R., (Eds.), Physical, Chemical, Biological and Geological Interactions Within Sea Floor Hydrothermal Systems. AGU, Washington, DC, pp. 317346.Google Scholar
Luyendyk, B. P., Kamerling, M. J., and Terres, R. R., 1980. Geometric model for Neogene crustal rotations in southern California. GSA Bulletin, 91: 211217.2.0.CO;2>CrossRefGoogle Scholar
Luyendyk, B. P., Kamerling, M. J., Terres, R. R., and Hornafius, J. S., 1985. Simple shear of southern California during Neogene time suggested by paleomagnetic declinations. Journal of Geophysical Research, 90: 1245412466.CrossRefGoogle Scholar
Lybéris, N., 1988. Tectonic evolution of the Gulf of Suez and the Gulf of Aqaba. Tectonophysics, 153: 209220.CrossRefGoogle Scholar
Lybéris, N., and Manby, G., 1993. The origin of the West Spitsbergen Fold Belt from geological constraints and plate kinematics: implications for the Arctic. Tectonophysics, 224: 371391.CrossRefGoogle Scholar
Macdonald, A. J., and Spooner, E. T. C., 1981. Calibrations of a LINKAM TH600 programmable heating–cooling stage for microthermometric examination of fluid inclusions. Economic Geology, 76: 12481258.CrossRefGoogle Scholar
Macdonald, K. C., and Fox, P. J., 1983. Overlapping spreading centers: new accretion geometry on the East Pacific Rise. Nature, 303: 5557.CrossRefGoogle Scholar
Macdonald, K. C., Fox, P. J., Perram, L. J., et al., 1988. A new view of the mid-ocean ridge from the behavior of ridge-axis discontinuities. Nature, 335: 217225.CrossRefGoogle Scholar
Macdonald, K. C., Scheirer, D. S., and Carbotte, S. M., 1991. Mid-oceanic ridges: discontinuities, segments and giant cracks. Science, 253: 986994.CrossRefGoogle Scholar
Macedo, J., and Marshak, S., 1999. Controls on the geometry of fold-thrust belt salients. Geological Society of America Bulletin, 111(12): 18081822.2.3.CO;2>CrossRefGoogle Scholar
Macgregor, D. S., 1993. Relationships between seepage, tectonics and subsurface petroleum reserves. Marine and Petroleum Geology, 10: 606619.CrossRefGoogle Scholar
Mackenzie, A. S., and Quigley, T. M., 1988. Principles of geochemical prospect appraisal. AAPG Bulletin, 72: 399415.Google 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(5): 11451154.CrossRefGoogle Scholar
MacLeod, C. J., Escartín, J., Banerji, D., et al., 2002. Direct geological evidence for oceanic detachment faulting: the Mid-Atlantic Ridge. Geology 30: 879882.2.0.CO;2>CrossRefGoogle Scholar
Magnavita, L. P., 1992. Geometry and kinematics of the Recôncavo-Tucano-Jatobá rift, NE Brazil. PhD Thesis. University of Oxford, England.Google 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. AAPG Bulletin, 89(9): 12031237.CrossRefGoogle Scholar
Magoon, L. B., Lillis, P. G., and Peters, K. E., 2007. Total petroleum systems used to determine assessment units in the San Joaquin Basin Province, California. In: Scheirer, A. H. (Ed.), Petroleum Systems and Geologic Assessment of Oil and Gas in the San Joaquin Basin Province, California. USGS, Reston, pp. 165.Google Scholar
Maheľ, M., (Ed.) 1973. Tectonic Map of the Carpathian–Balkan Mountain System and Adjacent Areas. Scale 1:1,000, 000. GÚDŠ, Bratislava.Google Scholar
Maher, L. J. 1998. Automating the dreary measurements for loss on ignition. INQUA Sub-Commission on Data-Handling Methods, Newsletter 18.Google Scholar
Maia, M., Sichel, S., Briais, A., et al., 2016. Extreme mantle uplift and exhumation along a transpressive transform fault. Nature Geoscience Letters, 9: 619623.CrossRefGoogle Scholar
Mailloux, B. J., Person, M., Kelley, S., et al., 1999. Tectonic controls on the hydrogeology of the Rio Grande rift, New Mexico. Water Resources Research, 35(9): 26412659.CrossRefGoogle Scholar
Mair, K., and Abe, S., 2011. Breaking up: comminution mechanisms in sheared simulated fault gouge. Pure and Applied Geophysics, 168:(12): 22772288.CrossRefGoogle Scholar
Malamud, B. D., and Turcotte, D. L., 1999. How many plumes are there? Earth and Planetary Science Letters, 174(1–2): 113124.CrossRefGoogle Scholar
Maltman, A. J., 1994. Prelithification deformation. In: Hancock, P. L. (Eds.), Continental Deformation. Pergamon Press, Tarrytown, pp. 143158.Google 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(3): 432466.Google Scholar
Mancktelow, N. S., and Grasemann, B., 1997. Time-dependent effects of heat advection and topography on cooling histories during erosion. Tectonophysics, 270: 167195.CrossRefGoogle Scholar
Mandic, O., Harzhauser, M., Spezzaferri, S., and Zuschin, M., 2003. The paleoenvironment of an early Middle Miocene Paratethys sequence in NE Austria with special emphasis on paleoecology of mollusks and foraminifera. Geobios, 35: 193206.CrossRefGoogle Scholar
Mandl, G., 1988. Mechanics of Tectonic Faulting: Models and Basic Concepts. Developments in Structural Geology. Elsevier, Amsterdam.Google Scholar
Mango, F. D., 1997. The light hydrocarbons in petroleum: a critical review. Organic Geochemistry, 26: 417440.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, U., 1994. An integrated approach to the study of primary petroleum migration. In: Parnell, J. (Ed.), Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins. Geological Society of London, London: pp. 233260.Google Scholar
Mann, U., Hantschel, T., Schaefer, R. G., et al., 1997. Petroleum migration: mechanisms, pathways, efficiencies and numerical simulations. In: Welte, D. H., Horsfield, B., and Baker, D. R. (Eds.), Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling. Springer, Berlin, pp. 405520.Google Scholar
Manspeizer, W., 1985. Dead Sea Rift: impact of climate and tectonism on Pleistocene and Holocene sedimentation. In: Biddle, K., and Christie-Blick, N., (Eds.), Strike-Slip Deformation, Basin Formation and Sedimentation. SEPM, Tulsa, pp. 143158.CrossRefGoogle Scholar
Marinho, M., Mascle, J., and Wannesson, J., 1988. Structural framework of the southern Guinean Margin (Central Atlantic). Journal of African Earth Sciences, 7(2): 401408.CrossRefGoogle Scholar
Markwick, P. J., Raddadi, M. C., Bailiff, R. G., et al., 2010. The evolution of global source-to-sink relationships during the Cretaceous and Tertiary using stage level palaeogeographies and DEMs. AAPG Search and Discovery 90104.Google Scholar
Marsh, A. J., Pachan, D. S., and Sims, D. W., 1990. Strike-slip duplexing: an analog model for transpressional plate margins. Geological Society of America – Abstracts with Programs 22: 143A.Google Scholar
Marshak, S., and Wilkerson, M. S., 1992. Effect of overburden thickness on thrust belt geometry and development. Tectonics, 11: 560566.CrossRefGoogle Scholar
Mart, Y., and Dauteuil, O., 2000. Analogue experiments of propagation of oblique rifts. Tectonophysics, 316(1–2): 121132.CrossRefGoogle Scholar
Mart, Y., Ryan, W. B. F., and Lunina, O. V., 2005. Review of the tectonics of the Levant Rift system: the structural significance of oblique continental breakup. Tectonophysics, 395(3–4): 209232.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: 14511465.2.3.CO;2>CrossRefGoogle Scholar
Martin, J. B., Orange, D., Lorenson, T. D., and Kvenvolden, K. A., 1997. Chemical and isotopic evidence of gas-influenced flow at a transform plate boundary: Monterey Bay, California. Journal of Geophysical Research, 102: 2032520345.CrossRefGoogle 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 Papers of the Geological Society of America, 380: 243–65.Google Scholar
Márton, E., 1993. Paleomagnetism in the Mediterranean from Spain to the Aegean: a review of data relevant to Cenozoic movements. In: Mantovani, E., Morelli, A., and Boschi, E. (Eds.), Recent Evolution and Seismicity of the Mediterranean Region. Kluwer, Dordrecht, pp. 367402.CrossRefGoogle Scholar
Masaryk, P., and Lintnerová, O., 1997. Diagenesis and porosity of the Upper Triassic carbonates of the pre-Neogene Vienna Basin basement. Geologica Carpathica, 48(6): 371386.Google Scholar
Mascle, J., 1976. Atlantic-type continental margins: distinction of two basic structural types. Anais da Academia Brasileira de Ciencas, 48: 191197.Google Scholar
Mascle, J., and Blarez, E., 1987. Evidence for transform margin evolution from the Ivory Coast Ghana continental margin. Nature, 326 (6111): 378381.CrossRefGoogle Scholar
Mascle, J., Blarez, E., and Marinho, M., 1988. The shallow structure of the Guinea and Cote d’Ivoire transform margins: their bearing on the equatorial Atlantic Mesozoic evolution. Tectonophysics, 155: 193209.Google Scholar
Mascle, J., Lohmann, P., Clift, P., and ODP 159 Scientific Party, 1997. Development of a passive transform margin: Côte d’Ivoire–Ghana transform margin – ODP Leg 159 preliminary results. Geo-Marine Letters, 17: 411.CrossRefGoogle Scholar
Massari, F., and Colella, A., 1988. Evolution and types of fan-delta systems in some major tectonic settings. In: Nemec, W. and Steel, R. J. (Eds.), Fan Deltas; Sedimentology and Tectonic Settings. Blackie and Son, London, pp. 103122.Google Scholar
Massini, E., Manatschal, G., and Mohn, G., 2013. The Alpine Tethys rifted margins: reconciling old and new ideas to understand the stratigraphic architecture of magma-poor rifted margins. Sedimentology, 60: 174196.CrossRefGoogle Scholar
Matenco, L., Bertotti, G., Cloetingh, S., and Dinu, C., 2003. Subsidence analysis and tectonic evolution of the external Carpathian–Moesian Platform region during Neogene times, Sedimentary Geology 156: 7194.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: Beaumont, E. A., and Foster, N. H. (Eds.), Exploring for Oil and Gas Traps. AAPG, Washington, DC, pp. 7-31–7-38.Google Scholar
Mattinson, J. M., 2005. Zircon U–Pb chemical abrasion (“CA-TIMS”) method: combined annealing and multi-step partial dissolution analysis for improved precision and accuracy of zircon ages. Chemical Geology, 220: 4766.CrossRefGoogle Scholar
Mauduit, T., and Dauteuil, O., 1996. Small-scale models of oceanic transform faults. Journal of Geophysical Research, 101: 2019520209.CrossRefGoogle Scholar
Maurin, J. C., 1985. Analyse des zones decrochantes dans le fosse de la Benoue (Nigeria) et systematiques U/Pb et Pb/Pb appliques aux mineralisations uraniferes et de Pb/Zn associees. PhD thesis.University Montpellier.Google Scholar
Maurin, J. C., and Guiraud, R., 1990. Relationships between tectonics and sedimentation in the Barremo–Aptian intracontinental basins of Northern Cameroon. Journal of African Earth Sciences, 10(1–2): 331340.CrossRefGoogle Scholar
Maurin, J. C., Benkhelil, J., and Robineau, B., 1986. Fault rocks of the Kaltungo lineament, NE Nigeria, and their relationships with Benue Trough tectonics. Journal of the Geological Society of London, 143: 587599.CrossRefGoogle Scholar
Mavko, G. M., and Nur, A., 1979. Wave attenuation in partially saturated rocks. Geophysics, 44: 161178.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(10): 13571372.2.3.CO;2>CrossRefGoogle Scholar
McAlister, E. J., Cann, J., and Spencer, S., 1995. The evolution of crustal deformation in an oceanic extensional environment. Journal of Structural Geology, 17: 183199.CrossRefGoogle Scholar
McAuliffe, C. D., 1979. Oil and gas migration: chemical and physical constraints. AAPG Bulletin, 73: 14551471.Google 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: 16681686.Google Scholar
McBride, J. H., Barazangi, M., Best, J., et al., 1990. Seismic reflection structure of intracratonic palmyride fold-thrust belt and surrounding Arabian platform, Syria. AAPG Bulletin, 74(3): 238259.Google Scholar
McCaffrey, R., 1992. Oblique plate convergence, slip vectors, and forearc deformation. Journal of Geophysical Research, 97: 89058915.CrossRefGoogle Scholar
McCaffrey, R., 1996. Slip partitioning at convergent plate boundaries of SE Asia. In: Hall, R., and Blundell, D. J. (Eds.), Tectonic Evolution of Southeast Asia. Geological Society of London, London, pp. 318.Google Scholar
McCaffrey, R., 2009. The tectonic framework of the Sumatran subduction zone. Annual Review of Earth and Planetary Sciences, 37: 345366.CrossRefGoogle Scholar
McCaffrey, R., Bock, R. Y., and Rais, J., 1990. Crustal deformation and oblique plate convergence in Sumatra. Eos Transactions AGU, 71: 637.Google Scholar
McCaffrey, R., Zwick, P. C., Bock, Y., et al., 2000. Strain partitioning during oblique plate convergence in northern Sumatra: geodetic and seismologic constraints and numerical modeling. Journal of Geophysical Research, 105(B12): 2836328376.CrossRefGoogle Scholar
McClay, K., and Dooley, T., 1995. Analogue models of pull-apart basins. Geology, 23(8): 711714.2.3.CO;2>CrossRefGoogle Scholar
McClay, K., and Khalil, S., 1998. Extensional hard linkages, eastern Gulf of Suez, Egypt. Geology (Boulder), 26(6): 563566.2.3.CO;2>CrossRefGoogle Scholar
McClay, K. R., Dooley, T., Whitehouse, P., and Mills, M., 2002. 4-D evolution of rift systems: insights from scaled physical models. AAPG Bulletin, 86: 935959.Google Scholar
McCluskey, S., Balassanian, S., Barka, A., et al., 2000. Global Positioning System constraints on plate kinematics and dynamics in the eastern Mediterranean and Caucasus. Journal of Geophysical Research: Solid Earth, 105(B3): 56955719.CrossRefGoogle Scholar
McCoss, A., 2017. Opening new oil basins: a pattern of discovery. AAPG Search and Discovery 70280.Google Scholar
McDougall, I., and Harrison, T. M., 1988. Geochronology and thermochronology by the 40Ar/ 39Ar method. Oxford Monographs on Geology and Geophysics, 9: 212.Google Scholar
McGuire, J. J., and Beroza, G. C., 2012. A rogue earthquake off Sumatra. Science, 336: 11181119.CrossRefGoogle ScholarPubMed
McGuire, J. J., Boettcher, M. S., and Jordan, T. H., 2005. Foreshock sequences and short-term earthquake predictability on East Pacific Rise transform faults. Nature, 434: 457461.CrossRefGoogle ScholarPubMed
McGuire, J. J., Collins, J. A., Gouedard, P., et al., 2012. Variations in earthquake rupture properties along the Gofar transform fault, East Pacific Rise. Nature Geoscience, 5: 336341.CrossRefGoogle 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(6): 937947.CrossRefGoogle Scholar
McKenzie, D. P., 1969. Speculations on the consequences and causes of plate motions. Geophysical Journal of the Royal Astronomic Society, 18: 132.CrossRefGoogle Scholar
McKenzie, D. P., 1972. Active tectonics of the Mediterranean region. Royal Astronomical Society Geophysical Journal, 30: 109185.CrossRefGoogle Scholar
McKenzie, D. P., 1984. A possible mechanism for epeirogenetic uplift: Nature, 307: 616619.CrossRefGoogle Scholar
McKenzie, D., Jackson, J., and Priestley, K., 2005. Thermal structure of oceanic and continental lithosphere. Earth Planetary Science Letters, 233: 337349.CrossRefGoogle Scholar
McKibben, M. A., and Hardie, L. A., 1997. Ore-forming brines in active continental rifts. In: Barnes, H. L. (Ed.), Geochemistry of Hydrothermal Ore Deposits. Wiley, New York, pp. 877935.Google 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. AAPG Bulletin, 86: 10611093.Google Scholar
McMaster, R. L., Schilling, J.-G. E., and Pinet, P. R., 1977. Plate boundary within the Tjornes Fracture Zone on Iceland’s northern insular margin. Nature 269: 663668.CrossRefGoogle Scholar
Mechie, J., and Kind, R., 2013. A model of the crust and mantle structure down to 700 km depth beneath the Lhasa to Golmud transect across the Tibetan plateau as derived from seismological data. Tectonophysics, 606: 187197.CrossRefGoogle Scholar
Mechie, J., Abu-Ayyash, K., Ben-Avraham, Z., et al., 2005. Crustal shear velocity structure across the Dead Sea Transform from two-dimensional modelling of DESERT project explosion seismic data. Geophysical Journal, 160(3): 910924.CrossRefGoogle Scholar
Medwedeff, D. A., 1989. Growth fault-bend folding at southeast Lost Hills, San Joaquin Valley, California. AAPG Bulletin, 73: 5467.Google Scholar
Medwedeff, D. A., 1992. Geometry and kinematics of an active, laterally propagating wedge thrust, Wheeler Ridge, California. In: Mitra, S. and Fisher, G. W. (Eds.), Structural Geology of Fold and Thrust Belts. John Hopkins University Press, Baltimore, pp. 328.Google 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(3–4): 333344.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(3–4): 347363.CrossRefGoogle Scholar
Meijerink, A. M. J., Rao, D. P., and Rupke, J., 1984. Stratigraphic and structural development of the Precambrian Cuddapah basin, SE India. Precambrian Research, 26: 57104.CrossRefGoogle Scholar
Meisling, K. E., Cobbold, P. R., and Mount, V. S., 2001. Segmentation of an obliquely rifted margin, Campos and Santos basins, southeastern Brazil. AAPG Bulletin, 85(11): 19031924.Google Scholar
Meissner, R., 1986. The Continental Crust: A Geophysical Approach. Academic Press, San Diego.Google Scholar
Mello, M. R., Telnaes, N., Gaglianone, P. C., et al., 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., 1987. Evolução termomecânica da Bacia Potiguar – RN, Brasil. Master’s thesis, Departamento de Geologia da Escola de Minas, Universidade de Ouro Preto.Google Scholar
Mello, U. T., and Bender, A. A., 1988. On isostasy at the equatorial margin of Brazil. Revista Brasileira de Geociencias, 18: 237246.CrossRefGoogle Scholar
Menard, H. V., 1967. Extension of northeastern-Pacific fracture zones. Science, 155: 7274.CrossRefGoogle ScholarPubMed
Menard, H. V., and Atwater, T., 1969. Origin of fracture zone topography. Nature, 222: 10371040.CrossRefGoogle Scholar
Menard, H. V., and Chase, T. E., 1970. Fracture zones. In: Maxwell, A. E. (Ed.), The Sea, Vol. 4. Wiley-Interscience, New York, pp. 421444.Google Scholar
Mercier De Lépinay, M., Loncke, L., Basile, C., et al., 2016. Transform continental margins: Part 2, worldwide review. Tectonophysics, 693: 96115.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: 157179.CrossRefGoogle Scholar
Merrihue, C., and Turner, G., 1966. Potassium–argon dating by activation with fast neutrons. Journal of Geophysical Research, 71(11): 28522859.CrossRefGoogle Scholar
Meschede, M., and Frisch, W., 1998. A plate-tectonic model for the Mesozoic and Early Cenozoic history of the Caribbean plate. Tectonophysics, 296: 269291.CrossRefGoogle 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. Doctoral 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(11): 10721076.2.3.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(7): 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(1–2): 3163.CrossRefGoogle Scholar
Mezger, K., and Cosca, M. A., 1999. The thermal history of the Eastern Ghats Belt (India) as revealed by U–Pb and 40Ar/39Ar dating of metamorphic and magmatic minerals: implications for SWEAT correlation. Precambrian Research, 94: 251271.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), North-east Canadian Arctic Island. In: Biddle, K. T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation and Sedimentation. SEPM, Tulsa, pp. 361374.CrossRefGoogle Scholar
Michon, L., and Merle, O., 2000. Crustal structures of the Rhine graben and the Massif Central grabens: an experimental approach. Tectonics, 19: 896904.CrossRefGoogle Scholar
Middleton, M. F., 1982. Tectonic history from vitrinite reflectance. Geophysical Journal of the Royal Astronomical Society, 68(1): 121132.CrossRefGoogle Scholar
Mihut, D., and Müller, R. D., 1998. Volcanic margin formation and Mesozoic rift propagators in the Cuvier Abyssal Plain off Western Australia. Journal of Geophysical Research, 103: 2713527149.CrossRefGoogle 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: 876888.CrossRefGoogle Scholar
Milkov, A. V., 2010. Methanogenic biodegradation of petroleum in the West Siberian Basin (Russia): significance for formation of giant Cenomanian gas pool. AAPG Bulletin, 94: 14851541.CrossRefGoogle Scholar
Miller, D. J., and Christensen, N., 1997. Seismic velocities of lower crustal and upper mantle rocks from the slow-spreading Mid-Atlantic Ridge, south of the Kane transform zone (MARK). In: Proceedings of the Ocean Drilling Program, Scientific Results, Volume 153. Texas A&M University, College Station, pp. 437454.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(1): 3549.CrossRefGoogle 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, Texas A&M University, College Station, pp. 119127.Google Scholar
Min, K., Mundil, R., Renne, P. R., and Ludwig, K. R., 2000. A test for systematic errors in 40Ar/39Ar geochronology through comparison with U/Pb analysis of 1.1-Ga rhyolite. Geochimica et Cosmochimica Acta, 64(1): 7398.CrossRefGoogle Scholar
Minakov, A., Faleide, J. I., Glebovsky, V. Y., and Mjelde, R., 2013. Structure and evolution of the northern Barents-Kara Sea continental margin from integrated analysis of potential fields, bathymetry and sparse seismic data. Geophysical Journal International, 188: 79102.CrossRefGoogle Scholar
Mining Awareness, 2016. California nuclear power stations and the earthquake rupture forecast. Available at: https://miningawareness.wordpress.com/2016/06/12/california-nuclear-power-stations-and-the-earthquake-rupture-forecast/image-3365/Google Scholar
Mitra, S., 1986. Duplex structures and imbricate thrust systems: Geometry, structural position, and hydrocarbon potential. AAPG Bulletin, 70: 10871112.Google Scholar
Mitra, S., 1987. Regional variations in deformation mechanisms and structural styles in the central Appalachian orogenic belt. Geological Society of America Bulletin, 98: 569590.2.0.CO;2>CrossRefGoogle Scholar
Mitra, S., 1988a. Three dimensional geometry and kinematic evolution of the Pine Mountain thrust system, southern Appalachians. GSA Bulletin, 100: 7295.2.3.CO;2>CrossRefGoogle Scholar
Mitra, S., 1988b. Effects of deformation mechanisms on reservoir potential in Central Appalachian overthrust belt. AAPG Bulletin, 72: 536554.Google 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
Moghal, M. A., Saqi, M. I., Hameed, A., and Bugti, M. N., 2007. Subsurface geometry of Potwar sub-basin in relation to structuration and entrapment. Pakistan Journal of Hydrocarbon Research, 17: 6172.Google Scholar
Mohn, G., Manatschal, G., Beltrando, M., Masini, E., and Kusznir, N., 2012. Necking of continental crust in magma-poor rifted margins: evidence from the fossil Alpine Tethys margins. Tectonics, 31. DOI: 10.1029/2011TC002961CrossRefGoogle Scholar
Mohriak, W. U., Rabelo, J. H. L., Matos, R. D., and Barros, M. C., 1995. Deep seismic reflection profiling of sedimentary basins offshore Brazil: geological objectives and preliminary results in the Sergipe Basin. Journal of Geodynamics, 20: 515539.CrossRefGoogle 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(1–4): 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: Pankhurst, R. J., Trouw, R. A. J., de Brito Neves, B. B., and De Witt, M. J. (Eds.), West Gondwana: Pre-Cenozoic Correlations Across the South Atlantic Region. Geological Society of London, London, pp. 365398.Google Scholar
Mohsen, A., Hofstetter, R., Bock, G., et al., 2005. A receiver function study across the Dead Sea Transform. Geophysical Journal, 160(3): 948960.CrossRefGoogle Scholar
Molina, P., 1988. Correlation geologique Afrique-Amerique du Sud et provinces uraniferes. Journal of African Earth Sciences, 7: 489497.CrossRefGoogle Scholar
Molnar, P., and Dayern, K. E., 2010. Major intracontinental strike-slip faults and contrasts in lithospheric strength. Geosphere, 6: 444467.CrossRefGoogle Scholar
Molnar, P., and Tapponier, P., 1975. Cenozoic tectonics of Asia: effects of a continental collision. Science, 189(4201): 419426.CrossRefGoogle ScholarPubMed
Momper, J. A., 1978. Oil migration limitations suggested by geological and geochemical considerations. AAPG Bulletin, 8: 160.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
Montési, G. L. J., 2004a. Postseismic deformation and the strength of ductile shear zones. Earth, Planets and Space, 56: 11351142.CrossRefGoogle Scholar
Montési, G. L. J., 2004b. Controls of shear zone rheology and tectonic loading on postseismic creep. Journal of Geophysical Research: Solid Earth, 109(B10). DOI: 10.1029/2003JB002925CrossRefGoogle Scholar
Montési, G. L. J., 2013. Fabric development as the key for forming ductile shear zones and enabling plate tectonics. Journal of Structural Geology, 50: 254266.CrossRefGoogle Scholar
Montési, G. L. J., and Zuber, M. T., 2002. A unified description of localization for application to large-scale tectonics. Journal of Geophysical Research, Solid Earth, 107. DOI: 10.1029/2001JB000465Google Scholar
Montgomery, D. R., and Brandon, M. T., 2002. Topographic controls on erosion rates in tectonically active mountain ranges. Earth and Planetary Science Letters 201: 481489.CrossRefGoogle Scholar
Moore, D. E., and Rymer, M. J., 2007. Talc-bearing serpentinite and the creeping section of the San Andreas fault. Nature, 448: 795797.CrossRefGoogle ScholarPubMed
Moore, D. E., Lockner, D. A., Shengli, M., Summers, R., and Byerlee, J. D., 1997. Strengths of serpentinite gouges at elevated temperatures. Journal of Geophysical Research, 102(B7): 1478714801.CrossRefGoogle Scholar
Moore, D. G., 1969. Reflection profiling studies in the California continental borderland: structure and Quaternary turbidite basins. Geological Society of America Special Paper 1007.CrossRefGoogle Scholar
Moore, D. G., and Curray, J. R., 1982. Geologic and tectonic history of the Gulf of California. In: Initial Reports of the Deep Sea Drilling Project, Volume 64. Texas A&M University, College Station, pp. 12791294.Google Scholar
Moore, D. M., and Reynolds, R. C. Jr., 1997. X-ray Diffraction and the Identification and Analysis of Clay Minerals, 2nd edition. Oxford University Press, Oxford.Google Scholar
Moore, J. C., 1989. Tectonics and hydrogeology of accretionary prisms: role of the décollement zone. Journal of Structural Geology, 11: 95106.CrossRefGoogle Scholar
Moore, J. C., and Vrolijk, P., 1992. Fluids in accretionary prisms. Reviews of Geophysics, 30: 130135.Google Scholar
Moore, J. N., 2004. Fluid inclusion study of Niko-1 well. In: Nemčok, M. (Ed.), Georgia Niko Well Fracture Analysis Study. EGI Technical Report 01-0059-5000-50501264. EGI Archive, Salt Lake City.Google 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(5–6): 695710.CrossRefGoogle Scholar
Moore, M. E., Gleadow, A. J. W., and Lovering, J. L., 1986. Thermal evolution of rifted continental margins: new evidence from fission track analysis. Earth and Planetary Science Letters, 93: 255270.CrossRefGoogle Scholar
Mora, A., Reyes, A., Rodriguez, G., et al., 2013. Inversion tectonics under increasing rates of shortening and sedimentation throughout the Cenozoic. In: Nemčok, M., Mora, A., and Cosgrove, J. W. (Eds.), Thick-Skin-Dominated Orogens: From Initial Inversion to Full Accretion. Geological Society of London, London.Google 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: Clark, D. B. (Ed.), Dynamic Rock Mechanics: Proceedings of the 12th Symposium of Rock Mechanics. American Institute of Mining, Metallurgical and Petroleum Engineers, Englewood, pp. 221240.Google Scholar
Moresi, L., and Muhlhaus, H.-B., 2006. Anisotropic viscous models of large-deformation Mohr-Coulomb failure. Philosophical Magazine, 86(21): 32873305.CrossRefGoogle Scholar
Morgan, P., Boulos, F. K., Hennin, S. F., et al., 1985. Heat flow in eastern Egypt: the thermal signature of a continental breakup. Journal of Geodynamics, 4(1–4): 107131.CrossRefGoogle Scholar
Morgan, W. J., 1968. Rises, trenches, great faults, and crustal blocks. Journal of Geophysics, 73(6): 19591982.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: 511526.CrossRefGoogle Scholar
Morin, R. H., Hess, A. E., and Becker, K., 1992. In situ measurements of fluid flow in DSDP holes 395A and 534A: results from the dianaut program. Geophysical Research Letters, 19: 509512.CrossRefGoogle Scholar
Mørk, M. B. E., and Duncan, R. A., 1993. Late Pliocene basaltic volcanism on the western Barents Shelf margin: implications from petrology and 40Ar–39Ar dating of volcaniclastic debris from a shallow drill core. Norsk Geologisk Tidsskrift 73(4): 209225.Google 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): 479492.CrossRefGoogle Scholar
Morley, C. K., 2015. Cenozoic structural evolution of the Andaman Sea: evolution from an extensional to a sheared margin. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London, pp. 3961.Google Scholar
Morley, C. K., Nelson, E. 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. AAPG Bulletin, 74(8): 12341253.Google Scholar
Morley, C. K., Wescott, W. A., Stone, D. M., et al., 1992. Tectonic evolution of the northern Kenyan Rift. Journal of the Geological Society of London, 149(3): 333348.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: 12291263.CrossRefGoogle Scholar
Morrison, J., Burgess, C., Cornford, C., and N’Zalasse, N., 2000. Hydrocarbon systems of the Abidjan margin, Cote D’Ivoire. Paper presented at the Offshore West Africa Conference, Abidjan, March 21–23.Google Scholar
Morrow, C., and Byerlee, J., 1988. Permeability of rock samples from Cajon Pass, California. Geophysical Research Letters, 15(9): 10331036.CrossRefGoogle Scholar
Morrow, T. A., Kim, S. S., Mittelstaedt, E. L., Olive, J.-A., and Howell, S. M., 2016. New bathymetry reveals detailed history of transform fault segmentation at the Clarion fracture zone. Abstract T33C-3048, AGU Fall Meeting Abstracts.Google Scholar
Mottl, M. J., Sansone, F. J., Wheat, C. G., et al., 1995. Manganese and methane in hydrothermal plumes along the East Pacific Rise, 8° 40′ to 11° 50′N. Geochimica et Cosmochimica Acta, 59: 41474165.CrossRefGoogle Scholar
Moulin, M., Aslanian, D., Olivet, J.-L., et al. 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
Moulin, M., Aslanian, D., and Unternehr, P., 2010. A new starting point for the South and Equatorial Atlantic Ocean. Earth-Science Reviews, 98: 137.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. AAPG Bulletin, 86: 9791002.Google Scholar
Mpodozis, C., and Ramos, V., 1990. The Andes of Chile and Argentina. In: Ericksen, G. E., Cañas-Pinochet, M. T., and Reinemund, J. A. (Eds.), Geology of the Andes and Its Relation to Hydrocarbon and Mineral Resources. AAPG, Washington, DC, pp. 5991.Google Scholar
Mueller, K., and Suppe, J., 1997. Growth of Wheeler Ridge anticline, California: geomorphic evidence for fault-bend folding behavior during earthquakes. Journal of Structural Geology, 19: 383396.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. Available at: 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: 32113214.CrossRefGoogle Scholar
Müller, R. D., Sdrolias, M., Gaina, C., and Roest, W. R., 2008. Age, spreading rates and spreading symmetry of the world’s ocean crust, Geochemistry, Geophysics, Geosystems, 9: Q04006. DOI: 10.1029/2007GC001743CrossRefGoogle Scholar
Mulugeta, G., and Koyi, H., 1987. Three-dimensional geometry and kinematics of experimental piggyback thrusting. Geology (Boulder), 15: 10521056.2.0.CO;2>CrossRefGoogle Scholar
Murton, B. J., 1986. Anomalous oceanic lithosphere formed in a leaky transform fault: evidence from the Western Limassol Forest Complex, Cyprus. Journal of the Geological Society of London, 143: 845854.CrossRefGoogle Scholar
Murton, B. J., and Gass, I. G., 1986. Western Limassol Forest complex, Cyprus: Part of an Upper Cretaceous leaky transform fault. Geology, 14: 255258.2.0.CO;2>CrossRefGoogle Scholar
Murton, B. J., and Parson, L. M., 1993. Segmentation, volcanism and deformation of oblique spreading centers: a quantitative study of the Reykjanes Ridge. Tectonophysics, 222: 237257.CrossRefGoogle Scholar
Mutter, J. C., and Karson, J. A., 1992. Structural processes at slow spreading ridges. Science, 257: 627634.CrossRefGoogle ScholarPubMed
Mutti, E., and Normark, W. R., 1987, Comparing examples of modern and ancient turbidite systems: problems and concepts. In Leggett, J. K., and Zuffa, G. G. (Eds.), Marine Clastic Sedimentology: Concepts and Case Studies. Graham and Trotman, London, pp. 138.Google Scholar
Naar, D. F., and Hey, R. N., 1989. Speed limit for oceanic transform faults. Geology, 17: 420422.2.3.CO;2>CrossRefGoogle Scholar
Nabelek, J., Chen, W. P., and Ye, H., 1987. The Thangshan earthquake sequence and its implications for the evolution of the North China Basin. Journal of Geophysical Research, 92: 1261512628.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, 88: 790814.CrossRefGoogle Scholar
NATFA a.s., 2012a. Vienna Basin – Slovak part. CEEC Scout Group Meeting, Brno, October 5–6, 2012, extended abstract.Google Scholar
NATFA a.s., 2012b. Eastern Slovakia Basin. CEEC Scout Group Meeting, Brno, October 5–6, 2012, extended abstract.Google Scholar
Nagle, A. N., Pickle, R. C., Saal, A. E., Hauri, E. H., and Forsyth, D. W., 2007. Volatiles in basalts from intra-transform spreading centers: implications for melt migration models. Eos Transactions, 88(52).Google Scholar
Namson, J. S., and Davis, T. L., 1988. Seismically active fold and thrust belt in the San Joaquin Valley, central California. Geological Society of America Bulletin, 100: 257273.2.3.CO;2>CrossRefGoogle Scholar
Namson, J. S., Davis, T. L., andLagoe, M. B., 1990. Tectonic history and thrust-fold deformation style of seismically active structures near Coalinga. In: Rymer, M. J. and Ellsworth, W. L. (Eds.), The Coalinga, California, Earthquake of May 2, 1983. US Geological Survey Professional Paper 1487, pp. 79–96.Google Scholar
Nana, 2002. Presentation at Africa Upstream Conference, Cape Town, September.Google Scholar
Nanda, J. K., Rath, S. C., and Behera, S. N., 1998. Collision related post-orogenic magmatism in the contact zone between the high grade and low grade terranes: example from northwestern Orissa, India. In: Seminar Abstracts. Volume of the International Seminar on Precambrian Crust in Eastern and Central India. BBSR, pp. 6468.Google Scholar
Naranjo, J., 1987. Interpretacion de la actividad cenozoica superior a lo largo de la Zona de Fallas de Atacama, norte de Chile. Revista Geologica De Chile, 31: 4355.Google Scholar
Naylor, M. A., Mandl, G., and Sijpesteijn, C. H., 1986. Fault geometries in basement-induced wrench faulting under different initial stress states. Journal of Structural Geology, 7: 737752.CrossRefGoogle Scholar
Nazeer, A., Solangi, S. H., Brohi, I. A., et al., 2013. Hydrocarbon potential of Zinda Pir Anticline, Eastern Sulaiman Fold Belt, Middle Indus Basin, Pakistan. Pakistan Journal of Hydrocarbon Research, 22–23: 7384.Google Scholar
Nederlof, M. H., and Mohler, H. P., 1981. Quantitative investigation of trapping effect of unfaulted caprock. AAPG Bulletin, 65: 964.Google Scholar
Neev, D., Hall, J. K., 1979. Geophysical investigation in the Dead Sea. Sedimentary Geology, 23: 209238.CrossRefGoogle Scholar
Nelson, C. H., and Kulm, LD., 1973. Submarine Fans and Deep-Sea Channels. Turbidites and Deep-Water Sedimentation. SEPM, Tulsa, pp. 3978.Google Scholar
Nemčok, J., and Nemčok, M., 1990. Significance of movement vectors in the Vihorlat-Cirocha Fault System [in Slovak]. In: Sýkora, M., Jablonský, J., and Samuel, O. (Eds.), Sedimentologické problémy Západných Karpát. Geologický ústav Dionýza Štúra, Bratislava, pp. 89106.Google Scholar
Nemčok, M., 1993. Transition from convergence to escape: field evidence from the West Carpathians. Tectonophysics, 217: 117142.CrossRefGoogle Scholar
Nemčok, M. (Ed.), 2004. Georgia Niko Well Fracture Analysis Study. EGI Technical Report 01-0059-5000-50501264. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., 2016. Rifts and Passive Margins: Structural Architecture, Thermal Regimes and Petroleum Systems. Cambridge University Press, Cambridge.Google Scholar
Nemčok, M., and Frost, B., 2021. Along-strike magma-poor/magma-rich spreading transitions. Nature, in prep.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: 13011314.CrossRefGoogle Scholar
Nemčok, M., and Kantor, J., 1990. Movement study in the selected area of the Veľký Bok Unit [in Slovak]. In: Regionálna geológia Západných Karpát. State Geological Institute Dionýza Štúra, Bratislava, pp. 7583.Google Scholar
Nemčok, M., and Nemčok, J., 1998. Dynamics and kinematics of the Cirocha, Trangoška and Zázrivá Strike-Slip Faults, Western Carpathians. Geologica Carpathica, 49(2): 8598.Google Scholar
Nemčok, M., and Rosendahl, B. R., 2006. CameroonSpan and GabonSpan Interpretation. EGI Report 01-00059-5000-50501399. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., and Rybár, S., 2016. Rift-Drift Transition in Magma-Rich System: Gop Rift/Laxmi Basin Case Study, West India. Geological Society of London, London, pp. 95117.Google Scholar
Nemčok, M., and Schamel, S., 2000. Structural Characterization of the Elk Hills Field, Southern San Joaquin Basin, California: An Exclusive Research Study for Occidental of Elk Hills, Inc. EGI Report I 00669. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Marko, F., Kováč, M., and Fodor, L., 1989. Neogene tectonics and paleostress changes in the Czechoslovakian part of the Vienna Basin. Jahrbuch der Geologischen Bundesanstalt Wien, 132(2): 443458.Google 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. In: Buchanan, J. G., and Buchanan, P. G. (Eds.), Basin Inversion. Geological Society of London, London, pp. 355392.Google Scholar
Nemčok, M., Pospíšil, L., Lexa, J., and Donelick, R. A., 1998a. Tertiary subduction and break-off model of the Carpathian–Pannonian region. Tectonophysics, 295: 307340.CrossRefGoogle Scholar
Nemčok, M., Houghton, J. J., and Coward, M. P., 1998b. Strain partitioning along the western margin of the Carpathians. Tectonophysics, 292: 119143.CrossRefGoogle Scholar
Nemčok, M., Hók, J., Kováč, P., et al., 1998c. Tertiary extension development and extension/compression interplay in the West Carpathian mountain belt. Tectonophysics, 290: 137167.CrossRefGoogle Scholar
Nemčok, M., Collister, J., James, R., and Schamel, S., 2000. Systematics of Hydrocarbon Exploration and Production in Thrustbelts. EGI Report 01-00059-5000-50500459-3-00. EGI Archive, Salt Lake City.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 modelling. Journal of Structural Geology, 24: 18851901.CrossRefGoogle Scholar
Nemčok, M., Rosendahl, B. R., Segall, M., et al., 2004a. 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. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Moore, J. N., Allis, R., and McCulloch, J., 2004b. Fracture development within a stratovolcano: the Karaha-Telaga Bodas geothermal field, Java volcanic arc. In: Cosgrove, J., and Engelder, T. (Eds.), The Initiation, Propagation, and Arrest of Joints and Other Fractures: A Field Workshop Dedicated to the Memory of Paul Hancock. Geological Society of London, London, pp. 223242.Google Scholar
Nemčok, M., Vangelov, D., Stuart, C., et al., 2004c. Bulgaria Bourgas Block Project Phase 2. EGI Technical Report 50501139. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Stuart, C., Segall, M. P., et al., 2005a. Structural development of South Morocco: interaction of tectonics and deposition. In: Post, P.J., Rosen, N., Olson, D.L., et al. (Eds), Petroleum Systems of Divergent Continental Margin Basins, 25th Annual GCSSEPM Foundation Bob F. Perkins Research Conference, Session I, Crustal Architecture of Divergent Margins. GCSSEPM, Houston, pp. 151202.CrossRefGoogle Scholar
Nemčok, M., Schamel, S., and Gayer, R. A., 2005b. Thrustbelts: Structural Architecture, Thermal Regimes and Petroleum Systems. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Nemčok, M., Pogácsás, G., and Pospíšil, L., 2006a. Activity timing of the main tectonic systems in the Carpathian–Pannonian region in relation to the roll-back destruction of the lithosphere. In: Pícha, F., and Golonka, J. (Eds.), The Carpathians and Their Foreland: Geology and Hydrocarbon Resources, AAPG, Washington, DC, pp. 743766.Google Scholar
Nemčok, M., Moore, J. N., Christensen, C., et al., 2006b. Controls on the Karaha-Telaga Bodas geothermal reservoir, Indonesia. Geothermics, 36: 946.CrossRefGoogle Scholar
Nemčok, M., Dilov, T., Wojtaszek, M., et al., 2007a. Dynamics of the Polish and Eastern Slovakian parts of the Carpathian accretionary wedge: insights from paleostress analyses. In: Ries, A. C., Butler, R. W. H., and Graham, R. H. (Eds.), Deformation of the Continental Crust: The Legacy of Mike Coward. Geological Society of London, London, pp. 271302.Google Scholar
Nemčok, M., Stuart, C., Welker, C., et al., 2007b. Crustal Types, Structural Architecture, and Plate Configurations Study of the Reliance East Coast Region. Phase 1 Project. EGI Report 01-00059-5000-50501546. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Stuart, C., Moore, J. N., et al., 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. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Stuart, C., Odegard, M., et al., 2009. Tectonic Development of the Black Sea Region: Phase 2 of the Circum-Black Sea Project. EGI Report 01-00059-5000-50501563. EGI Archive, Salt Lake City.Google 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 Study. EGI Report 01-00059-5000-50501853. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Greb, M., Allen, J., et al., 2011. Central Atlantic Margins Project, Phase 1 Study Report. EGI Technical Report 01-0059-5000-50501978. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Henk, A., Allen, R., Sikora, P. J., and Stuart, C., 2012a. Continental break-up along strike-slip fault zones: observations from Equatorial Atlantic. In: Mohriak, W. U., Danforth, A., Post, P. J., et al. (Eds.), Conjugate Divergent Margins. Geological Society of London, London, pp. 537556.Google Scholar
Nemčok, M., Sinha, S. T., Stuart, C., et al., 2012b. East Indian margin evolution and crustal architecture: integration of deep reflection seismic interpretation and gravity modeling. In: Mohriak, W. U., Danforth, A., Post, P. J., et al. (Eds.), Conjugate Divergent Margins. Geological Society of London, London, pp. 477496.Google Scholar
Nemčok, M., Stuart, C. J., Rosendahl, B. R., et al., 2012c. Continental break-up mechanism; lessons from intermediate- and fast-extension settings. In: Mohriak, W. U., Danforth, A., Post, P. J., et al. (Eds.), Conjugate Divergent Margins. Geological Society of London, London. pp. 373401.Google Scholar
Nemčok, M., Kotulová, J., Sheya, C., et al., 2012d. Tectonic Development Model and Hydrocarbon Potential of the Guyana-Suriname Basin. EGI Project 01-0059-5000-50502394. Archive EGI, Salt Lake City.Google Scholar
Nemčok, M., Kotulová, J., Matejová, M., et al., 2012e. Northern Brazil Project. EGI Project 01-0059-5000-50502571. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Kotulová, J., Rybár, S., et al., 2013. Tectonic Development Model and Hydrocarbon Potential of the Guyana-Suriname Basin. EGI Technical Report 01-0059-5000-50502394. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Henk, A., Kotulová, J., et al., 2014 Petroleum Systems of Sheared Margins. Phase 1 Study. EGI Project 01-0059-5000-50502487. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Rybár, S., Odegard, M., et al., 2015a. Development history of the southern terminus of the Central Atlantic: Guyana–Suriname case study. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London, pp. 145178.Google Scholar
Nemčok, M., Rybár, S., Ekkertová, P., et al., 2015b. Transform margin model of hydrocarbon migration: Guyana–Suriname case study. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L. (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London, pp. 199217.Google Scholar
Nemčok, M., Stuart, C., Kotulová, J., et al., 2015c. Sheared Margins of Western Australia. Phase 1 Study. EGI Technical Report 01-0059-5000-50502717. EGI Archive, Salt Lake City.Google Scholar
Nemčok, M., Henk, A., and Molčan, M., 2015d. Role of pre-break-up heat flow on thermal history of transform margin. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London, pp. 249271.Google Scholar
Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., 2016a. Transform margins: development, controls and petroleum systems: an introduction. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London, pp. 138.Google Scholar
Nemčok, M., Sinha, S. T., Doré, A. G., et al., 2016b. Mechanisms of microcontinent release associated with wrenching-involved continental breakup: a review. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London, pp. 323359.Google Scholar
Nemčok, M., Welker, C., Sinha, S., et al., 2018. Proximal and distal margin examples from the eastern continental passive margin of India. In: Misra, A. A., and Mukherjee, S. (Eds.), Atlas of Structural Geological Interpretation from Seismic Images. Wiley., Hoboken, pp. 241252.CrossRefGoogle Scholar
Nemčok, M., Ledvényiová, L., Henk, A., et al., 2021a. Early post-breakup kinematic adjustments of transform fault zones. Geology, in prep.Google Scholar
Nemčok, M., Pospíšil, L., Melnik, A., et al., 2021b. Breakup of the last layer during lithospheric separation in magma-poor and magma-rich settings. Nature Reports, in prep.Google Scholar
Nemec, W., 1990. Aspects of sediment movement on steep delta slopes. In: Colella, A. and Prior, D. B. (Eds.), Coarse Grained Deltas. International Association of Sedimentologists, Gent, pp. 2973.CrossRefGoogle Scholar
Neubauer, F., and Genser, J., 1990. Architecture and kinematics of the eastern Central Alps, an overview [in German]. Mitteilungen des Naturwissenschaftlichen Vereines für Steiermark, 120: 203219.Google 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, B, Solid Earth and Planets, 93( 11): 1300513023.CrossRefGoogle Scholar
Ngah, K., Madon, M., and Tjia, H. D., 1996. Role of pre-Tertiary fracture in formation and development of the Malay and Penyu basins. In: Hall, R., and Blundell, D. J. (Eds.), Tectonic Evolution of Southeast Asia. Geological Society of London, London, pp. 281289.Google Scholar
Nichols, A. R. L., Carroll, M. R., and Hoskuldsson, A., 2002. Is the Iceland hot spot also wet? Evidence from the water contents of undegassed submarine and subglacial pillow basalts. Earth and Planetary Science Letters, 202: 7787.CrossRefGoogle Scholar
Nicholson, C., Sorlien, C. C., Atwater, T., Crowell, J. C., and Luyendyk, B. P., 1994. Microplate capture, rotation of the western Transverse Ranges, and initiation of the San Andreas transform as a low-angle fault system. Geology, 22(6): 491495.2.3.CO;2>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. SEPM, Tulsa, pp. 79103.Google Scholar
Nirrengarten, M., Gernigon, L., and Manatschal, G., 2014. Lower crustal bodies in the Møre volcanic rifted margin: geophysical determination and geological implications. Tectonophysics, 636: 143157CrossRefGoogle Scholar
Niu, Y., 2004. Bulk rock major and trace element compositions of abyssal peridotites: implications for mantle melting, melt extraction and post-melting processes beneath Mid-Ocean Ridges. Journal of Petrology, 45(12): 24232458.CrossRefGoogle 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(3): 265283.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. AAPG Bulletin, 90: 16091630.CrossRefGoogle Scholar
Normark, W. R., Piper, D. J. W., Romans, B. W., et al., 2009. Submarine canyon and fan systems of the California continental borderland. In: Lee, H. J., and Normark, W. R. (Eds.), Earth Science in Urban Ocean: the Southern California Continental Borderland. Geological Society of America, Boulder, pp. 141168.Google Scholar
Norris, R. J., 2004. Strain localisation within ductile shear zones beneath active faults: the Alpine Fault contrasted with the adjacent Otago fault system, New Zealand. Earth, Planets and Space, 56: 10951101.CrossRefGoogle Scholar
Norris, R. J., and Carter, R. M., 1980. Offshore sedimentary basins at the Southern End of the Alpine Fault, New Zealand. In: Ballance, P. F., and Reading, H. G. (Eds.), Sedimentation in Oblique-Slip Mobile Zones. International Association of Sedimentologists, Gent, pp. 237265.CrossRefGoogle Scholar
Norris, R. J., and Cooper, A. T., 2001. Late Quaternary slip rates and slip partitioning on the Alpine Fault, New Zealand. Journal of Structural Geology, 23(2): 507520.CrossRefGoogle Scholar
Norris, R. J., and Cooper, A. T., 2003. Very high strains recorded in mylonites along the Alpine Fault, New Zealand: implications for the deep structure of plate boundary faults. Journal of Structural Geology, 25(12): 21412157.CrossRefGoogle Scholar
Norris, R. J., and Cooper, A. F., 2007. The Alpine Fault, New Zealand: surface geology and field relationships. In: Okaya, D., Stern, T., and Davey, F. (Eds.), A Continental Plate Boundary: Tectonics at South Island, New Zealand. AGU, Washington, DC, pp. 157175.CrossRefGoogle Scholar
Norris, R. J., and Toy, V. G., 2014. Continental transforms: a view from the Alpine Fault. Journal of Structural Geology, 64: 331.CrossRefGoogle Scholar
Norton, I., Carruthers, D. T., and Hudec, M. R., 2016. Rift to drift transition in the South Atlantic salt basins: a new flavor of oceanic crust. Geology. DOI: 10.1130/G37265.1CrossRefGoogle Scholar
Nuevo, 2000. Presentation at Africa Upstream Conference, Cape Town, September.Google Scholar
Nunn, J. A., and Deming, D., 1991. Thermal constraints on basin-scale flow systems. Geophysical Research Letters, 18: 967970.CrossRefGoogle 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. AAPG Bulletin, 86: 907918.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: 257262.Google Scholar
Nuriel, P., Weinberger, R., Kylander-Clark, A. R. C., Hacker, B. R., and Craddock, J. P., 2017. The onset of the Dead Sea transform based on calcite age-strain analyses. Geology, 45(7): 587590.CrossRefGoogle 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
Ocean Drilling Program, 2003. Leg 207 Preliminary Report Demerara Rise: Equatorial Cretaceous and Paleogene Paleooceanographic Transect, Western Atlantic, 4 January-6 March 2003. Texas A&M University, College Station.Google Scholar
O’Connell, R. J., Budiansky, B., 1974. Seismic velocities in dry and saturated cracked solids. Journal of Geophysical Research, 79: 54125426.CrossRefGoogle Scholar
O’Hanley, D. S., 1996. Serpentinites: Recorders of Tectonic and Petrological History. Oxford University Press, New York.Google 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(7): 927944.CrossRefGoogle Scholar
Ojeda, H. A. O., 1982. Structural framework, stratigraphy, and evolution of Brazilian marginal basins. AAPG Bulletin, 66: 732749.Google Scholar
Okal, E. A., and Langenhorst, A. R., 2000. Seismic properties of the Eltanin Transform System, South Pacific. Physics of the Earth and Planetary Interiors, 119: 185208.CrossRefGoogle Scholar
Okal, E. A., and Stewart, L. M., 1992. Slow earthquakes along oceanic fracture zones: evidence for asthenospheric flow away from hotspots? Earth and Planetary Science Letters, 57: 7587.CrossRefGoogle Scholar
Olafsson, J., Thors, K., Stefansson, U., et al., 1990. Geochemical observations from a boiling hydrothermal site on the Kolbeinsey Ride, EOS, 71: 1650.Google Scholar
Oldenburg, D. W., and Brune, J. N., 1972. Ridge transform fault spreading pattern in freezing wax. Science 178: 301.CrossRefGoogle ScholarPubMed
Olierook, H. K. H., Merle, R. E., Jourdan, F., et al., 2015. Age and geochemistry of magmatism on the oceanic Wallaby Plateau and implications for the opening of the Indian Ocean. Geology. DOI: 10.1130/G37044.1CrossRefGoogle 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
Olson, S. F., 1989. The stratigraphic and structural setting of the Potrerillos porphyry copper district, northern Chile. Revista Geológica de Chile, 16: 329.Google Scholar
Orange, D. L., Greene, H. G., Reed, D., et al., 1999. Widespread fluid expulsion on a translational continental margin: mud volcanoes, fault zones, headless canyons, and organic-rich substrate in Monterey Bay, California. GSA Bulletin, 111 (7): 9921009.2.3.CO;2>CrossRefGoogle Scholar
Osborne, M. J., and Swarbrick, R. E., 1997. Mechanisms for generating overpressures in sedimentary basins: a reevaluation. AAPG Bulletin, 81: 10231041.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
Ostrolucký, P., 1994: Geological structure and hydrocarbon accumulations in the pre-Neogene base of the Slovak portion of the Vienna Basin. PhD Thesis. Comenius University.Google Scholar
Otofuji, Y., Hayashida, A., and Torii, M., 1985. When was the Japan Sea opened? Paleomagnetic evidence from Southwest Japan. In: Nasu, N., Uyeda, S., Kushiro, I., Kobayashi, K., and Kagami, H. (Eds.), Formation of Active Ocean Margins. Terra Publishing, Tokyo, pp. 551566.CrossRefGoogle Scholar
OTTER Scientific Team 1985. The geology of the Oceanographer transform: the transform domain. Marine Geophysical Research, 7: 329358.CrossRefGoogle 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: Jones, G., Fisher, Q. L., and Knipe, R. J. (Eds.), Faulting, Fault Sealing and Fluid Flow in Hydrocarbon Reservoirs. Geological Society of London, London, pp. 269282.Google Scholar
Palciauskas, V. V., 1986. Models for thermal conductivity and permeability in normally compacting basins. In: Burrus, J. (Ed.), Thermal Modeling of Sedimentary Basins. Technip, Paris, pp. 323336.Google Scholar
Pálmason, G., 1973. Comments on origin and structure of the Iceland Plateau and Kolbeinsey Ridge by Johnson, G. L., Southall, J. R., Young, P. W., and Voght, P. R. Journal of Geophysical Research, 29, 7019.CrossRefGoogle Scholar
Pálmason, G., 1981. Crustal rifting, and related thermo-mechanical processes in the lithosphere beneath Iceland. Geologische Rundschau, 70: 244260.CrossRefGoogle Scholar
Palmer, S. E., 1993. Effect of biodegradation and water washing on crude oil composition. In: Engel, M. H., and Macko, S. A. (Eds.), Organic Geochemistry. Plenum Press, New York.Google Scholar
Parkinson, C., and Dooley, T., 1996. Basin formation and strain partitioning along strike-slip fault zones. Bulletin of the Geological Survey of Japan, 47(8): 427436.Google Scholar
Parmentier, E. M., 1987. Dynamic topography in rift zones: implications for lithospheric heating. Philosophical Transactions of the Royal Society of London, Series A: Mathematical and Physical Sciences, 321(1557): 2325.Google Scholar
Parmentier, E. M., and Forsyth, D. W., 1985. Three-dimensional flow beneath a slow spreading ridge axis: a dynamic contribution to the deepening of the median valley toward fracture zones. Journal of Geophysical Research, 90(B1): 678684.CrossRefGoogle Scholar
Parnell, J., 1994, Geofluids: Origin, Migration and Evolution of Fluids in Sedimentary Basins. Geological Society of London, London.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: 601631.CrossRefGoogle Scholar
Parson, L. M., Murton, B. J., Searle, R. C., et al., 1993. En echelon axial volcanic ridges at the Reykjanes Ridge: a life cycle of volcanism and tectonics. Earth and Planetary Science Letters, 117: 7387.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: 803829.CrossRefGoogle Scholar
Parsons, W., Hawke, C., and Malborg, P., 2017. Turning exploration inside out: facies prediction for Ayame-1X, Côte d’Ivoire. In: Abstract Volume of the HGS PESGB 16th African Conference, 31st August–1st September. Design Centre, London.Google Scholar
Partridge, T. C., and Maud, R. R., 1987. Geomorphic evolution of southern Africa since the Mesozoic. South African Journal of Geology, 90: 179208.Google Scholar
Patruno, S., and Helland-Hansen, W., 2018. A dynamic classification of clinoforms. Geophysical Research Abstracts, 20: EGU2018–2755.Google Scholar
Paul, D. K., Ray Barman, T. K., McNaughton, N.J, et al., 1990. Archaean–Proterozoic evolution of Indian charnockites: isotopic and geochemical evidence from granulites of Eastern Ghat belt. Journal of Geology, 98: 253263.CrossRefGoogle Scholar
Pécskay, Z., Lexa, J., Szakacs, A., et al., 1995. Space and time distribution of Neogene–Quaternary volcanism in the Carpatho-Pannonian region. Acta Vulcanologica, 7(2): 1528.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
Pepper, A. S., 1991. Estimating the petroleum expulsion behavior of source rocks: a novel quantitative approach. In: England, W. A. and Fleet, A. J. (Eds.), Petroleum Migration. Geological Society of London, London, pp. 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: 321340.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: 217226.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: 451466.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
Pereira, M., Vergani, G., Cambon, I., et al., 2018. Andean deformation and its control on hydrocarbon generation, migration, and charge in the wedge-top of southern Bolivia. In: Zamora, G. McClay, K. R., and Ramos, V. A. (Eds.), Petroleum Basins and Hydrocarbon Potential of the Andes of Peru and Bolivia. AAPG, Washington, DC, pp. 531554.Google 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: 39613976.CrossRefGoogle Scholar
Perfit, M. R., Fornani, D. J., Ridley, W. I., et al., 1996. Recent volcanism in the Siqueiros transform fault: picritic basalts and implications for MORB magma genesis, Earth and Planetary Science Letters, 141: 91108.CrossRefGoogle Scholar
Perincek, D., and Cemen, I., 1990. The structural relationship between the East Anatolian and Dead Sea fault zones in southeastern Turkey. Tectonophysics, 172: 331340.CrossRefGoogle Scholar
Perner, M., Kuever, J., Seifert, R., et al., 2007. The influence of ultramafic rocks on microbial communities at the Logatchev hydrothermal field, located 15 degrees N on the Mid-Atlantic Ridge. FEMS Microbiology Ecology, 61(1): 97109.CrossRefGoogle ScholarPubMed
Péron-Pinvidic, G., Manatschal, G., and Osmundsen, P. T., 2013. Structural comparison of archetypal Atlantic rifted margins: a review of observation and concepts. Marine and Petroleum Geology 43: 2147.CrossRefGoogle Scholar
Person, M., 1990. Hydrologic constraints on the thermal evolution of continental rift basins: implications for petroleum maturation. PhD Thesis. Johns Hopkins University.Google Scholar
Person, M., and Garven, G., 1992. Hydrologic constraints on petroleum generation within continental rift basins: theory and application to the Rhine Graben. AAPG Bulletin, 76(4): 468488.Google Scholar
Peters, K. E., 1986. Guidelines for evaluating petroleum source rock using programmed pyrolysis. AAPG Bulletin, 70: 318329.Google Scholar
Peters, K. E., and Cassa, M. R., 1994. Applied source rock geochemistry. In: Magoon, L. B. and Dow, W. G. (Eds.), The Petroleum System: From Source to Trap, AAPG, Washington, DC, pp. 93120.Google Scholar
Peters, K. E., and Moldowan, J. M., 1993. The Biomarker Guide. Interpreting Molecular Fossils in Petroleum and Ancient Sediments. Prentice Hall, Englewood Cliffs.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, U.S.A. In: Magoon, L. B., and Dow, W. E. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 423436.Google Scholar
Peters, K. E., Walters, C. C., and Moldowan, J. M., 2005. The Biomarker Guide, 2nd edition. Cambridge University Press, Cambridge.Google Scholar
Petford, N., Cruden, A. R., McCaffrey, K. J. W., and Vignerese, J.-L., 2000. Granite magma formation, transport and emplacement in the Earth’s crust. Nature, 408: 669673.CrossRefGoogle ScholarPubMed
Petit, J.-P., 1987. Criteria for the sense of movement on fault surfaces in brittle rocks. Journal of Structural Geology, 9(5–6): 597608.CrossRefGoogle Scholar
Petit, J.-P., Wibberley, C. A., and Ruiz, G., 1999. “Crack-seal”, slip: a new fault valve mechanism? Journal of Structural Geology, 21: 11991207.CrossRefGoogle Scholar
Philip, H., Rogozhin, E., Cisternas, A., et al., 1992. The Armenian earthquake of 1988 December 7: faulting and folding, neotectonics and palaeoseismicity. Geophysical Journal International, 110: 141158.CrossRefGoogle Scholar
Phipps Morgan, J., and Chen, Y. J., 1993. The genesis of oceanic crust: magma injection, hydrothermal circulation, and crustal flow. Journal of Geophysical research, 98: 62836297.CrossRefGoogle Scholar
Phipps Morgan, J. P., and Forsyth, D., 1988. Three-dimensional flow and temperature perturbations due to a transform offset: effects on oceanic crustal and upper mantle structure. Journal of Geophysical Research, 93: 29552966.CrossRefGoogle Scholar
Phipps Morgan, J., Parmentier, E. M., and Lin, J., 1987. Mechanisms for the origin of mid-ocean ridge axial topography: implications for the thermal and mechanical structure of accreting plate boundaries. Journal of Geophysical Research, 92: 1282312836.CrossRefGoogle Scholar
Pícha, F. J., and Peters, K. E., 1998. Biomarker oil-to-source rock correlation in the Western Carpathians and their foreland, Czech Republic. Petroleum Geoscience, 4: 289302.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 (Boulder), 24(12): 10791082.2.3.CO;2>CrossRefGoogle Scholar
Pierce, C., Whitmarsh, R. B., Scrutton, R. A., et al., 1996. Cote d’Ivoire–Ghana margin: seismic imaging of passive rifted crust adjacent to a transform continental margin. Geophysical Journal International, 125: 781795.CrossRefGoogle Scholar
Pindell, J., and Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and northern South America in the mantle reference frame: an update. In: James, K. H., Lorente, M. A., and Pindell, J. L. (Eds.), Origin and Evolution of the Caribbean Plate. Geological Society of London, London, pp. 155.Google Scholar
Pinet, P., and Souriau, M., 1988. Continental erosion and large-scale relief. Tectonics, 7: 563582.CrossRefGoogle Scholar
Piper, D. J. W., and Normark, W. R., 2001. Sandy fans: from Amazon to Hueneme and beyond. AAPG Bulletin, 85(8): 14071438.Google Scholar
Piper, D. J. W., and Normark, W. R., 2009. Processes that initiate turbidity currents and their influence on turbidites: a marine geology perspective. Journal of Sedimentary Research, 79: 347362.CrossRefGoogle Scholar
Piper, D. J. W., Shaw, J., and Skene, K. I., 2007. Stratigraphic and sedimentological evidence for late Wisconsinan subglacial outburst floods to Laurentian Fan. Palaeogeography, Palaeoclimatology, Palaeoecology, 246: 101119.CrossRefGoogle Scholar
Pirmez, C., Beaubouef, R. T., Friedmann, S.J., and Mohrig, D. C., 2000. Equilibrium profile and base level in submarine channels: examples from late Pleistocene systems and implications for the architecture of deepwater reservoirs. Gulf Coast Section SEPM (Society of Sedimentary Geology) 20th Annual Research Conference Proceedings, pp. 782–805.CrossRefGoogle Scholar
Pitman, W. C., and Andrews, J. A., 1985. Subsidence and thermal history of small pull-apart basins. In: Biddle, K. T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation and Sedimentation. Society of Economic Paleontologists and Mineralogists, San Antonio, 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, Reston.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: 1933519351.CrossRefGoogle Scholar
Platte River Associates, 1995. BasinMod* 1-D for WindowsTM. Basin Modeling System. Document Version: 5.0.Google Scholar
Platzman, E. S., Platt, J. P., Tapirdamaz, C., Sanver, M., and Rundle, C. C., 1994. Why are there no clockwise rotations along the Northern Anatolian Fault Zone? Journal of Geophysical Research, Solid Earth, 99: 2170521715.CrossRefGoogle Scholar
Pockalny, R. A., Gente, P., and Buck, W. R., 1996. Oceanic transverse ridges: a flexural response to fracture-zone-normal extension. Geology, 24: 7174.2.3.CO;2>CrossRefGoogle Scholar
Pockalny, R. A., Fox, P. J., Fornari, D. J., Macdonald, K.C, and Perfit, M. R., 1997. Tectonic reconstruction of the Clipperton and Siqueiros Fracture Zones: evidence and consequences of plate motion change for the last 3 Myr. Journal of Geophysical Research, 102(B2): 31673181.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: Welte, D. H., Horsfield, B., and Baker, D. R. (Eds.), Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling. Springer, Berlin, pp. 370.CrossRefGoogle Scholar
Pogacsás, G., Mattick, R. E., Tari, G., and Varnai, P., 1994. Structural control on hydrocarbon accumulation in the Pannonian Basin, Hungary. In: Teleki, P. G., Mattick, R. E., and Kokay, J. (Eds.), Basin Analysis for Oil and Gas Exploration: A Case History from Hungary. Kluwer, Amsterdam, pp. 221235.Google Scholar
Pollack, H. N., and Chapman, D. S., 1977. On the regional variation of heat flow, geotherms, and lithospheric thickness. Tectonophysics, 38: 279296.CrossRefGoogle Scholar
Popescu, B. M., 1995. The Guyanas (Guyana, Suriname, French Guyana). In: Kulke, H. (Ed.), Regional Petroleum Geology of the World. Vol. 22, Part 2. G. Borntraeger, Stuttgart, pp. 603612.Google Scholar
Popoff, M., 1988. Du Gondwana a l’Atlantique sud; les connexions du fosse de la Benoue, avec les bassins du nord-est bresilien jusqu’ a l’ouverture du golfe de Guinee au Cretace inferieur. 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, 7: 409431.CrossRefGoogle Scholar
Popov, A. A., Sobolev, S. V., and Zoback, M. D., 2012. Modeling evolution of the San Andreas Fault system in northern and central California. Geochemistry, Geophysics, Geosystems 13. DOI: 10.1029/2012GC004086CrossRefGoogle Scholar
Poreda, R. J., Craig, H., Arnorsson, S., and Welhan, J. A., 1992. Helium isotopes in Icelandic geothermal systems: I. 3He, gas chemistry, and 13C relations. Geochimica et Cosmochimica Acta, 56: 42214228.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. SEPM, Tulsa.CrossRefGoogle Scholar
Posamentier, H. W., Erskine, R. D., and Mitchum, R. M., Jr., 1991. Submarine fan deposition within a sequence stratigraphic framework. In: Weimer, P., and Link, M. H. (Eds.), Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems. Springer, New York, pp. 127136.CrossRefGoogle Scholar
Powell, W., Burrell, A., Diaby, I., and Gauly, M., 2019. Côte d’Ivoire: regional understanding to unlock hydrocarbon potential. GeoEcPro, 16(5): 4244.Google Scholar
Power, W. L., and Tullis, T. E., 1989. The relationship between slickenside surfaces in fine-grained quartz and the seismic cycle. Journal of Structural Geology, 11(7): 879893.CrossRefGoogle Scholar
Powley, D. E., 1980. Pressures, normal and abnormal. AAPG 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 [in French]. Documents et Travaux de l’Institut Geologique Albert de Lapparent, 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(1): 5984.CrossRefGoogle Scholar
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.Google Scholar
Prins, M. A., Postma, G., Cleveringa, J., Cramp, A., and Kenyon, N. H., 2000. Controls on terrigenous sediment supply to the Arabian Sea during the late Quaternary; the Indus Fan. Marine Geology, 169: 327349.CrossRefGoogle Scholar
Prinzhofer, A., Santos Neto, E. V., Takaki, T., Mello, M. R., and Roos, S., 1998. The gas system in the Potiguar Basin: relations with the petroleum genesis and migration (abs). In: 6th Latin American Congress on Organic Geochemistry, Isla Margarita, Venezuela.Google Scholar
Prochác, R., Pereszlényi, M., and Sopková, B., 2012. Tectono-sedimentary features in 3D seismic data from the Moravian part of the Vienna Basin. First Break, 30: 4956.CrossRefGoogle Scholar
Provost, A.-S., Chery, J., and Hassani, R., 2003. 3D mechanical modeling of the GPS velocity field along the North Anatolian fault. Earth and Planetary Science Letters, 209(3–4): 361377.CrossRefGoogle Scholar
Puthe, C., and Gerya, T., 2014. Dependence of mid-ocean ridge morphology on spreading rate in numerical 3-D models. Gondwana Research, 25: 270283.CrossRefGoogle Scholar
Quadri, V. N., and Quadri, S. M. G. J., 1996. Exploration anatomy of success in oil and gas exploration in Pakistan, 1915–94. Oil and Gas Journal, 94: 120.Google Scholar
Quennell, A. M., 1956. The structural and geomorphic evolution of the Dead Sea Rift. Quarterly Journal of the Geological Society of London, 114: 124.CrossRefGoogle Scholar
Quennell, A. M., 1984. The Western Arabia Rift System. In: Dixon, J. E., and Robertson, A.H.F. (Eds.), Geological Evolution of the Eastern Mediterranean. Geological Society, London, pp. 775788.Google Scholar
Quigley, T. M., and Mackenzie, A. S., 1988. The temperatures of oil and gas formation in the subsurface. Nature, 333: 549552.CrossRefGoogle Scholar
Quittmeyer, R. C., Farah, A., and Jacob, K. H., 1979. The seismicity of Pakistan and its relation to surface faults. In: Farah, A., and DeJong, K.A. (Eds.), Geodynamics of Pakistan. Geological Survey of Pakistan, Balochistan, pp. 271284.Google 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: Roth, S., and Rueggeberg, A. (Eds.), International Conference and Annual Meeting of the Deutsche Geologische Gesellschaft and Geologische Vereinigung. Deutschen Geologischen Gesellschaft, Kiel, pp. 157.Google Scholar
Rabinowitz, N., and Steinberg, J., 1996. New evidence of magmatic diapirs in the intermediate crust under the Dead Sea, Israel. Tectonics, 15(2): 237243.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
Rădulescu, F., 1988. Seismic models of the crustal structure in Romania. Revue roumaine de géologie, géophysique et géographie, 32: 1319.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
Ramakrishnan, M., 1990. Crustal development in southern Bastar central Indian craton. Geological Survey of India Special Publications, 28: 4466.Google Scholar
Ramakrishnan, M., Nanda, J. K., and Augustine, P. F., 1998. Geological evolution of the Proterozoic Eastern Ghats Mobile Belt. Geological Survey of India Special Publications, 44: 121.Google Scholar
Ramana, M. V., Ramprasad, T., and Desa, M., 2001. Seafloor spreading magnetic anomalies in the Enderby Basin, East Antarctica. Earth and Planetary Science Letters, 191: 241255.CrossRefGoogle Scholar
Ramsay, J. G., 1980. Shear zone geometry: a review. Journal of Structural Geology, 2: 8399.CrossRefGoogle Scholar
Ramsay, J. G., and Huber, M. I., 1987. The Techniques of Modern Structural Geology, volume 2. Academic Press, London.Google Scholar
Ranalli, G., 1995. Rheology of the Earth, 2nd edition. Springer, Berlin-Heidelberg.Google 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
Randive, K., Hari, K. R., Dora, M. L., Malpe, D. B., and Bhondwe, A. A., 2014. Study of fluid inclusions: methods, techniques and applications. Gondwana Geological Magazine, 29(1–2): 1928.Google Scholar
Ranero, C., and Reston, T. J., 1999. Detachment faulting at ocean core complexes. Geology, 27: 983986.2.3.CO;2>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
Rasbury, E. T., and Cole, J. M., 2009. Directly dating geologic events: U–Pb dating of carbonates. Reviews of Geophysics, 47. DOI: 10.1029/2007RG000246CrossRefGoogle Scholar
Rasskazov, S. V., Boven, A., Ivanov, A. V., and Semenova, V. G., 2000. Middle Quaternary volcanic impulse in the Olekma-Stanovoy mobile system: 40Ar–39Ar dating of volcanics from the Tokinsky Stanovik. Geology of the Pacific Ocean, 19(4): 1928.Google Scholar
Ratschbacher, L., Frisch, W., and Linzer, H. G., 1991a. Lateral extrusion in the Eastern Alps, Part 2: structural analysis. Tectonics, 10(2): 257271.CrossRefGoogle Scholar
Ratschbacher, L., Merle, O., Davy, P., and Cobbold, P., 1991b. Lateral extrusion in the Eastern Alps, Part 1: boundary conditions and experiments scaled for gravity. Tectonics, 10(2): 245256.CrossRefGoogle Scholar
Ray, D. K., 1963. Tectonic Map of India, scale 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: 645672.CrossRefGoogle Scholar
Raynolds, R. G., 1980. The Plio-Pleistocene structural and stratigraphic evolution of the eastern Potwar Plateau, Pakistan. PhD thesis, Dartmouth College, Hanover.Google Scholar
Raza, H. A., Ahmed, R., Manshoor, S., and Ahmad, J., 1989. Petroleum prospects: Sulaiman sub-basin, Pakistan. Pakistan Journal of Hydrocarbon Research, 1(2): 2156.Google Scholar
Redin, T., 1991. Oil and gas production from submarine fans of the Los Angeles Basin. In: Biddle, K. T. (Ed.), Active Margin Basins. AAPG, Washington, DC, pp. 239259.Google Scholar
Reeves, C. V., Sahu, B. K., and de Wit, M., 2002. A re-examination of the paleo-position of Africa’s eastern neighbours in Gondwana. Journal of African Earth Sciences, 34(3–4): 101108.CrossRefGoogle Scholar
Reeves, C. V., Teasdale, J. P., and Mahanjane, E. S., 2016. Insight into the Eastern Margin of Africa from a new tectonic model of the Indian Ocean. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., (Eds), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London, pp. 299322.Google 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. Bol. IG-USP, Publ. Esp. 7: 139143.CrossRefGoogle Scholar
Regenauer-Lieb, K., and Yuen, D. A., 2000. Fast mechanisms for the formation of new plate boundaries. Tectonophysics, 322(1): 5367.CrossRefGoogle Scholar
Regenauer-Lieb, K., and Yuen, D. A., 2003. Modeling shear zones in geological and planetary sciences: solid- and fluid-thermal-mechanical approaches. Earth Science Reviews, 63(3): 295349.CrossRefGoogle Scholar
Regenauer-Lieb, K., Hobbs, B., Yuen, D. A., et al., 2006. From point defects to plate tectonic faults. Philosophical Magazine, 86(21): 33733392.CrossRefGoogle Scholar
Reid, I., 1989. Effects of lithospheric flow on the formation and evolution of a transform margin. Earth and Planetary Science Letters, 95: 3852.CrossRefGoogle Scholar
Reid, I., and Jackson, H. R., 1997. Crustal structure of northern Baffin Bay: seismic refraction results and tectonic implications. Journal of Geophysical Research, 102: 523542.CrossRefGoogle Scholar
Reilinger, R., McClusky, S., Paradissis, D., Ergintav, S., and Vernant, P., 2009. Geodetic constraints on the tectonic evolution of the Aegean region and strain accumulation along the Hellenic subduction zone. Tectonophysics, 488: 2230.CrossRefGoogle Scholar
Reilinger, R., McClusky, S. N., Vernant, P., et al., 2006. GPS constraints on continental deformation in the Africa–Arabia–Eurasia continental collision zone and implications for the dynamics of plate interactions. Journal of Geophysical Research: Solid Earth, 111(B5). DOI: 10.1029/2005JB004051CrossRefGoogle Scholar
Reinen, L. A., Weeks, J. D., and Tullis, T. E., 1994. The frictional behavior of lizardite and antigorite serpentinites: experiments, constitutive models, and implications for natural faults. Pure and Applied Geophysics, 143: 318358.CrossRefGoogle Scholar
Reiners, P. W., Campbell, I., Nicolescu, S., et al., 2002. Single crystal helium–lead dating of detrital zircon. Abstracts with Programs: Geological Society of America, 34(6): 484.Google Scholar
Reston, T. J., 2007. The formation of non-volcanic rifted margins by the progressive extension of the lithosphere. In: Karner, G. D., Manatschal, G., and Pinheiro, L. M. (Eds.), Imaging, Mapping and Modeling Continental Lithosphere Extension and Breakup. Geological Society of London, London, pp. 77110.Google Scholar
Reutter, K.-J., Heinsohn, W. D., Scheuber, E., and Wigger, P., 1991. Crustal structure of the Coastal Cordillera near Antofagasta, northern Chile: 6th Congreso Geológico Chileno. Actas, 1: 862866.Google Scholar
Reyners, M., 2013. The central role of the Hikurangi Plateau in the Cenozoic tectonics of New Zealand and the Southwest Pacific. Earth and Planetary Science Letters, 361: 460468.CrossRefGoogle Scholar
Reynolds, A. D., Simmons, M. D., Bowman, M. B. J., et al., 1998. Implications of outcrop geology for reservoirs in the Neogene productive series: Apsheron Peninsula, Azerbaijan. AAPG Bulletin, 82: 2549.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, J. R., and Langmuir, C. H., 1997. Petrological systematic of the Mid-Atlantic Ridge south of Kane: implications for ocean crust formation. Journal of Geophysical Research, 102: 1491514946.CrossRefGoogle Scholar
Reynolds, J. R., Langmuir, C. H., Bender, J. F., Kastens, K. A., and Ryan, W. B. F., 1992. Spatial and temporal variability in the geochemistry of basalts from the East Pacific Rise. Nature, 359: 493499.CrossRefGoogle Scholar
Reynolds, K., Copley, A., and Hussain, E., 2015. Evolution and dynamics of a fold-thrust belt: the Sulaiman Range of Pakistan. Geophysical Journal International, 201(2): 683710.CrossRefGoogle Scholar
Rialto Energy Ltd., 2012. Good Oil conference presentation. September 5.Google Scholar
Rice, J. R., 1992. Fault stress states, pore pressure distributions, and the weakness of the San Andreas fault. In: Evans, B.W. (Ed.), Earthquake Mechanics, Rock Deformation, and Transport Properties of Rocks. Academic Press, San Diego, pp. 475503.Google Scholar
Richard, P. D., Naylor, M. A., and Koopman, A., 1995. Experimental models of strike-slip tectonics. Petroleum Geoscience, 1: 7180.CrossRefGoogle Scholar
Riedel, C., Schmidt, M., Botz, R., and Theilen, F., 2001. The Grimsey hydrothermal field offshore North Iceland: crustal structure, faulting and related gas venting. Earth and Planetary Science Letters. 193(3–4): 409421.CrossRefGoogle Scholar
Riedel, C., Petersen, T., Theilen, F., and Neben, S., 2003. High b-values in the leaky segment of the Tjörnes Fracture Zone north of Iceland: are they evidence for shallow magmatic heat sources? Journal of Volcanology and Geothermal Research, 128: 1529.CrossRefGoogle Scholar
Riedel, W., 1929. Zum Mechanik geologischer Brucherscheinungen. Zentralblatt fur Mineralogie, Geologie und Paleontologie B, 1929: 354368.Google Scholar
Riegl, B., and Piller, W. E., 2000. Biostromal coral facies: a Miocene example from the Leitha Limestone (Austria) and its actualistic interpretation. Palaios, 15: 399413.2.0.CO;2>CrossRefGoogle Scholar
Riffault, R., 1969. Catalogue des characteristiques geologiques et mechaniques de quelques roches Francaise. Laboratoire Central des Ponts et Chaussees, Paris.Google Scholar
Ring, U., and Gerdes, A., 2016. Kinematics of the Alpenrhein-Bodensee graben system in the Central Alps: Oligocene/Miocene transtension due to formation of the Western Alps arc. Tectonics. DOI: 10.1002/2015TC004085CrossRefGoogle Scholar
Ring, U., Brandon, M. T., Willett, S. D., and Lister, G. S., 1998. Exhumation processes. In: Ring, U., Brandon, M. T., Lister, G. S., and Willett, S. D. (Eds.), Exhumation Processes: Normal Faulting, Ductile Flow and Erosion. Geological Society of London, London, pp. 129.Google Scholar
Ritter, O., Ryberg, T., Weckmann, U., et al., 2003. Geophysical images of Dead Sea Transform in Jordan reveal an impermeable barrier for fluid flow. Geophysics Letters 30. DOI: 10/1029/2003GL017541Google Scholar
Robb, M. S., Taylor, B., and Goodlife, A. M., 2005. Re-examination of the magnetic lineations of the Gascoyne and Cuvier Abyssal Plains, off NW Australia. Geophysical Journal International, 163(1): 4255.CrossRefGoogle Scholar
Roberts, D. G., Thompson, M., Mitchener, B., et al., 1999. Palaeozoic to Tertiary rift and basin dynamics: mid-Norway to the Bay of Biscay – a new context for hydrocarbon prospectivity in the deep water frontier. In: Fleet, A. J., and Boldy, S. A. R. (Eds.), Petroleum Geology of Northwest Europe, Proceedings of the 5th Conference, pp. 7–40.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: 863879.CrossRefGoogle Scholar
Roberts, N. M. W., and Walker, R. J., 2016. U–Pb geochronology of calcite-mineralized faults: absolute timing of rift-related fault events on the northeast Atlantic margin. Geology, 44(7): 531534.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: 2823128252.CrossRefGoogle Scholar
Robinson, A. C., Yin, A., Manning, C. E., et al., 2007. Cenozoic evolution of the eastern Pamir: implications for strain-accommodation mechanisms at the western end of the Himalayan–Tibetan orogen. Geological Society of America Bulletin, 119: 882896.CrossRefGoogle Scholar
Robinson, D. P., 2011. A rare great earthquake on an oceanic fossil fracture zone. Geophysical Journal International, 186: 11211134.CrossRefGoogle Scholar
Robinson, D., and Santana de Zamora, A., 1999. The smectite to chlorite transition in the Chipilapa geothermal system, El Salvador. American Mineralogist, 84: 606619.CrossRefGoogle Scholar
Robinson, G. R., Jr., and Woodruff, L. G., 1988, Characteristics of base-metal and barite vein deposits associated with rift basins, with examples from some early Mesozoic basins of eastern North America. In: Froelich, A. J., and Robinson, G. R., Jr. (Eds.), Studies of the Early Mesozoic Basins of the Eastern U.S. US Geological Survey, Reston, pp. 377390.Google Scholar
Robinson, P. T., and Whitford, D. J., 1974. Basalts from the Indian Ocean, DSDP Leg 27. Initial Reports of the Deep Sea Drilling Project Vol. 27, US Government Printing Office, Washington, DC, pp. 551559.Google Scholar
Robinson, P. T., Elders, W. A., and Muffler, L. J. P., 1976. Quaternary volcanism in the Salton Sea geothermal field, Imperial Valley, California. Geological Society of America Bulletin, 87(3): 347360.2.0.CO;2>CrossRefGoogle Scholar
Robinson, P. T., Dick, H. J. B., Natland, J. H., et al., 2000. Lower oceanic crust formed at an ultra-slow-spreading ridge; Ocean Drilling Program Hole 735B, Southwest Indian Ridge. Geological Society of America Special Paper 349.CrossRefGoogle 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: 850858.2.3.CO;2>CrossRefGoogle Scholar
Rodolfo, K. S., 1969. Bathymetry and marine geology of Andaman Basin, and tectonic implications for southeast Asia. GSA Bulletin, 80: 12031230.CrossRefGoogle Scholar
Rodrigues, R., Françolin, J. B. L., and Lima, H. P., 1983, Avaliação geoquímica preliminary da Bacia Potiguar terrestre. Petrobras internal report.Google Scholar
Rodrigues, R., Santos, A. S., and Costa, L. A., 1984. Avaliacao geoquimica da Bacia de Barreirinhas. Petrobras internal report.Google Scholar
Roedder, E., 1981. Problems in the use of fluid inclusions to investigate fluid–rock interaction in igneous and metamorphic processes. Fortschritte der Mineralogie, 59: 267302.Google Scholar
Roedder, E., 1984. Fluid Inclusions. Mineralogical Society of America. Washington, DC.CrossRefGoogle Scholar
Roedder, E., 1990. Fluid inclusion analysis: prologue and epilogue. Geochimica et Cosmochimica Acta, 54: 495507.CrossRefGoogle Scholar
Rohr, K., 1994. Increase of seismic velocities in upper oceanic crust and hydrothermal circulation in the Juan de Fuca plate. Geophysical Research Letters, 21: 21632166.CrossRefGoogle Scholar
Rohr, K., Schmidt, U., Lowe, C., and Milkereit, B., 1994. Multichannel Seismic Reflection Data Across Endeavour Segment, Juan De Fuca Ridge. Open File 2847. Geological Survey of Canada, Sidney, BC.CrossRefGoogle Scholar
Rohrman, M., 2013a. Greater Exmouth LIP (NW Australia). LIP of the Month, November 2013. Large Igneous Provinces Commission, International Association of Volcanology and Chemistry of the Earth’s Interior. Available at: www.largeigneousprovinces.org/13nov.Google Scholar
Rohrman, M., 2013b. Intrusive large igneous provinces below sedimentary basins: an example from the Exmouth Plateau (NW Australia). Journal of Geophysical Research, 118(8): 44774487.CrossRefGoogle Scholar
Roland, E., Behn, M. D., and Hirth, G., 2010. Thermal-mechanical behavior of oceanic transform fault: implications for the spatial distribution of seismicity. Geochemistry, Geophysics, Geosystems, 11(7). DOI: 10.1029/2010GC003034CrossRefGoogle Scholar
Román-Berdiel, T., Aranguren, A., Cuevas, J., et al., 2000. Experiments on granite intrusions in transtension. In: Vigneresse, J. L., Mart, Y., and Vendeville, B. (Eds.), Salt, Shale and Igneous Diapirs in and Around Europe. Geological Society of London, London, pp. 2142.Google Scholar
Rona, P. A., Bougault, H., Charlou, J. L., et al., 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: 783786.2.3.CO;2>CrossRefGoogle Scholar
Rongvaldsson, S. T., Gudmundsson, A., and Slunga, R., 1998. Seismotectonic analysis of the Tjornes Fracture Zone, an active transform fault in North Iceland. Journal of Geophysical Research, 103: 3011730129.CrossRefGoogle Scholar
Roscoe, K. H., and Burland, J. B., 1968. On the generalized stress–strain behavior of “wet” clay. In: Heyman, J., and Leckie, F. (Eds.), Engineering Plasticity. Cambridge University Press, Cambridge, pp. 535609.Google Scholar
Roscoe, K. H., Schofield, A. N., and Wroth, C. P., 1958. On the yielding of soils. Geotechnique, 8: 2253.CrossRefGoogle 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. In: Cameron, N. R., Bate, R. H., and Clure, V. S. (Eds.), The Oil and Gas Habitats of the South Atlantic. Geological Society of London, London, pp. 7583.Google Scholar
Rosendahl, B. R., Reynolds, D. J., Lorber, P. M., et al., 1986. Structural expressions of rifting: lessons from Lake Tanganyika, Africa. In: Frostick, L. E., Renault, R. W., Reid, I., and Tiercelin, J. J., (Eds.), Sedimentation in the African Rifts. Geological Society of London, London, pp. 2943.Google 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(4): 291295.2.3.CO;2>CrossRefGoogle Scholar
Rosendahl, B. R., Kilembe, E., and Kaczmarcik, K., 1992. Comparison of the Tanganyika, Malawi, Rukwa and Turkana rift zones from analyses of seismic reflection data. Tectonophysics, 213: 235256.CrossRefGoogle Scholar
Rosendahl, B. R., Mohriak, W. U., Nemčok, M., et al., 2005. West African and Brazilian conjugate margins: crustal types, architecture, and plate configurations. In: Programs and Abstracts – Society of Economic Paleontologists. Gulf Coast Section. Research Conference, volume 25, pp. 1314.Google Scholar
Rosensaft, M., and Sneh, A., 2017. The Geological Map of Israel, scale 1:200,000, version 2014. GSI WebApp #1/2017. Geological Survey of Israel, Jerusalem.Google Scholar
Rosenthal, M., Ben-Avraham, Z., and Schattner, U., 2019. Almost a sharp cut: a case study of the cross point between a continental transform and a rift, based on 3D gravity modelling. Tectonophysics, 761: 4664.CrossRefGoogle Scholar
Ross, M., and Scotese, C., 1988. A hierarchical tectonic model of the Gulf of Mexico and Caribbean region. Tectonophysics, 155: 139168.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. AAPG Bulletin, 83(9): 14541484.Google Scholar
Roy, R. F., Beck, A. E., and Toulokian, Y. S., 1981. Thermophysical properties of rocks. In: Toulokian, Y. S., Judd, W. R., and Roy, R. F. (Eds.), Physical Properties of Rocks and Minerals. McGraw-Hill, New York, pp. 409488.Google Scholar
Royden, L., 1988. Late Cenozoic tectonics of the Pannonian basin system. In: Royden, L., and Horváth, F. (Eds.), The Pannonian Basin: A Study in Basin Evolution. AAPG, Washington, DC, pp. 2748.Google Scholar
Royden, L. H., 1985. The Vienna Basin: a thin-skinned pull-apart basin. In: Biddle, K. T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation, and Sedimentation. SEPM, Tulsa, pp. 319338.CrossRefGoogle 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(2): 343361.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., 2003. New insights on the Green River petroleum system in the Uinta basin from hydrous pyrolysis experiments. Reply. AAPG Bulletin, 87: 15351541.CrossRefGoogle 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
Rullkötter, J., and Mukhopadhyay, P. K., 1986. Comparison of Mesozoic carbonaceous claystones in the western and eastern North Atlantic (DSDP Legs 76, 79, and 93). In: Summerhayes, C. P., and Shackleton, N. J. (Eds.), North Atlantic Paleoceanography. Geological Society of London, London, pp. 377387.Google Scholar
Rümpker, G., Ryberg, T., Bock, G., and Desert Seismology Group, 2003. Boundary-layer mantle flow under the Dead Sea transform fault inferred from seismic anisotropy. Nature, 425(6957): 497501.CrossRefGoogle ScholarPubMed
Ruppel, C., 1995. Extensional processes in continental lithosphere. Journal of Geophysical Research, 100(B12): 2418724215.CrossRefGoogle Scholar
Rupprecht, B. J., Sachsenhofer, R. F., Zach, C., et al., 2018. Oil and gas in the Vienna Basin: hydrocarbon generation and alteration in a classical hydrocarbon province. Petroleum Geoscience, 25: 329.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: 706730.CrossRefGoogle Scholar
Rust, D. J., and Summerfield, M. A., 1990. Isopach and borehole data as indicators of rifted margin evolution in southwestern Africa. Marine and Petroleum Geology, 7(3): 277287.CrossRefGoogle Scholar
Rutter, E. H., 1976. The kinetics of rock deformation by pressure solution. Philosophical Transactions of the Royal Society of London, A283: 203220.Google Scholar
Rutter, E. H., 1983. Pressure solution in nature, theory and experiment. Journal of Geological Society of London, 140: 725740.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: 102109.Google Scholar
Ryan, M. C., Helland-Hansen, W., Johannessen, E. P., and Steel, R. J., 2009, Erosional versus accretionary shelf margins: the influence of margin type of deepwater sedimentation– an example from the Porcupine Basin, offshore western Ireland. Basin Research. DOI: 10.1111/j.1365-2117.2009.00424.xCrossRefGoogle Scholar
Rybach, L., 1976. Radioactive heat production in rocks and its relation to other petrophysical parameters. Pure and Applied Geophysics, 114: 309317.CrossRefGoogle Scholar
Rybach, L., 1986. Amount and significance of radioactive heat sources in sediments. In: Burrus, J. (Ed.), Thermal Modeling in Sedimentary Basins. Technip, Paris, pp. 311322.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: 335344.CrossRefGoogle Scholar
Rybakov, M., and Segev, A., 2004. Top of crystalline basement in Levant. Geochemistry, Geophysics, Geosystems, 5: Q09001.CrossRefGoogle Scholar
Rybár, S., Nemčok, M., Sinha, S. T., et al., 2021. Distribution of coarse-grained syn-rift strata in a transform margin linking with rifted margins: seismic-based study of the Coromandal transform margin, offshore E India. In: Nemčok, M., Doran, H., Doré, A. G., Ledvényiová, L., and Rybár, S. (Eds.), Tectonic Development, Thermal History and Hydrocarbon Habitat Models of Transform Margins, and Their Differences from Rifted Margins. Geological Society of London, London.Google Scholar
Ryseth, A., Augustson, J. H., Charnock, M., et al., 2003. Cenozoic stratigraphy and evolution of the Sørvestsnaget Basin, southwestern Barents Sea. Norwegian Journal of Geology, 83: 107130.Google Scholar
Saemundsson, K., 1974. Evolution of the axial rifting zone in Northern Iceland and the Tjörnes fracture zone. Geological Society of America Bulletin, 85: 495504.2.0.CO;2>CrossRefGoogle Scholar
Sage, F., 1994. Crustal structure of the transform margin and adjacent oceanic domain: example from the Cote d’Ivoire–Ghana margin [in French]. PhD Thesis, University of Pierre and Marie Curie, Paris.Google 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. Geophysical Journal International, 143(3): 662678.CrossRefGoogle Scholar
Sakai, H., Funaki, M., Sato, T., et al., 1997. Paleomagnetic study with 40Ar/39Ar dating of Rajmahal Hills and Mahanadi Graben in India: reconstruction of Gondwanaland. Chishitsugaku Zasshi = Journal of the Geological Society of Japan, 103(3): 192202.Google Scholar
Sales, J. K., 1997. Seal strength vs. trap closure: a fundamental control on the distribution of oil and gas. In: Surdam, R. C. (Ed.), Seals, Traps, and the Petroleum System. AAPG, Washington, DC, pp. 5783.Google Scholar
Sanderson, D. J., and Marchini, W. R. D., 1984. Transpression. Journal of Structural Geology, 6: 449458.CrossRefGoogle Scholar
Sandwell, D. T., 1984. Thermomechanical evolution of oceanic fracture zones. Journal of Geophysical Research, 89(B13): 1140111413.CrossRefGoogle Scholar
Sandwell, D. T., and Schubert, G., 1982. Geoid height–age relation from SEASAT altimeter profiles across the Mendocino Fracture Zone. Journal of Geophysical Research, Solid Earth 87(B5): 39493958.CrossRefGoogle 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(B5): 1003910054.CrossRefGoogle Scholar
Sant, K., Kuipfer, K. F., Rybár, S., et al., 2020. 40Ar/39Ar geochronology using high sensitivity mass spectrometry: examples from middle Miocene horizons of the Central Paratethys. Geologica Carpathica, 71(2): 166182.CrossRefGoogle Scholar
Sant, K., Palcu, D. V., Turco, E., et al., 2019. The mid-Langhian flooding in the eastern Central Paratethys: integrated stratigraphic data from the Transylvanian Basin and SE Carpathian Foredeep. International Journal of Earth Sciences. DOI: 10.1007/s00531-019-01757-zCrossRefGoogle Scholar
Santos Neto, E. V., 1999, Use of hydrogen and carbon stable isotopes characterizing oils from the Potiguar Basin (onshore), northeastern Brazil: AAPG Bulletin, 83: 496518.Google Scholar
Santos Neto, E. V., Mello, M. R., and Rodrigues, R., 1990. Caracterizacao geoquimica dos oleos da Bacia Potiguar. In: XXXVI Congresso Brasileiro de Geologia, vol. 2, pp. 974985.Google Scholar
Šarinová, K., Rybár, S., Halásová, E., et al., 2018. Integrated biostratigraphical, sedimentological and provenance analyses with implications for lithostratigraphic ranking: the Miocene Komjatice depression of the Danube Basin. Geologica Carpathica, 69(4): 382409.CrossRefGoogle Scholar
Sarkar, M., Corfu, F., Paul, D. K., et al., 1993. Early Archean crust in Bastar Craton, Central India: a geochemical and isotopic study. Precambrian Research, 62: 127137.CrossRefGoogle Scholar
Şaroğlu, F., 1988. Age and offset of the North Anatolian Fault. METU Journal of Pure and Applied Sciences, 21: 6579.Google Scholar
Sarwar, G., and DeJong, K. A., 1979. Arcs, oroclines, syntaxis: the curvatures of mountain belts in Pakistan. In: Farah, A., and DeJong, K. A. (Eds.), Geodynamics of Pakistan. Geological Survey of Pakistan, Balochistan, pp. 342349.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
Saunders, J. B., Edgar, N. T., Donnelly, T. W., and Hay, W. W., 1973. Cruise Synthesis. Initial Reports of the Deep Sea Drilling Project 15. US Government Printing Office, Washington, DC, pp. 10771113.Google Scholar
Sawyer, J. H., 1975. Latin America after 1920. In: Owen, E.W. (Ed.), Trek of the Oil Finders: a History of Exploration for Petroleum. AAPG, Washington, DC, pp. 9601251.Google Scholar
Sayers, J., Borissova, I., Ramsay, D., et al., 2002. Geological framework of the Wallaby Plateau and adjacent areas. Record-Australian Geological Survey Organisation, Report No. 2002/021.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: 103134.CrossRefGoogle Scholar
Schattner, U., and Weinberger, R., 2008. A mid-Pleistocene deformation transition in the Hula basin, northern Israel: implications for the tectonic evolution of the Dead Sea Fault. Geochemistry, Geophysics, Geosystems, 9. DOI: 10.1029/2007GC001937CrossRefGoogle Scholar
Scheck, M., and Bayer, U. 1999. Evolution of the Northeast German Basin: inferences from a 3D structural model and subsidence analysis. Tectonophysics, 313: 145169.CrossRefGoogle Scholar
Schenk, C. J., Higley, D. K., and Magoon, L. B., 2003. Region 6 Assessment Summary: Central and South America. US Geological Survey, Reston.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: 137–134.CrossRefGoogle Scholar
Schenk, H. J., Horsfield, B., Krooss, B., Schaefer, R. G., and Schwochau, K., 1997. Kinetics of petroleum formation and cracking. In: Welte, D. H., Horsfield, B., and Baker, D. R. (Eds.), Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling. Springer, Berlin, pp. 233269.Google Scholar
Scherer, E. E., Cameron, K. L., and Blichert-Toft, J., 2000. Lu-Hf garnet geochronology: closure temperature relative to the Sm-Nd system and the effects of trace mineral inclusions. Geochimica et Cosmochimica Acta, 64: 34133432.CrossRefGoogle Scholar
Schettino, A., and Turco, E., 2009. Breakup of Pangaea and plate kinematics of the Central Atlantic and Atlas regions. Geophysics, 178: 10781097.Google Scholar
Scheuber, E., and Andriessen, P. M., 1990. The kinematics significance of the Atacama Fault zone, northern, Chile. Journal of Structural Geology, 21: 243257.CrossRefGoogle Scholar
Schiffries, C. M., 1990. Liquid-absent aqueous fluid inclusions and phase equilibria in the system CaCl2–NaCl–H2O. Geochimica et Cosmochimica Acta, 54: 611619.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(B5): 1054310569.CrossRefGoogle Scholar
Schlanger, S. O., and Jenkyns, H. C., 1976. Cretaceous oceanic anoxic events: causes and consequences. Geologie en Mijnbouw, 55: 179184.Google 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: L16313.CrossRefGoogle Scholar
Schlische, R. W., 1993. Anatomy and evolution of the Triassic–Jurassic continental rift system, eastern North America. Tectonics, 12: 10261042.CrossRefGoogle Scholar
Schmid, H. P., Harzhauser, M., and Kroh, A., 2001. Hypoxic events in a Middle Miocene carbonate platform of the Central Paratethys (Austria, Badenian, 14 Ma). Annalen des Naturhistorischen Museums in Wien, 102A: 150.Google Scholar
Schmid, S. M., and Casey, M., 1986. Complete fabric analysis of some commonly observed quartz c-axis patterns. In: Hobbs, B., and Heard, H. C. (Eds.), Mineral and Rock Deformation: Laboratory Studies – The Patterson Volume. AGU, Washington, DC, pp. 246261.Google Scholar
Schmidt, J., Hacker, B., Ratschbacher, L., et al., 2011. Cenozoic deep crust in the Pamir. Earth and Planetary Science Letters, 312: 411421.CrossRefGoogle Scholar
Schmoker, J. W., and Halley, R. B., 1982. Carbonate porosity versus depth: a predictable relation for South Florida. AAPG Bulletin, 66: 25612570.Google Scholar
Schmucker, U., 1969. Conductivity anomalies, with special reference to the Andes. In: Runcorn, S. K. (Ed.), The Application of Modern Physics to the Earth and Planetary Interiors. Wiley-Interscience, London, pp. 125138.Google Scholar
Schoell, M., 1980. The hydrogen and carbon isotopic composition of methane from natural gases of various origins. Geochimica et Cosmochimica Acta, 44: 649661.CrossRefGoogle Scholar
Scholl, D. W., 1999. East to west transition from orthogonal to oblique, to strike-slip, to orthogonal subduction along the northern rim of the Pacific Basin. Paper presented at the Penrose Conference on Subduction to strike-slip transitions on plate boundaries. GSA, Puerta Plata, Dominican Republic.Google Scholar
Scholz, C. H., 1968. Microfracturing and the inelastic deformation of rock in compression. Journal of Geophysical Research, 73: 14171432.CrossRefGoogle Scholar
Scholz, C. H., Beavan, J., and Hanks, T. C., 1979. Frictional metamorphism, argon depletion, and tectonic stress on the Alpine Fault, New Zealand. Journal of Geophysical Research: Solid Earth, 84(B12): 67706782.CrossRefGoogle Scholar
Schon, J., 1983. Petrophysik: Physikalische Eigenschaften von Gesteinen und Mineralen. Ferdinand Enke Verlag, Stuttgart.CrossRefGoogle Scholar
Schouten, H., and White, R., 1980. Zero offset fracture zones. Geology, 8: 175179.2.0.CO;2>CrossRefGoogle Scholar
Schowalter, T. T. 1979. Mechanics of secondary hydrocarbon migration and entrapment. AAPG Bulletin, 63: 723760.Google Scholar
Schrauder, M., and Navon, O., 1993. Solid carbon dioxide in a natural diamond. Nature, 365: 4244.CrossRefGoogle Scholar
Schretter, I., Zamolyi, A., and Finsterwalder, R., 2020. Geological re-evaluation of the mature reservoirs of the Hochleiten-Pirawarth fields (Vienna Basin). Search and Discovery, 20493.CrossRefGoogle Scholar
Schubert, C., 1980. Late-Cenozoic pull-apart basins, Bocono fault zone, Venezuelan Andes. Journal of Structural Geology, 2(4): 463468.CrossRefGoogle Scholar
Schumacher, B. A., 2002. Methods for the determination of total organic carbon (TOC) in soils and sediments. US EPA.Google Scholar
Sclater, J. G., Grindlay, N. R., Madsen, J. A., and Rommevaux-Jestin, C., 2005. Tectonic interpretation of the Andrew Bain transform fault: southwest Indian Ocean. Geochemistry, Geophysics, Geosystems, 6(9). DOI: 10.1029/2005GC000951CrossRefGoogle Scholar
Scoates, J. S., and Chamberlain, K. R., 1995. Baddeleyite (ZrO2) and zircon (ZrSiO4) from anorthositic rocks of the Laramie anorthosite complex, Wyoming; petrologic consequences and U–Pb ages. American Mineralogist, 80: 13171327.CrossRefGoogle Scholar
Scotese, C. R., 1998. Quicktime computer animations. PALEOMAP Project, Department of Geology, University of Texas at Arlington, Arlington.Google Scholar
Searle, R. C., 1983. Multiple, closely spaced transform faults in fast-slipping fracture-zones. Geology, 11: 607610.2.0.CO;2>CrossRefGoogle Scholar
Searle, R. C., and Laughton, A. S., 1981. Fine-scale sonar study of tectonics and volcanism on the Reykjanes ridge. Oceanologica Acta, 4: 513.Google Scholar
Searle, R. C., Fujioka, H., Cannat, M., et al., 1999. FUJI Dome, a large detachment fault near 64° E on the very slow spreading South West Indian Ridge. AGU Eos Transactions 80: F956.Google Scholar
Secor, D. T., 1965. Role of fluid pressure in jointing. American Journal of Science, 263: 633646.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: 15991617.CrossRefGoogle Scholar
Segall, M. P., and Pollard, D. D., 1980. Mechanics of discontinuous faults. Journal of Geophysical Research, 85: 43374350.CrossRefGoogle Scholar
Sempere, J.-C., Lin, J., Brown, H., Schouten, H., and Purdy, G. M., 1993. Segmentation and morphotectonic variations along a slow-spreading center: the Mid-Atlantic Ridge (24°00′N-30°40′N). Marine Geophysical Researches, 15(3): 153200.CrossRefGoogle Scholar
Sen, S., and Yillar, S., 2008. Is the largest petroleum trap of the world in NW Turkey: Korudag Anticlinorium in the South Trace Basin? Search and Discovery, 70058.Google Scholar
Sengör, A. M. C., 1979. The North Anatolian transform fault: its age, offset and tectonic significance. Journal of the Geological Society of London., 136: 269282.CrossRefGoogle Scholar
Şengör, A. M. C., Görür, N., and Şaroğlu, F., 1985. Strike-slip faulting and related basin formation in zones of tectonic escape: Turkey as a case study. In: Biddle, K. T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation, and Sedimentation. SEPM, Tulsa, pp. 227264.CrossRefGoogle Scholar
Şengör, A. M. C., and Yılmaz, Y., 1981. Tethyan evolution of Turkey: a plate tectonic approach. Tectonophysics, 75: 181241.CrossRefGoogle Scholar
Şengör, A. M. C., Tuysuz, O., Imren, C., et al., 2005. The North Anatolian Fault: a new look. Annual Reviews in Earth and Planetary Sciences, 33: 37112.CrossRefGoogle Scholar
Seton, M., Müller, R. D., Zahirovic, S., et al., 2012. Global continental and ocean basin reconstructions since 200 Ma. Earth-Science Reviews, 113: 212270.CrossRefGoogle Scholar
Shah, F., 2010. Centrifuge modelling study of contrasting structural styles in the Salt Range and the Potwar Plateau, Pakistan. MS Thesis. Queen’s University, Kingston, Ontario.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. Jr., 1983. Permeability controls on aquathermal pressuring. AAPG Bulletin, 67: 20572061.Google Scholar
Shaw, R. K., Arima, M., Kagami, H., Fanning, C. M., and Motoyoshi, Y., 1997. Proterozoic events in the Eastern Ghats Granulite Belt, India: evidence from Rb-Sr, Sm-Nd systematics and SHRIMP dating. Journal of Geology, 105: 645656.CrossRefGoogle 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: 1797717993.CrossRefGoogle Scholar
Sheffels, B., and McNutt, M., 1986. Role of subsurface loads and regional compensation in the isostatic balance of the Transverse Ranges, California: evidence for intracontinental subduction. Journal of Geophysical Research, Solid Earth, 91: 64196431.CrossRefGoogle Scholar
Shelley, D., and Bossiere, G., 2001 The Ancenis Terrane: an exotic duplex in the Hercynian Belt of Armorica. Journal of Structural Geology, 23: 15971614.CrossRefGoogle Scholar
Shemenda, A. I., and Grocholsky, A. L., 1994. Physical modeling of slow seafloor spreading. Journal of Geophysical Research, 99: 91379153.CrossRefGoogle Scholar
Shen, Y., and Forsyth, D. W., 1992, The effects of temperature- and pressure-dependent viscosity on three-dimensional passive flow of the mantle beneath a ridge-transform system. Journal of Geophysical Research, 97: 1971719719.CrossRefGoogle Scholar
Shepard, F., and Emery, K. O., 1941. Submarine topography off the California Coast: Canyons and tectonic interpretations. GSA Special Paper 31.Google Scholar
Shi, J. Y., Mackenzie, A. S., Alexander, R., et al., 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: 21532162.CrossRefGoogle Scholar
Shi, Y., and Wang, C., 1987. Two-dimensional models of the P–T–t paths of regional metamorphism in simple overthrust terrains. Geology, 15(11): 10481051.2.0.CO;2>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: 320328.Google Scholar
Shock, E. L., and Holland, M. E., 2004. Geochemical energy sources that support the subsurface biosphere. In: Wilcock, W. S. D., DeLong, E. F., Kelley, D. S., Baross, J. A., and Cary, S. C. (Eds.), The Sub-Seafloor Biosphere at Mid-Ocean Ridges. AGU, Washington, DC, pp. 153165.CrossRefGoogle Scholar
Shulman, H., Reshef, M., and Ben-Avraham, Z., 2004. The structure of the Golan Heights and its tectonic linkage to the Dead Sea Transform and the Palniyrides folding. Israel Journal of Earth Science, 53: 225237.CrossRefGoogle Scholar
Sibson, R. H., 1983. Continental fault structure and the shallow earthquake source. Journal of the Geological Society of London, 140: 741767.CrossRefGoogle Scholar
Sibson, R. H., 1986. Brecciation processes in fault zones: inferences from earthquake rupturing. Pure and Applied Geophysics, 124: 159175.CrossRefGoogle Scholar
Sibson, R. H., 1990. Faulting and fluid flow. In: Nesbitt, B. E. (Ed.), Fluids in Tectonically Active Regimes of the Continental Crust. Mineralogical Association of Canada, Quebec City, pp. 93132.Google Scholar
Sibson, R. H., 1996. Structural permeability of fluid-driven fault-fracture meshes. Journal of Structural Geology, 18: 10311042.CrossRefGoogle Scholar
Sibson, R. H., Robert, F., Poulsen, K. H., 1988. High-angle reverse faults, fluid-pressure cycling, and mesothermal gold-quartz deposits. Geology, 16: 551555.2.3.CO;2>CrossRefGoogle Scholar
Sibuet, J. C., and Mascle, J., 1978. Plate kinematic implications of Atlantic equatorial fracture zone trends. Journal of Geophysical Research, 83(B7): 34013421.CrossRefGoogle Scholar
Sibuet, J. C., Srivastava, S., and Manatschal, G., 2007. Exhumed mantle-forming transitional crust in the Newfoundland–Iberia rift and associated magnetic anomalies. Journal of Geophysical Research, 112: B06105.CrossRefGoogle Scholar
Siddiqui, F. I., 1996. A dynamic theory of hydrocarbon migration and trapping. PhD Thesis. University of Texas Austin.CrossRefGoogle Scholar
Siddiqui, F. I., and Lake, L. W., 1997. A comprehensive dynamic theory of hydrocarbon migration and trapping. In: Proceedings of the SPE 72nd Annual Technical Conference, San Antonio, 5–8 October 1997. Society of Petroleum Engineers, Richardson, pp. 395410.Google Scholar
Siedl, W., Strauss, P., Sachsenhofer, R. F., et al., 2020. Revised Badenian (middle Miocene) depositional systems of the Austrian Vienna Basin based on a new sequence stratigraphic framework. Austrian Journal of Earth Sciences, 113(1): 87110.CrossRefGoogle Scholar
Sigmundsson, F., Einarsson, P., Bilham, R., and Sturkell, E., 1995. Rift-transform kinematics in south Iceland: deformation from Global Positioning System measurements, 1986 and 1992. Journal of Geophysical Research, 100: 62356248.CrossRefGoogle Scholar
Sikora, P. J., 2004. West African chronostratigraphy. In: Nemčok, M., Rosendahl, B. R., Segall, M., et al. (Eds.), 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. EGI Archive, Salt Lake City.Google Scholar
Sikora, P., Segall, M., Stuart, C., et al., 2004. 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. EGI Archive, Salt Lake City.Google Scholar
Sills, S., and Agyapong, D., 2012. Jubilee field reservoir description & waterflood performance overview. Paper presented at the Offshore Technology Conference, 30 April–3 May, Houston, Texas.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(9): 961964.CrossRefGoogle Scholar
Simancas, J. F., Carbonnell, R., Gonzalez Lodeiro, F., et al. 2003. Crustal structure of the transpression Variscan orogen of SW Iberia: SW Iberia deep seismic reflection profile (IBERSEIS). Tectonics, 22: 19.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(5): 533554.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: 609622.CrossRefGoogle Scholar
Singh, R. P., Rawat, S., and Chandra, K., 1999. Hydrocarbon potential in India deep waters. Exploration Geophysics, 30: 8395.CrossRefGoogle Scholar
Sinha, S. T., Nemčok, M., Choudhuri, M., Sinha, N., and Pundarika Rao, D., 2016. The role of breakup localization in microcontinent separation along a strike-slip margin: East India–Elan Bank case study. In: Nemčok, M., Rybár, S., Sinha, S. T., Hermeston, S. A., and Ledvényiová, L., (Eds.), Transform Margins: Development, Controls and Petroleum Systems. Geological Society of London, London, pp. 95123.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. AAPG Bulletin, 92: 513547.CrossRefGoogle Scholar
Skelton, A. D. L., Whitmarsh, R., Arghe, F., Crill, P., and Koyi, H., 2005. Constraining the rate and extent of mantle serpentinization from seismic and petrological data: implications for chemosynthesis and tectonic processes. Geofluids, 5: 153164.CrossRefGoogle Scholar
Skempton, A. W., 1985. Residual strength of clays in landslides, folded strata and the laboratory. Geotechnique, 35: 318.CrossRefGoogle Scholar
Skene, K. I., and Piper, D. J. W., 2006, Late Cenozoic evolution of Laurentian Fan: development of a glacially-fed submarine fan: Marine Geology, 227: 6792.CrossRefGoogle Scholar
Slater, R. A., Gorsline, D. S., Kolpack, R. L., and Shiller, G. I., 2002, Post-glacial sediments of the Californian shelf from Cape San Martin to the US–Mexico border. Quaternary International, 92: 4561.CrossRefGoogle Scholar
Sleep, N., and Barth, G. A., 1997. The nature of the lower crust and shallow mantle emplacement at low spreading rates. Tectonophysics, 279: 181191.CrossRefGoogle Scholar
Sleep, N. H., and Bird, D. K., 2007. Niches of the pre-photosynthetic biosphere and geologic preservation of Earth’s earliest ecology. Geobiology, 5(2): 101117.CrossRefGoogle Scholar
Small, C., 1998. Global systematic of mid-ocean ridge morphology. In: Buck, W. R., Delaney, P. T., Karson, J. A., and Lagabrielle, Y. (Eds.), Faulting and Magmatism at Mid-Ocean Ridges. AGU, Washington, DC, pp. 126.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: 2288522902.CrossRefGoogle Scholar
Smit, M. A., Ratschbacher, L., Kooijman, E., and Stearns, M. A., 2017. Early evolution of the Pamir deep crust from Lu–Hf and U–Pb geochronology and garnet thermometry. Geology, 42(12): 10471050.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. In: Newman, S. P., and Neretnieks, I. (Eds.), Hydrogeology of Low Permeability Environments. A A Balkema, Amsterdam, pp. 4167.Google Scholar
Smith, P. E., Farquhar, R. M., and Hancock, R. G., 1991. Direct radiometric age determination of carbonate diagenesis using U–Pb in secondary calcite. Earth and Planetary Science Letters, 105: 474491.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: 951970.CrossRefGoogle Scholar
Smith, W. H. F., and Sandwell, D. T., 1997. Seafloor topography from satellite altimetry and ship depth soundings. Science, 277(5334): 19561962.CrossRefGoogle Scholar
Sneh, A., and Weinberger, R., 2014. Major Geological Structures of Israel and Environs. Geological Survey of Israel, Jerusalem.Google Scholar
Snow, J. E., and Dick, H. J. B., 1955. Pervasive magnesium loss by weathering of peridotite. Geochimica et Cosmochimica Acta, 59: 42194235.CrossRefGoogle Scholar
Snowdon, L. R., 1991. Oil from type III organic matter: resinite revisited. Organic Geochemistry, 17: 743747.CrossRefGoogle Scholar
Sobolev, S. V., Petrunin, A., Garfunkel, Z., Babeyko, A. Y., and the Desert Group, 2005. Thermo-mechanical model of the Dead Sea Transform. Earth and Planetary Science Letters, 238(1–2): 7895.CrossRefGoogle Scholar
Société Nationale d’Operations Pétrolières de la Côte d’Ivoire (PETROCI), 2009. Deepwater Côte d’Ivoire potential. Presentation at Africa Upstream 2009, Cape Town.Google Scholar
Souza, J. M., 1982. Transmission of seismic energy through the Brazilian Parana Basin layered basalt stack. Master’s thesis. University of Texas at Austin, Austin.Google Scholar
Sparks, D. W., and Parmentier, E. M., 1993. The structure of three-dimensional convection beneath spreading centers. Geophysical Journal International, 112: 8191.CrossRefGoogle Scholar
Sperner, B., Ratschbacher, L., and Nemčok, M., 2002. Interplay between subduction retreat and lateral extrusion: tectonics of the Western Carpathians. Tectonics, 21(6): 10511075.CrossRefGoogle Scholar
Špička, V., 1969. Analysis of thickness, areal extent and evolution of the Neogene strata in the Vienna Basin region [in Czech]. Zborník geologických vied, Západné Karpaty, 11: 128155.Google Scholar
Spiers, C. J., and Schutjens, P. M. T., 1990. Densification of crystalline aggregates by fluid phase diffusional creep. In: Barber, D. J., and Meredith, P. G. (Eds.), Deformation Processes in Minerals, Ceramics and Rock. Springer, New York, pp. 334352.CrossRefGoogle Scholar
Srivastava, D. C., and Engelder, T., 1990. Crack-propagation sequence and pore-fluid conditions during fault-bend folding in the Appalachian Valley and Ridge, central Pennsylvania. Geological Society of America Bulletin 102(1): 116128.2.3.CO;2>CrossRefGoogle Scholar
Srivastava, S., Sibuet, J.-C., Cande, S., Roest, W. R., and Reid, I. R., 2000. Magnetic evidence for slow seafloor 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: Vogt, P. R., and Tucholke, B. (Eds.), The Geology of North America, Volume M, The Western North Atlantic Region. Geological Society of America, Boulder, pp. 379404.CrossRefGoogle Scholar
St. John, W., 2000. The role of transform faulting in formation of hydrocarbon traps in the Gulf of Guinea, West Africa. Paper presented at the Offshore West Africa Conference, Abidjan, March, 21–23, 2000.Google Scholar
Staatsolie, 2001. Petroleum investment opportunities in Surinam. Available at: www.staatsolie.com.Google Scholar
Stacey, J. S., and Kramers, J. D., 1975. Approximation of terrestrial lead isotope evolution by a two-stage model. Earth and Planetary Science Letters, 26: 207221.CrossRefGoogle Scholar
Stach, E., Mackowski, M. T., Teichmüller, M., et al., 1982. Stach’s Textbook of Coal Petrology; Third Revised and Enlarged Edition. Borntraeger, Berlin.Google Scholar
Stagg, H. M. J., Alcock, M. B., Bernadel, G., et al., 2004. Geological Framework of the Outer Exmouth Plateau and Adjacent Ocean Basins. Geoscience Australia, Canberra.Google Scholar
Stakes, D. S., Orange, D., Paduan, J. B., Salamy, K. A., and Maher, N. M., 1999. Cold-seeps and authigenic carbonate formation in Monterey Basin, California. Marine Geology, 159: 93109.CrossRefGoogle Scholar
Staniaszek, S., 2019. Transform faults and evolution of the Tjörnes Transform Fault Zone, Northern Iceland. Internship in structural geology of oceanic transforms, unpublished report. EGL Bratislava.Google Scholar
Stearns, M. A., Cottle, J. M., Hacker, B. R., and Kylander-Clark, A. R. C., 2016. Extracting thermal histories from the near-rim zoning in titanite using coupled U/Pb and trace element depth profiles by single shot laser ablation split stream (SS-LASS) ICP-MS. Chemical Geology, 422: 1324.CrossRefGoogle Scholar
Steckler, M. S., 1981. A thermomechanical model of the subsidence of continental margins. Eos, Transactions, American Geophysical Union, 62(17): 390.Google Scholar
Steckler, M. S., 1985. Uplift and extension at the Gulf of Suez: indications of induced mantle convection. Nature (London), 317(6033): 135139.CrossRefGoogle Scholar
Steckler, M. S., Berthelot, F., Lybéris, N., and Le Pichon, X., 1988. Subsidence in the Gulf of Suez: implications for rifting and plate kinematics. Tectonophysics, 153(1–4): 249270.CrossRefGoogle 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: 735738.2.3.CO;2>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(44): 589.Google Scholar
Steel, R., Gjelberg, J., Helland-Hansen, W., et al., 1985. Tertiary strike-slip basins and orogenic belt of Spitsbergen. In: Biddle, K. T., and Christie-Blick, N. (Eds.), Strike-Slip Deformation, Basin Formation, and Sedimentation. SEPM, Tulsa, pp. 339359.CrossRefGoogle Scholar
Stefansson, R., Bergerat, F., Bonafede, M., et al., 2000. PRENLAB project. Final report. Available at: http://hraun.vedur.is/ja/prenlab.Google Scholar
Stefansson, R., Gudmundsson, G. B., and Roberts, M., 2006a. Long-term and short-term earthquake warnings based on seismic information in the SISZ. Icelandic Meteorological Office report 06006. Available at: http://andvari.vedur.is/utgafa/greinargerdir/2006/06006.pdf.Google Scholar
Stefansson, R., Bonafede, M., Roth, F., et al., 2006b. Modelling and parametrizing the SW Iceland earthquake release and deformation process. Icelandic Meteorological Office report 06005. Available at: http://andvari.vedur.is/utgafa/greinargerdir/2006/06005.pdf.Google Scholar
Stefansson, R., Gudmundsson, G. B., and Halldorsson, P., 2008. Tjörnes fracture zone: new and old seismic evidences for the link between the North Iceland rift zone and the Mid-Atlantic ridge. Tectonophysics. 447 (1–4): 117126.CrossRefGoogle Scholar
Stein, C. A., and Stein, S., 1992. A model for the global variation in oceanic depth and heat flow with lithospheric age. Nature, 359: 123129.CrossRefGoogle Scholar
Stein, C. A., and Stein, S., 1994. Constraints on hydrothermal heat flux through the oceanic lithosphere from global heat flow. Journal of Geophysical Research, 99: 30813095.CrossRefGoogle Scholar
Stein, R., Rullkötter, J., and Welte, D. H., 1986. Accumulation of organic carbon-rich sediments in the late Jurassic and Cretaceous Atlantic Ocean: a synthesis. Chemical Geology, 56:132.CrossRefGoogle Scholar
Stel, H., 1985. Crystal growth in cataclasites: diagnostic microstructures and implications. Tectonophysics, 78: 585600.CrossRefGoogle Scholar
Stephenson, L. P., 1977. Porosity dependence on temperature: limits on maximum possible effect. AAPG Bulletin, 61: 407415.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(1): 3155.CrossRefGoogle Scholar
Stern, T. A., and McBride, J. H., 1998. Seismic exploration of continental strike-slip zones. Tectonophysics, 286(1): 6378.CrossRefGoogle Scholar
Stern, T., Molnar, P., Okaya, D., and Eberhart-Phillips, D., 2000. Teleseismic P wave delays and models of shortening the mantle lithosphere beneath South Island, New Zealand. Journal of Geophysical Research, Solid Earth, 105: 2161521631.CrossRefGoogle Scholar
Stern, T., Kleffmann, S., Okaya, D., Scherwath, M., and Bannister, S., 2001. Low seismic-wave speeds and enhanced fluid pressure beneath the Southern Alps of New Zealand. Geology, 29(8): 679682.2.0.CO;2>CrossRefGoogle Scholar
Stern, T., Okaya, D., Kleffmann, S., et al., 2007. Geophysical exploration and dynamics of the Alpine Fault Zone. In: Okaya, D., Stern, T., and Davey, F. (Eds.), A Continental Plate Boundary: Tectonics at South Island, New Zealand. AGU, Washington, DC, pp. 207233.CrossRefGoogle Scholar
Stewart, S. A., and Reeds, A., 2003. Geomorphology of kilometer-scale extensional fault scarps: factors that impact seismic interpretation. AAPG Bulletin, 87(2): 251272.CrossRefGoogle Scholar
Stirling, E., Fugelli, E., and Thompson, M., 2017. The edges of the wedges: a systematic approach to trap definition and risking for stratigraphic, combination and sub-unconformity traps. In: Bowman, M., and Levell, B. (Eds.), Petroleum Geology of NW Europe: 50 Years of Learning. Proceedings of the 8th Petroleum Geology Conference, London, pp. 273–286.CrossRefGoogle Scholar
Stitt, L. T., 1986. Structural history of the San Gabriel Fault and other Neogene structures of the central Transverse Ranges. In: Ehlig, P. (Ed.), Neotectonics and Faulting in Southern California (Cordilleran Section Guidebook and Volume). Geological Society of America, Boulder, pp. 43102.Google Scholar
Stockli, D. F., Dumitru, T. A., McWilliams, M. O., and Farley, K. A., 2003. Cenozoic tectonic evolution of the White Mountains, California and Nevada. Geological Society of America Bulletin, 115(7): 788816.2.0.CO;2>CrossRefGoogle Scholar
Stoffers, P., Botz, R., Garbe-Schonberg, D., et al., 1997. Cruise Report Poseidon 229a, Kolbeinsey Ridge. Geologisch-Palaontologisches Institut, Kiel.Google Scholar
Stoll, R. D., and Bryan, G. M., 1979. Physical properties of sediments containing gas hydrates. Journal of Geophysical Research, 84: 16291634.CrossRefGoogle Scholar
Stone, D. S., 1995. Structure and kinematic genesis of the Quealy wrench duplex: transpressional reactivation of the Precambrian Cheyenne Belt in the Laramie Basin, Wyoming. AAPG Bulletin, 79: 13491376.Google Scholar
Stráník, Z., Adámek, J., and Cipris, V., 1979. Geological profile through the Carpathian foredeep, Flysch Belt and Vienna Basin in the Pavlov Hills region [in Czech]. In: Maheľ, M. (Ed.), Tectonic Profiles of West Carpathians [in Slovak]. GÚDŠ, Bratislava, pp. 715.Google Scholar
Stráník, Z., Dvořák, J., Krejčí, O., et al., 1993. The contact of the North European Epivariscan Platform with the West Carpathians. Journal of the Czech Geological Society, 38(102): 2130.Google Scholar
Strauss, P., Harzhauser, M., Hinsch, R., and Wagreich, M., 2006. Sequence stratigraphy in a classic pull-apart basin (Neogene, Vienna Basin): a 3D seismic based integrated approach. Geologica Carpathica, 57: 185197.Google Scholar
Strehlau, J. 2006. Earthquakes and nonvolcanic tremor in the lower continental crust: open questions. EGU annual meeting, April 2–7, Vienna.Google Scholar
Stuart, C. J., Nemčok, M., Vangelov, D., et al., 2011. Structural and depositional evolution of the Eastern Balkan thrustbelt, Bulgaria. AAPG Bulletin, 95(4): 649673.CrossRefGoogle Scholar
Stubner, K., Ratschbacher, L., Rutte, D., et al., 2013a. The giant Shakhdara magmatic gneiss dome, Pamir, India-Asia collision zone, I: Geometry and kinematics. Tectonics, 32: 948979.CrossRefGoogle Scholar
Stubner, K., Ratschbacher, L., Weise, C., et al., 2013b. The giant Shakhdara magmatic gneiss dome, Pamir, India-Asia collision zone, II: Timing of dome formation. Tectonics, 32: 14041431.CrossRefGoogle Scholar
Stuiver, M., Reimer, P. J., and Braziunas, T. F., 1998. High-precision radiocarbon age calibration for terrestrial and marine samples. Radiocarbon, 40(3): 11271151.CrossRefGoogle Scholar
Sturm, R., 2012. Internal morphology and crystal growth of accessory zircon from igneous rocks. In: Van Dijk, G., and Van den Berg, V. (Eds.), Zircon and Olivine: Characteristics, Types and Uses. Nova Science Publishers, Hauppauge, pp. 3766.Google Scholar
Stüwe, K., 2002. Geodynamics of the Lithosphere: An Introduction. Springer, Berlin.CrossRefGoogle Scholar
Stüwe, K., 2013. Geodynamics of the Lithosphere, 2nd edition. Springer, Berlin.Google 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
Subrahmanyam, C., Thakur, N. K., Gangadhara Rao, T., et al., 1999. Tectonics of the Bay of Bengal; new insights from satellite-gravity and ship-borne geophysical data. Earth and Planetary Science Letters, 171(2): 237251.CrossRefGoogle Scholar
Summerfield, M. A., 1985. Plate tectonics and landscape development on the African continent. In: Morisawa, M., and Hack, J. T. (Eds.), Tectonic Geomorphology. Allen and Unwin, Concord, pp. 2751.Google Scholar
Summerfield, M. A., 2000. Geomorphology and global tectonics: introduction. In: Summerfield, M. A. (Ed.), Geomorphology and Global Tectonics. Wiley, New York, pp. 1527.Google Scholar
Sutherland, R., 1995. Late Cenozoic Tectonics in the SW Pacific, and Development of the Alpine Fault Through Southern South Island. University of Otago, Dunedin.Google Scholar
Sutherland, R., Davey, F., and Beavan, J., 2000. Plate boundary deformation in South Island, New Zealand, is related to inherited lithospheric structure. Earth and Planetary Science Letters, 177(3–4): 141151.CrossRefGoogle Scholar
Sutter, J. F., 1988. Innovative approaches to the dating of igneous events in the early Mesozoic basins of the eastern U.S. In: Froelich, A. J., and Robinson, G. R., Jr. (Eds.), Studies of the Early Mesozoic Basins of the Eastern U.S. US Geological Survey, Reston, pp. 194200.Google Scholar
Swanson, M. T., 1988. Pseudotachylite-bearing strike-slip duplex structures in the Fort Foster Brittle Zone, S. Maine. Journal of Structural Geology, 10: 813828.CrossRefGoogle Scholar
Swanson, M. T., 1989. Sidewall ripouts in strike-slip faults. Journal of Structural Geology, 11: 933948.CrossRefGoogle Scholar
Swanson, M. T., 2005. Geometry and kinematics of adhesive wear in brittle strike-slip fault zones. Journal of Structural Geology, 27: 871887.CrossRefGoogle Scholar
Sweeney, J. J., and Burnham, A. K., 1990. Evaluation of a simple model of vitrinite reflectance based on chemical kinetics. AAPG Bulletin, 74: 15591570.Google Scholar
Sylvester, A. G., 1988. Strike-slip faults. Geological Society of America Bulletin, 100(11): 16661703.2.3.CO;2>CrossRefGoogle Scholar
Sylvester, A. G., and Smith, R. R., 1976. Tectonic transpression and basement-controlled deformation in San Andreas fault zone, Salton Trough, California. AAPG Bulletin, 60(12): 20812102.Google Scholar
Szalay, A., and Koncz, I., 1993. Migration and accumulation of oil and natural gas generated from Neogene source rocks in the Hungarian part of the Pannonian Basin. In: Spencer, A. M. (Ed.), Generation, Accumulation and Production of Europe’s Hydrocarbons III. Springer, Berlin, pp. 303309.CrossRefGoogle Scholar
Szatmari, P., 2000. Habitat of petroleum along the South Atlantic margins. In: Mello, M. R., and Katz, B. J. (Eds.), Petroleum Systems of South Atlantic Margins. AAPG, Washington, DC, pp. 6975.Google Scholar
Szatmari, P., and Milani, E. J., 1999. Microplate rotation in northeastern Brazil during South Atlantic rifting: analogies with the Sinai microplate. Geology, 27: 11151118.2.3.CO;2>CrossRefGoogle Scholar
Tainish, H. R., Stringer, K. V., and Azad, J., 1959. Major gas fields of West Pakistan. American Association of Petroleum Geologist Bulletin, 43: 26752700.Google Scholar
Tamaki, K., Suyehiro, K., Allan, J., Ingle, J. C., and Pisciotto, K., 1992. Tectonic synthesis and implications of Japan Sea ODP drilling. In: Proceedings of the Ocean Drilling Program: Scientific Results, Volume 127–128. Texas A&M University, College Station, pp. 13331350.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. In: Tankard, A. J., and Balkwill, H. R. (Eds.), Extensional Tectonics and Stratigraphy of the North Atlantic Margins. AAPG, Washington, DC, pp. 265282.CrossRefGoogle Scholar
Tari, G. C., and Horváth, F., 2006. Alpine evolution and hydrocarbon geology of the Pannonian Basin: an overview. In: Golonka, J., and Pícha, F. J. (Eds.), The Carpathians and Their Foreland: Geology and Hydrocarbon Resources. AAPG, Washington, DC, pp. 605618.Google Scholar
Tari, G., Horváth, F., and Rumpler, J., 1992. Styles of extension in the Pannonian Basin. Tectonophysics, 208: 203219.CrossRefGoogle Scholar
Tari, G., Molnar, J., Ashton, P., and Hedley, R., 2000. Salt tectonics in the Atlantic margin of Morocco. Leading Edge, 19(10): 10741076.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: AAPG, 2001: 198.Google Scholar
Tari, G., Ashton, P., Coterill, K., et al., 2002, Are West Africa deepwater salt tectonics analogous to the Gulf of Mexico? Oil and Gas Journal, March 4: 1–16.Google Scholar
Tari, G., Molnar, J., and Ashton, P., 2003. Examples of salt tectonics from West Africa: a comparative approach. In: Arthur, T. J., MacGregor, D. S., and Cameron, N. R. (Eds.), Petroleum Geology of Africa: New Themes and Developing Technologies. Geological Society of London, London, pp. 85104.Google Scholar
Tatar, Y., 1975. Tectonic structures along the North Anatolian Fault Zone, northeast of Refahiye (Erzincan). Tectonophysics, 29: 401409.CrossRefGoogle 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: 645660.CrossRefGoogle 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 atlasique. PhD thesis. University of d’Aix-Marseille III.Google Scholar
Taylor, B., Crook, K., and Sinton, J., 1994. Extensional transform zones and oblique spreading centers. Journal of Geophysical Research, 99(B10): 1970719718.CrossRefGoogle Scholar
Taylor, B., Goodlife, A., Martinez, F., and Hey, R. N., 1995. Continental rifting and initial seafloor spreading in the Woodlark Basin. Nature, 374: 534537.CrossRefGoogle Scholar
Taylor, B., Goodlife, A., and Martinez, F., 2009. Initiation of transform faults at rifted continental margins. Comptes Rendus Geoscience, 341: 428438.CrossRefGoogle Scholar
Tchalenko, J. S., 1970. The evolution of kink-bands and the development of compressional textures in sheared clay. Tectonophysics, 6: 159174.CrossRefGoogle Scholar
Tchalenko, J. S., and Ambraseys, N. N., 1970. Structural analyses of the Dasht-e-Bayaz (Iran) earthquake fractures. GSA Bulletin, 81: 4160.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: 543574.CrossRefGoogle Scholar
Teisserenc, P., and Villemin, J., 1990. Sedimentary basin of Gabon: geology and oil systems. In: Edwards, J. D., and Santogrossi, P. A. (Eds.), Divergent/Passive Margin Basins. AAPG, Washington, DC, pp. 117199.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(2): 333350.CrossRefGoogle Scholar
ten Brink, U. S., and Stern, T. A., 1992. Rift flank uplifts and hinterland basins: comparison of the Transantarctic Mountains with the Great Escarpment of Southern Africa. Journal of Geophysical Research, 97(B1): 569585.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: 133144.CrossRefGoogle Scholar
Tesauro, M., Kaban, M. K., and Cloetingh, S. A. P. L., 2012. Global model for the lithospheric strength and effective elastic thickness. Tectonophysics, 602: 7886.CrossRefGoogle Scholar
Tesón, E., Mora, A., Silva, A., et al., 2013. Relationship of Mesozoic graben development, stress, shortening magnitude, and structural style in the Eastern Cordillera of the Colombian Andes. In: Nemčok, M., Mora, A., and Cosgrove, J. W. (Eds.), Thick-Skin-Dominated Orogens: From Initial Inversion to Full Accretion. Geological Society of London, London, pp. 257283.Google Scholar
Tetreault, J. L., and Buiter, S. J. H., 2018. The influence of extension rate and crustal rheology on the evolution of passive margins from rifting to break-up. Tectonophysics, 746(30): 155172.CrossRefGoogle Scholar
Thomas, J. B., Bodnar, R. J., Shimizu, N., and Sinha, A. K., 2002. Determination of zircon/melt trace element partition coefficients from SIMS analysis of melt inclusions in zircon. Geochimica et Cosmochimica Acta, 66: 28872902.CrossRefGoogle Scholar
Thomas, M. M., and Clouse, J. A., 1995. Scaled physical model of secondary oil migration. AAPG 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: 625694.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(5): 526538.CrossRefGoogle Scholar
Thompson, K. F. M., 1988. Gas-condensate migration and oil fractionation in deltaic systems. Marine and Petroleum Geology, 5: 237246.CrossRefGoogle 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.Google Scholar
Thurber, C., Roecker, S., Ellworth, W. L., et al., 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(13): 15911594.CrossRefGoogle Scholar
Thurber, C., Roecker, S., Roberts, K., et al., 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(3): 1112.CrossRefGoogle Scholar
Thurber, C., Roecker, S., Zhang, H., Baher, S., and Ellsworth, W., 2004. Fine-scale structure of the San Andreas fault zone and location of the SAFOD target earthquakes. Geophysical Research Letters, 31: L12S02.CrossRefGoogle Scholar
Thurow, J., Brumsack, H.-J., Rullkötter, J., Littke, R., and Meyers, P., 1992. The Cenomanian/Turonian boundary event in the Indian Ocean: a key to understand the global picture. In: Duncan, R. A., Rea, D. K., Kidd, R. B., von Rad, U., and Weissel, J.K. (Eds.), Synthesis of Results from Scientific Drilling in the Indian Ocean. AGU, Washington, DC, pp. 253273.Google Scholar
Tiessen, H., and Moir, J. O., 1993. Total and organic carbon. In: Carter, M. E. (Ed.), Soil Sampling and Methods of Analysis. Lewis Publishers, Ann Arbor, pp. 187211.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: 743777.CrossRefGoogle Scholar
Tissot, B. P., and Welte, D. H., 1984. Petroleum Formation and Occurrence, 2nd edition. Springer, New York.CrossRefGoogle Scholar
Tissot, B. P., Pelet, R., and Ungerer, P., 1987. Thermal history of sedimentary basins, maturation indices, and kinetics of oil and gas generation. AAPG Bulletin, 71: 14451466.Google Scholar
Todd, B. J., and Keen, C. E., 1989. Temperature effects and their geological consequences at transform margins. Canadian Journal of Earth Sciences, 26: 25912603.CrossRefGoogle Scholar
Todd, B. J., Reid, I., and Keen, C. E., 1988. Crustal structure across the Southwest Newfoundland transform margin. Canadian Journal of Earth Sciences, 25: 744759.CrossRefGoogle Scholar
Toksoz, M. N., Cheng, C. H., and Timur, A., 1976. Velocities of seismic waves in porous rocks. Geophysics, 41: 621643.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. AAPG Bulletin, 84: 6574.Google Scholar
Tolstoy, M., Harding, A., and Orcutt, J., 1993. Crustal thickness of the Mid-Atlantic Ridge: bulls-eye gravity anomalies and focused accretion. Science, 262: 726729.CrossRefGoogle ScholarPubMed
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(B4): 50315038.CrossRefGoogle Scholar
Torsvik, T. H., Rousse, S., Labails, C., and Smethurst, M. A., 2009. A new scheme for the opening of the South Atlantic Ocean and the dissection of an Aptian salt basin. Geophysical Journal International 177(3): 13151333.CrossRefGoogle Scholar
Toulokian, Y. S., Judd, W. R., and Roy, R. F. (Eds.), 1981. Physical Properties of Rocks and Minerals. McGraw-Hill, New York.Google Scholar
Touret, J. L. R., 2001. Fluid inclusions in metamorphic rocks. Lithos, 55: 125.CrossRefGoogle Scholar
Towle, P., Addis, D., Brown, A., Layman, J., and Mahon, K., 2012. Early opening history of the African transform margin and its influence on Upper Cretaceous deepwater depositional systems in the Deep Ivorian Basin. In: Third Central and North Atlantic Conjugate Margins Conference, Trinity College Dublin, 22–24 August 2012, abstracts volume, pp. 94–97.Google Scholar
Toy, V. G., Prior, D. J., and Norris, R. J., 2008. Quartz fabrics in the Alpine Fault mylonites: influence of pre-existing preferred orientations on fabric development during progressive uplift. Journal of Structural Geology, 30(5): 602621.CrossRefGoogle Scholar
Toy, V. G., Prior, D. J., Norris, R. J., Cooper, A. F., and Walrond, M., 2012. Relationships between kinematic indicators and strain during syn-deformational exhumation of an oblique slip, transpressive, plate boundary shear zone: The Alpine Fault, New Zealand. Earth and Planetary Science Letters, 333: 282292.CrossRefGoogle Scholar
Toy, V. G., Norris, R. J., Prior, D. J., Walrond, M., and Cooper, A. F., 2013. How do lineations reflect the strain history of transpressive shear zones? The example of the active Alpine Fault zone, New Zealand. Journal of Structural Geology, 50: 187198.CrossRefGoogle Scholar
Treiman, A. H., 2012. Eruption age of the Sverrefjellet volcano, Spitsbergen Island, Norway. Polar Research, 31: 17320.CrossRefGoogle Scholar
Treiman, J. E., 1995. Surface faulting near Santa Clarita. In: Woods, M. C., and Seiple, W. R. (Eds.), Northridge, California Earthquake, January 17, 1994. California Department of Conservation, Division of Mines and Geology, Sacramento, pp. 103110.Google Scholar
Trindade, L. A. F., Brassell, S. C., and Santos Neto, E. V. 1992, Petroleum migration and mixing in the Potiguar Basin, Brasil. AAPG Bulletin, 76: 19031924.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. AAPG Bulletin, 86: 10951112.Google Scholar
Tryggvason, E., 1994. Surface deformation at the Krafla volcano, north Iceland, 1982–1992. Bulletin of Volcanology, 56: 98107.CrossRefGoogle Scholar
Tseng, H. Y., Burrus, R. C., Onstott, T. C., and Gomaa, O., 1999. Paleofluid-flow circulation within a Triassic rift basin: evidence from oil inclusions and thermal histories. Geological Society of America Bulletin, 111(2): 275290.2.3.CO;2>CrossRefGoogle Scholar
Tsunongai, U., Yoshida, N., Ishibashi, J., and Gamo, T., 2000. Carbon isotopic distribution of methane in deep-sea hydrothermal plume, Myojin Knoll Caldera, Izu-Bonin arc: implications for microbial methane oxidation in the oceans and applications to heat flux estimation. Geochimica et Cosmochimica Acta, 64: 24392452.CrossRefGoogle Scholar
Tucholke, B. E., and Lin, J., 1994. A geological model for the structure of ridge segments in slow spreading ocean crust. Journal of Geophysical Research, 99: 1193711958.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(B5): 98579866.CrossRefGoogle Scholar
Tucholke, B. E., Fujioka, K., Ishihara, T., Hirth, G., and Kinoshita, M., 2001. Submersible study of an oceanic megamullion in the central North Atlantic. Journal of Geophysical Research 106: 1614516161.CrossRefGoogle Scholar
Tucholke, B. E., and Sibuet, J.-C., 2007. Leg 210 synthesis: tectonic, magmatic, and sedimentary evolution of the Newfoundland–Iberia rift. In: Tucholke, B. E., Sibuet, J.-C., and Klaus, A. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results, 2010. Texas A&M University, College Station, pp. 156.Google Scholar
Tuckwell, G. W., Bull, J. M., and Sanderson, D. J., 1998. Numerical models of faulting of oblique spreading centers. Journal of Geophysical Research, 103: 1547315482.CrossRefGoogle Scholar
Tullow Oil plc., 2008. Tullow Capital Markets Day. A Step change for Tullow and Ghana. October 1.Google Scholar
Tullow Oil plc., 2018. Capital markets day 2018. November 29. Available at: www.tullowoil.com/application/files/7515/7960/2778/tullow-capital-markets-day-2018_presentation.pdf.Google Scholar
Turcotte, D. L., and Schubert, G., 1982. Geodynamics Applications of Continuum Physics to Geological Problems. Wiley, New York.Google Scholar
Turcotte, D. L., and Schubert, G., 2002. Geodynamics, 2nd edition. Cambridge University Press, Cambridge.CrossRefGoogle Scholar
Twiss, R. J., 1977. Theory and applicability of a recrystallized grain size paleopiezometer. Pure and Applied Geophysics PAGEOPH, 115( 1–2): 227244.CrossRefGoogle Scholar
Twiss, R. J., and Moores, E. M., 1992. Structural Geology. W. H. Freeman and Company, New York.Google Scholar
Twiss, R. J., and Moores, E. M., 2007. Structural Geology, 2nd edition. W. H. Freeman and Company, New York.Google Scholar
Tye, R. S., 1991. Fluvial-sandstone reservoirs of the Travis Peak Formation, East Texas Basin. Concepts in Sedimentology and Paleontology, 3: 172188.Google Scholar
Tye, R. S., 2004a. Geomorphology: an approach to determining subsurface reservoir dimensions. AAPG Bulletin, 88: 11231147.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., 1989. Depositional processes and stratigraphy of fluvially dominated lacustrine deltas: Mississippi delta plain. Journal of Sedimentary Petrology, 59: 973996.Google Scholar
UgwuOju, O., 2018. Clinothems of the Cretaceous Berbice Canyon, offshore Guyana. MS thesis. Colorado School of Mines, Golden.Google Scholar
Ulmer, P., and Trommsdorff, V., 1995. Serpentine stability to mantle depths and subduction-related magmatism. Science, 268: 858861.CrossRefGoogle ScholarPubMed
Ulmishek, G. F., and Klemme, H. D., 1990. Depositional Controls, Distribution, and Effectiveness of World’s Petroleum Source Rocks. US Geological Survey, Reston.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 (Boulder), 15(3): 216219.2.0.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, P., and Pelet, R., 1987. Extrapolation of the kinetics of oil and gas formation from laboratory to sedimentary basin. Nature, 327: 5254.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: Bjoroy, M., Albrecht, P., Cornford, C., et al. (Eds.), Advances in Organic Geochemistry 1981. Wiley, Chichester, pp. 129135.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. AAPG Bulletin, 74: 309335.Google Scholar
Unsworth, M., and Bedrosian, P. A., 2004. Electrical resistivity structure at the SAFOD site from magnetotelluric exploration. Geophysical Research Letters, 31: L12S05.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(4): 359362.2.3.CO;2>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(B1): 11311150.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(18): 30213024.CrossRefGoogle Scholar
Unsworth, M., Wenbo, W., Jones, A. G., et al., 2004. Crustal and upper mantle structure of northern Tibet imaged with magnetotelluric data. Journal of Geophysical Research: Solid Earth, 109(B2). DOI: 10.1029/2002JB002305CrossRefGoogle Scholar
Vågnes, E., 1997. Uplift at thermo-mechanically coupled ocean–continent transforms: modelled at the Senja Fracture Zone, southwestern Barents Sea. Geo-Marine Letters, 17: 100109.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. In: Application of Seismic Reflection Configuration to Stratigraphic Interpretation. AAPG, Washington, DC, pp. 99116.Google 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(3–4): 527544.CrossRefGoogle Scholar
Van Den Kerkhof, A. M., and Olsen, S. N., 1990. A natural example of superdense CO2 inclusions: microthermometry and Raman analysis. Geochimica et Cosmochimica Acta, 54: 885901.CrossRefGoogle Scholar
van der Beek, P., and Braun, J., 1998. Numerical modelling of landscape evolution on geological time-scales: a parameter analysis and comparison with the south-eastern highlands of Australia. Basin Research, 10: 4968.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(3–4): 417433.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(2): 406421.CrossRefGoogle Scholar
van der Beek, P., Summerfield, M. A., Braun, J., Brown, R. W., and Fleming, A., 2002. Modeling postbreakup landscape development and denudational history across the southeast African (Drakensberg Escarpment) margin. Journal of Geophysical Research. DOI: 10.1029/2001JB000744CrossRefGoogle Scholar
Van der Kamp, G., and Bachu, S., 1989. Use of dimensional analysis in the study of thermal effects of various hydrogeological regimes. In: Beck, A. E., Garven, G., and Stegena, L. (Eds.), Hydrogeological Regimes and Their Subsurface Thermal Effects. Wiley, New York, pp. 2328.Google Scholar
Van Krevelen, D. W., 1950. Graphical-statistical method for the study of structure and reaction processes of coal. Fuel, 29: 269284.Google Scholar
Van Schmus, W. R., 1984. Radioactivity properties of minerals and rocks. In: Carmichael, R. S. (Ed.), CRC Handbook of Physical Properties of Rocks. CRC Press, Boca Raton, pp. 281293.Google Scholar
van Wijk, J., Axen, G., and Abera, R., 2017. Initiation, evolution and extinction of pull-apart basins: implications for opening of the Gulf of California. Tectonophysics, 719–720: 3750.CrossRefGoogle Scholar
Vandenbroucke, M., 1993. Migration of hydrocarbons. In: Bordenave, M. L. (Ed.), Applied Petroleum Geochemistry. Technip, Paris, pp. 125148.Google Scholar
Vandenbroucke, M., and Largeau, C., 2007. Kerogen origin, evolution and structure. Organic Geochemistry, 38: 719833.CrossRefGoogle 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: 11051125.CrossRefGoogle Scholar
Vass, D., Nagy, A., Kohút, M., and Kraus, I., 1988. Devínska Nová Ves Formation: coarse-clastic sediments on the SE flank of the Vienna Basin. Mineralia Slovaca, 20(2): 97108.Google Scholar
Vavra, C. L., Kaldi, J. G., and Sneider, R. M., 1992. Geological applications of capillary pressure: a review. AAPG Bulletin, 76: 840850.Google Scholar
Veevers, J. J., and Johnstone, M. H., 1974. Comparative stratigraphy and structure of the Western Australian margin and the adjacent deep ocean floor. In: Initial Reports of the Deep Sea Drilling Project, Vol. 27. Texas A&M University, College Station, pp. 571585.Google Scholar
Vergés, J., Munoz, J. A., and Martinez, A., 1992. South Pyrenean fold and thrust belt: the role of foreland evaporitic levels in thrust geometry. In: McClay, K. R. (Ed.), Thrust Tectonics. Chapman and Hall, London, pp. 255264.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. In: Tankard, A. J., and Balkwill, H. R. (Eds.), Extensional Tectonics and Stratigraphy of the North Atlantic Margins. AAPG, Washington, DC, pp. 131147.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 (London), 337(6205): 354357.CrossRefGoogle Scholar
Villemin, T., Alvarez, F., and Angelier, J., 1986. The Rhinegraben: extension, subsidence and shoulder uplift. Tectonophysics, 128(1–2): 4759.CrossRefGoogle Scholar
Villinger, H., Grevemeyer, I., Kaul, N., Hauschild, J., and Pfender, M., 2002. Hydrothermal heat flux through aged oceanic crust: where does the heat escape? Earth and Planetary Science Letters, 202(1): 159170.CrossRefGoogle 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(9): 977980.CrossRefGoogle Scholar
Vinogradov, A., Tugarinov, A., Zhykov, C., et al., 1964. Geochronology of Indian Precambrian. In: XXII International Geological Congress, Rep. Pt, X, pp. 553–567.Google Scholar
Voight, D. E., and Brantley, S. L., 1991. Inclusions in synthetic quartz. Journal of Crystal Growth, 113: 527539.CrossRefGoogle Scholar
Von Herzen, R. P., 2004. Geothermal evidence for continuing hydrothermal circulation in older (>60 Ma) ocean crust. In: Davis, E. E., and Elderfield, H. (Eds.), Hydrogeology of the Oceanic Lithosphere. Cambridge University Press, Cambridge, pp. 414450.Google Scholar
von Stackelberg, U., 1979. Geologische Untersuchungen des 3W-australischen Kontinentalrandes Forschungsfahrt SO-B mit M.S. “Sonne”. Bundesanstaldt für Geowissenschaften und Fohstoffe, Fdrd. erungsvorhaben MF 0249 4, Archiv-Nr.: 82574, Tagebuch-Nr.: 6811/79Google Scholar
von Stackelberg, U., Exon, N. F., von Rad, U., et al., 1980. Geology of the Exmouth and Wallaby Plateaus off northwest Australia: sampling of seismic sequences. BMR Journal of Australian Geology and Geophysics, 5: 113140.Google Scholar
Vrolijk, P., Donelick, R. A., Queng, J., and Cloos, M., 1992. Testing models of fission track annealing in apatite in a simple thermal setting: Site 800. Leg 129 In: Proceedings of the Ocean Drilling Program Scientific Results, Volume 129. Texas A&M University, College Station, pp. 169176.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(2): 223230.CrossRefGoogle Scholar
Wade, J. A., and MacLean, B. C., 1990. The geology of the southeastern margin of Canada. InKeen, M. J. and Williams, G. L. (Eds.), Geology of the Continental Margin of Eastern Canada. Geological Society of America, Boulder, pp. 167238.CrossRefGoogle Scholar
Wagner, T., and Pletsch, T., 2001. No major thermal event on the mid-Cretaceous Côte d’Ivoire–Ghana Transform Margin. Terra Nova, 13(3): 165171.CrossRefGoogle Scholar
Walder, A. J., Platznert, I., and Freedman, P. A., 1993. Isotope ratio measurement of lead, neodymium and neodymium–samarium mixtures, hafnium and hafnium–lutetium mixtures with a double focusing multiple collector inductively coupled plasma mass spectrometer. Journal of Analytical Atomic Spectrometry, 8(1): 1923.CrossRefGoogle Scholar
Walker, R. J., Holdsworth, R. E., Imber, J., and Ellis, D., 2011. Onshore evidence for progressive changes in rifting directions during continental break-up in the NE Atlantic. Journal of the Geological Society, 168: 2748.CrossRefGoogle Scholar
Wallace, R. E., 1951. Geometry of shearing stress and relation to faulting. Journal of Geology, 59: 118130.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: 429435.CrossRefGoogle Scholar
Walsh, J. J., and Watterson, J., 1990. New methods of fault projection for coalmine planning. Proceedings of the Yorkshire Geological Society, 48: 209219.CrossRefGoogle 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
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(1–2): 127140.CrossRefGoogle Scholar
Wang, K., and Davis, E. E., 1992. Thermal effects of marine sedimentation in hydrothermally active areas. Geophysical Journal International, 110: 7078.CrossRefGoogle Scholar
Wang, Y., Mooney, W. D., Yuan, X., and Okaya, N., 2013. Crustal structure of the northeastern Tibetan Plateau from the southern Tarim Basin to the Sichuan Basin, China. Tectonophysics, 584: 191208.CrossRefGoogle Scholar
Wannamaker, P. E., Jiracek, G. R., Stodt, J. A., et al., 2002. Fluid generation and pathways beneath an active compressional orogen, the New Zealand Southern Alps, inferred from magnetotelluric data. Journal of Geophysical Research (Solid Earth), 107(B6). DOI: 10.1029/2001JB000186Google Scholar
Wannamaker, P. E., Caldwell, T. G., Jiracek, G. R., et al., 2009. Fluid and deformation regime of an advancing subduction system at Marlborough, New Zealand. Nature, 460(7256): 733736.CrossRefGoogle ScholarPubMed
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: Meissner, R., Brown, L., Duerbaum, H.-J., et al. (Eds.), Continental Lithosphere: Deep Seismic Reflections. AGU, Washington, DC, pp. 275289.CrossRefGoogle Scholar
Waples, W., 1980. Time and temperature in petroleum formation: application of Lopatin’s method to petroleum exploration. AAPG Bulletin, 64: 916926.Google Scholar
Waples, D. W., 1994a. Maturity modeling: thermal indicators, hydrocarbon generation, and oil cracking. In: Magoon, L. B., and Dow, W. E. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 285306.Google Scholar
Waples, D. W., 1994b. Modeling of sedimentary basins and petroleum systems. In: Magoon, L. B., and Dow, W. E. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 307322.Google Scholar
Waples, D. W., Kamata, H., and Suizu, M., 1992a. The art of maturity modeling, Part 1: finding a satisfactory geologic model. AAPG 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. AAPG Bulletin, 76: 4766.Google Scholar
Ward, P. L., 1971. New Interpretation of the geology of Iceland. Geological Society of America Bulletin. 82(11): 29913012.CrossRefGoogle 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(24): 53715377.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(B12): 59896004.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(3–4): 239252.CrossRefGoogle Scholar
Watts, A. B., and Thorne, J., 1984. Tectonics, global changes in sea level and their relationship to stratigraphical sequences at the US Atlantic continental margin. Marine and Petroleum Geology, 1(4): 319339.CrossRefGoogle 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: 249281.Google 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
Weast, R. C., 1974. CRC Handbook of Chemistry and Physics. CRC Press, Boca Raton.Google Scholar
Weber, M. E., Wiedicke, M. H., Kudrass, H. R., Hübscher, C., and Erlenkeuser, H., 1997. Active growth of the Bengal Fan during sea-level rise and highstand. Geology, 25: 315318.2.3.CO;2>CrossRefGoogle Scholar
Weber, M., Abu-Ayyash, K., Abueladas, A., et al., 2009. Anatomy of the Dead Sea Transform from lithospheric to microscopic scale. Reviews of Geophysics, 47:(2). DOI: 10.1029/2008RG000264CrossRefGoogle Scholar
Wech, A. G., Boese, C. M., Stern, T. A., and Townend, J., 2012. Tectonic tremor and deep slow slip on the Alpine Fault. Geophysical Research Letters, 39(10). DOI: 10.1029/2012GL051751CrossRefGoogle Scholar
Weinstein, Y., 2012. Transform faults as lithospheric boundaries, an example from the Dead Sea Transform. Journal of Geodynamics, 54: 2128CrossRefGoogle 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(B10): 1391913950.CrossRefGoogle Scholar
Weissel, J. K., and Karner, G. D., 1994, Flexural uplift of rift flanks due to mechanical unloading of the lithosphere during extension. Journal of Geophysical Research, 94: 1391913950.CrossRefGoogle Scholar
Weissenbäck, M., 1995. Sedimentological model of the Lower–Middle Miocene (Karpatian–Badenian) of the central Vienna Basin (in German). PhD thesis. University of Vienna, Vienna.CrossRefGoogle Scholar
Welhan, J. A., 1988. Origins of methane in hydrothermal systems. Chemical Geology, 71: 183198.CrossRefGoogle Scholar
Welhan, J. A., and Craig, H., 1983. Methane, hydrogen and helium in hydrothermal fluids of 21° N on the East Pacific Rise. In: Rona, P. A., Bostrom, K., Laubier, L., and Smith, K. L., Jr. (Eds.), Hydrothermal Processes at Seafloor Spreading Centers. Plenum, New York, pp. 391409.CrossRefGoogle Scholar
Wellman, H. W., 1953. Data for the study of recent and late Pleistocene faulting in the South Island of New Zealand. New Zealand Journal of Science and Technology, 34B: 270288.Google Scholar
Welsink, H. J., Dwyer, J. D., and Knight, R. J., 1989. Tectono-stratigraphy of the passive margin off Nova Scotia. In: Tankard, A. J., and Balkwill, H. R. (Eds.), Extensional Tectonics and Stratigraphy of the North Atlantic Margins. AAPG, Washington, DC, pp. 215231.Google Scholar
Wenk, H.-R., Johnson, L. R., and Ratschbacher, L., 2000. Pseudotachylites in the Eastern Peninsular Ranges of California. Tectonophysics, 321: 253277.CrossRefGoogle Scholar
Wernicke, B., 1990. The fluid crustal layer and its implications for continental dynamics. In: Salisbury, M. H. and Fountain, D. M. (Eds.), Exposed Cross-Sections of the Continental Crust. Springer, Dordrecht.Google Scholar
Wesnousky, S. G., 2006. Predicting the endpoints of earthquake ruptures. Nature, 444: 358360.CrossRefGoogle ScholarPubMed
Wessely, G., 1993. Der Untergrund des Wiener Beckens. In: Brix, F., and Schultz, O. (Eds.), Erdol und Erdgas in Osterreich. Naturhistorisches NYseum und F. Berger, Horn, pp. 249280.Google Scholar
Wessely, G., 1999. Oil and gas occurrences in the Vienna Basin. In: FOREGS’99, Dachstein-Hallstatt-Salzkammergut Region, abstract volume, pp. 20–21.Google Scholar
Wessely, G., 2000. Sediments of the Vienna Basin and their Alpine and subalpine basement [in German]. Mitteilungen der Gesellschaft für Geologie und Bergbaustudenten Österreichs, 44: 191214.Google Scholar
Wessely, G., 2006. Geology of Austrian states: Lower Austria [in German]. Geologischen Bundesanstalt, Wien.Google Scholar
Wesson, R. L., 1988. Dynamics of fault creep. Journal of Geophysical Research, 93: 89298951.CrossRefGoogle Scholar
West, J., and Lewis, H., 1982. Structure and palinspastic reconstruction of the Absaroka Thrust, Anschutz Ranch area, Utah and Wyoming. In: Powers, R. B. (Ed.), Geologic Studies of the Cordilleran Thrust Belt. Rocky Mountain Association of Geologists, Boulder, pp. 633639.Google Scholar
Westaway, R., 1994. Evidence for dynamic coupling of surface processes with isostatic compensation in the lower crust during active extension of western Turkey. Journal of Geophysical Research, 99. 2020320223.CrossRefGoogle Scholar
Westwood 2019. Exploration Potential in the Tano Basin. Westwood Wildcat report. February 2019.Google 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 (Boulder), 16(3): 250253.2.3.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: 841854.CrossRefGoogle 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, Solid Earth, 94(B6): 76857729.CrossRefGoogle Scholar
White, R. S., Detrick, R. S., Sinha, M. C., and Cormier, M. H., 1984. Anomalous crustal seismic structure of oceanic fracture zones. Geophysical Journal of the Royal Astronomic Society, 79: 779798.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(B13): 1968319715.CrossRefGoogle Scholar
Whiticar, M. J., 1994. Correlation of natural gases with their sources. In: Magoon, L. B., and Dow, W. E. (Eds.), The Petroleum System: From Source to Trap. AAPG, Washington, DC, pp. 261283.Google Scholar
Whitmarsh, R. B., White, R. S., Sibuet, J.-C., et al., 1996. The ocean–continent boundary off the western continental margin of Iberia: crustal structure west of Galicia Bank. Journal of Geophysical Research, 101(B12): 2829128314.CrossRefGoogle 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(5): 919942.CrossRefGoogle Scholar
Widdowson, M., 1997. Tertiary palaeosurfaces of the SW Deccan, Western India: implications for passive margin uplift. In: Widdowson, M. (Ed.), Palaeosurfaces: Recognition, Reconstruction, and Palaeoenvironmental Interpretation. Geological Society of London, London, pp. 221248.Google Scholar
Wiens, D. A., and Stein, S., 1983. Age dependence of oceanic intraplate seismicity and implications for lithospheric evolution. Journal of Geophysical Research, 88: 64556468.Google Scholar
Wilcock, W. S. D., and Delaney, J. R., 1996. Mid ocean ridge sulfide deposits: evidence for heat extraction from magma chambers or cracking fronts? Earth and Planetary Science Letters, 145: 4964.CrossRefGoogle Scholar
Wilcock, W. S. D., Solomon, S. C., Purdy, G. M., Toomey, D. R., and Dougherty, M. E., 1992. Images of seismic attenuation at 9° 30′N on the East Pacific Rise: implications for the distribution of melt within the crust. AGU Eos Transactions, 73: 290.Google Scholar
Wilcox, R. E., Harding, T. P., and Seely, D. R., 1973. Basic wrench tectonics. AAPG Bulletin, 57: 7496.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: 283294.CrossRefGoogle Scholar
Wilkinson, W. B., and Shipley, E. L., 1972. Vertical and horizontal laboratory permeability measurements in clay soils. In: Bear, J., and Yavuz Corapcioglu, M. (Eds.), Fundamentals of Transport Through Porous Media. Elsevier, Amsterdam, pp. 285298.CrossRefGoogle Scholar
Wilks, K. R., and Carter, N. L., 1990. Rheology of some continental lower crustal rocks. Tectonophysics, 182: 5777.CrossRefGoogle Scholar
Willemse, E. J. M., Peacock, D. C. P., and Aydin, A., 1997. Nucleation and growth of strike-slip faults in limestones from Somerset, U. K. Journal of Structural Geology, 19: 14611477.CrossRefGoogle Scholar
Willett, S. D., 1997. Inverse modeling of annealing of fission tracks in apatite 1: a controlled random search method. American Journal of Science, 297: 939969.CrossRefGoogle Scholar
Williams, C. A., and Richardson, R. M., 1991. A rheologically layered three-dimensional model of the San Andreas Fault in central and southern California. Journal of Geophysical Research, 96(B10): 1659716623CrossRefGoogle Scholar
Williams, H. H., and Eubank, R. T., 1995. Hydrocarbon habitat in the rift graben of the central Sumatra Basin, Indonesia. In: Lambiase, J. J. (Ed.), Hydrocarbon Habitat in Rift Basin. Geological Society of London, London, pp. 331371.Google Scholar
Wilson, D. S., Teagle, D. A. H., Alt, J., et al., 2005a. Drilling to gabbro in intact oceanic crust. Science, 312: 10161020.CrossRefGoogle Scholar
Wilson, D. S., Teagle, D. A. H., Alt, J., et al., 2005b. Drilling a Complete in Situ Section of Upper Oceanic Crust Formed at Superfast Spreading Rate: Hole 1256D. EGU General Assembly, Vienna.Google Scholar
Wilson, J. T. 1965. A new class of faults and their bearing on continental drift. Nature, 207: 343347.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(1–2): 1940.CrossRefGoogle Scholar
Wilson, W. R., 1980. Thermal studies in a geothermal area. PhD thesis. University of Utah, Salt Lake City.Google Scholar
Winker, C. D., 1987, Neogene stratigraphy of the Fish Creek-Vallecito section, Southern California: implications for early history of the northern Gulf of California and Colorado Delta. Doctoral thesis. University of Arizona, Tucson.Google Scholar
Winker, C. D., and Kidwell, S. M., 1986, Paleocurrent evidence for lateral displacement of the Pliocene Colorado River delta by the San Andreas fault system, southeastern California. Geology, 14(9): 788791.2.0.CO;2>CrossRefGoogle 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(2): 390405.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. AAPG Bulletin, 82: 817835.Google Scholar
Wittlinger, G., Tapponnier, P., Poupinet, G., et al., 1998. Tomographic evidence for localized lithospheric shear along the Altyn Tagh Fault. Science,. 282(5386): 7476.CrossRefGoogle ScholarPubMed
Wittlinger, G., Vergne, J., Tapponnier, P., et al., 2004. Teleseismic imaging of subducting lithosphere and Moho offsets beneath western Tibet. Earth and Planetary Science Letters, 221(1–4): 117130.CrossRefGoogle Scholar
Wojtal, S., and Mitra, G., 1986. Strain hardening and strain softening in fault zones from foreland thrusts. Geological Society of America Bulletin, 97: 674687.2.0.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(21): 42314240.CrossRefGoogle Scholar
Wolfson Schwehr, M., 2015. The relationship between oceanic transform fault segmentation, seismicity, and thermal structure. Doctoral dissertation. University of New Hampshire, Durham.Google Scholar
Woodcock, N. H., 1986. The role of strike-slip fault systems at plate boundaries. Royal Society of London Philosophical Transactions, A., 317: 1329.Google Scholar
Woodcock, N. H., and Fischer, M., 1986. Strike-slip duplexes. Journal of Structural Geology, 8: 725735.CrossRefGoogle Scholar
Worman, S. L., Pratson, L. F., Karson, J. A., and Klein, E. M., 2016. Global rate and distribution of H2 gas produced by serpentinization within oceanic lithosphere. Geophysical Research Letters, 43(12): 64356443.CrossRefGoogle Scholar
Wright, J. B., 1976. Fracture systems in Nigeria and initiation of fracture zones in the South Atlantic. Tectonophysics, 34: T43T47.CrossRefGoogle Scholar
Wright, L. D., 1977. Sediment transport and deposition at river mouths: a synthesis. AAPG Bulletin, 88: 857868.Google Scholar
Wright, T. L., 1991. Structural geology and tectonic evolution of the Los Angeles basin, California. In: Biddle, K. T. (Ed.), Active Margin Basins. AAPG, Washington, DC, pp. 35134.Google Scholar
Wright, T., Parsons, B., and Fielding, E., 2001. Measurement of interseismic strain accumulation across the North Anatolian Fault by satellite radar interferometry Geophysical Research Letters, 28(10): 21172120.CrossRefGoogle Scholar
Wygrala, B. P., 1989. Integrated Study of an Oil Field in the Southern Po Basin, Northern Italy. University of Koln, Koln.Google Scholar
Xiao, Q., Zhao, G., and Dong, Z., 2011. Electrical resistivity structure at the northern margin of the Tibetan Plateau and tectonic implications. Journal of Geophysical Research: Solid Earth, 116(B12). DOI: 10.1029/2010JB008163CrossRefGoogle Scholar
Yalcin, M. N., Littke, R., and Sachsenhofer, R. F., 1997. Thermal history of sedimentary basins. In: Welte, D. H., Horsfield, B., and Baker, D. R. (Eds.), Petroleum and Basin Evolution: Insights from Petroleum Geochemistry, Geology and Basin Modeling. Springer, Berlin, pp. 71167.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: 627649.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: 12751292.CrossRefGoogle Scholar
Yardley, B. W. D., 1986. Is there water in the deep continental crust? Nature, 323: 111.CrossRefGoogle Scholar
Yeats, R. S., Khan, S. H., and Akhtar, M., 1984. Late Quaternary deformation of the Salt Range of Pakistan. Geological Society of America Bulletin, 95: 958966.2.0.CO;2>CrossRefGoogle Scholar
Yerkes, R., McCulloch, T., Schoellhamer, J., and Vedder, J., 1965. Geology of the Los Angeles Basin, California: an introduction. US Geological Survey Professional Paper 420-A.CrossRefGoogle Scholar
Yielding, G., Jackson, J. A., King, G. C. P., et al., 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
Yilmaz, Y., 2011. Geological and structural framework of Turkey. Presented at the Central-Eastern Europe and Caspian Scout Group Meeting, Belek, October 7–8.Google Scholar
Yin, A., Rumelhart, P. E., Butler, R., et al., 2002. Tectonic history of the Altyn Tagh fault system in northern Tibet inferred from Cenozoic sedimentation. Geological Society of America Bulletin, 114(10): 12571295.2.0.CO;2>CrossRefGoogle Scholar
Younes, A. I., 1996. Fracture distribution on faulted basement blocks, Gulf of Suez, Egypt: reservoir characterization and tectonic implications. Doctoral thesis. Pennsylvania State University at University Park, University Park.Google Scholar
Younes, A. I., and McClay, K. R., 2002. Development of accommodation zones in the Gulf of Suez-Red Sea Rift, Egypt. AAPG Bulletin, 86(6): 10031026.Google Scholar
Yu, W., and Sephrnoori, K., 2018. CO2 injection for enhanced oil recovery in tight oil reservoirs. Shale Gas and Tight Oil Reservoir Simulation, 2018: 333376.CrossRefGoogle Scholar
Yue, Y., and Liou, J. G., 1999. Two-stage evolution model for the Altyn Tagh fault, China. Geology,. 27(3): 227230.2.3.CO;2>CrossRefGoogle Scholar
Zak, I., and Freund, R., 1981. Asymmetry and basin migration in the Dead Sea Rift. Tectonophysics, 80(1–4, Special issue): 2738.CrossRefGoogle Scholar
Zalán, P. V., Nelson, E. P., Warme, J. E., and Davis, T., 1985. The Piaui Basin, rifting and wrenching in an equatorial Atlantic transform basin. In: Biddle, K. T., and Christie-Blick, N. (Eds,), Strike-Slip Deformation, Basin Formation, and Sedimentation. SEPM. Tulsa, pp. 177192.CrossRefGoogle Scholar
Zanella, E., and Coward, M. P., 2003. Structural framework. In: Evans, D., Graham, C., Armour, A., and Bathurst, P. (Eds.), The Millennium Atlas: Petroleum Geology of the Central and Northern North Sea. Geological Society of London, London, pp. 4559.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(10): 28652868.CrossRefGoogle Scholar
Zencher, F., Bonafede, M., and Stefansson, R., 2006. Near-lithostatic pore pressure at seismogenic depths: a thermo-poro-elastic model. Geophysics Journal International, 166: 13181334.CrossRefGoogle Scholar
Zhang, J.-W., and Nancollas, G. H., 1990. Mechanisms of growth and dissolution of sparingly soluble salts. In: Hochella, M. F., and White, A. F. (Eds.), Mineral-Water Interface Geochemistry. De Gruyter, Berlin, pp. 365396.CrossRefGoogle Scholar
Zhao, G., and Johnson, A. M., 1991. Sequential and incremental formation of conjugate faults. Journal of Structural Geology, 13: 887896.CrossRefGoogle 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). In: Geodynamics of rifting; Volume I, Case History Studies on Rifts; Europe and Asia. Geodynamics of Rifting Symposium, Glion-sur-Montreux, Switzerland, pp. 159171.Google Scholar
Zimmerman, R. W., Somerton, W. H., and King, M. S., 1986. Compressibility of porous rocks. Journal of Geophysical Research, 91: 1276512777.CrossRefGoogle Scholar
Znidarcic, D., and Aiban, S. A., 1988. Comment on Al-Tabbaa and Wood “Some measurements of the permeability of kaolin.” Geotechnique, 38: 453454.Google Scholar
Zoetemeijer, R., 1993. Tectonic Modeling of Foreland Basins: Thin-Skinned Thrusting, Syntectonic Sedimentation and Lithospheric Flexure. Vrije University, Amsterdam.Google Scholar
Zoth, G., 1979. Temperaturmessungen in der Bohrung Hanigsen sowie Warmeleitfahigeitbestimmungen von Gesteinproben. Report BGR/NLfB Hannover, Archive No. 81 828.Google Scholar
Zoth, G., and Hänel, R., 1988. Appendix. In: Hänel, R., Rybach, L., and Stegena, L. (Eds.), Handbook of Terrestrial Heat Flow Density Determination. Kluwer Academic, Norwell, pp. 449466.CrossRefGoogle Scholar
Zuber, M. T., and Parmentier, E. M., 1986. Lithospheric necking: a dynamic model for rift morphology. Earth and Planetary Science Letters, 77(3–4): 373383.CrossRefGoogle Scholar
Zuehlsdorf, L., Hutnak, M., Fisher, A. T., et al., 2005. Data report: Site surveys related to IODP Expedition 301: ImageFlux (SO149) and RetroFlux (TN116) Expeditions and earlier studies. In: Proceedings of Integrated Ocean Drilling Program, Volume 301. Texas A&M University, College Station.Google 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? Paper presented at the Conference on Basin Modeling, London.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure [email protected] 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.

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.

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.

Available formats
×